Next Article in Journal
Release Behavior of Benzimidazole-Intercalated α-Zirconium Phosphate as a Latent Thermal Initiator in the Reaction of Epoxy Resin
Previous Article in Journal
Optimization of Biodiesel Production from Waste Cooking Oil Using S–TiO2/SBA-15 Heterogeneous Acid Catalyst
Previous Article in Special Issue
Steam Reforming of Methanol over Nanostructured Pt/TiO2 and Pt/CeO2 Catalysts for Fuel Cell Applications
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Dry Reforming of Propane over γ-Al2O3 and Nickel Foam Supported Novel SrNiO3 Perovskite Catalyst

by
Sudhakaran M.S.P
,
Md. Mokter Hossain
,
Gnanaselvan Gnanasekaran
and
Young Sun Mok
*
Department of Chemical and Biological Engineering, Jeju National University, Jeju 63243, Korea
*
Author to whom correspondence should be addressed.
Catalysts 2019, 9(1), 68; https://doi.org/10.3390/catal9010068
Submission received: 28 November 2018 / Revised: 5 January 2019 / Accepted: 7 January 2019 / Published: 10 January 2019
(This article belongs to the Special Issue Catalysis in Steam Reforming)

Abstract

:
The SrNiO3 perovskite catalyst was synthesized by the citrate sol-gel method and supported on γ-Al2O3 and Nickel foam, which was used to produce syngas (CO and H2) via dry reforming of propane (DRP). Several techniques characterized the physicochemical properties of the fresh and spent perovskite catalyst. The X-ray diffractograms (XRD) characterization confirmed the formation of the perovskite compound. Before the catalytic activity test, SrNiO3 perovskite catalyst was reduced in the H2 atmosphere. Results indicated that the H2 reduction slightly increased the activity of the SrNiO3 perovskite catalyst. The catalytic activity was examined for the CO2/C3H8 ratio of 3 and reaction temperatures in the range of 550 °C–700 °C. The results from the catalytic study achieved 88% conversion of C3H8 and 66% conversion of CO2 with SrNiO3/NiF at 700 °C. Also, syngas with a maximum concentration of 21 vol.% of CO and 29 vol.% of H2 was produced from the DRP. The strong basicity of SrNiO3 perovskite enhanced the CO selectivity, resulting in minimal carbon formation. Post reaction catalyst characterization showed the presence of carbon deposition which could have originated from propane decomposition.

1. Introduction

The conventional production of synthesis gas (syngas) by methane steam reforming (1) regularly produces a product with higher H2/CO values greater than 3 [1,2]. Recently, the attention has been drawn to the conversion of light hydrocarbons with carbon dioxide into the valuable product (syngas) by catalytic reactions, which is known as dry reforming (2) [3]. The result of a product is with a lower H2/CO ratio ≤2. This ratio is more suitable for the synthesis of liquid fuels, and chemicals such as olefins, methanol synthesis and Fisher–Tropsch synthesis [4,5,6,7,8]. Propane is a by-product of natural gas and majorly produced in a variety of petroleum refining operations. Primarily, it readily activates at a lower reaction temperature than methane [9].
CH 4 + H 2 O = CO + 3 H 2 ( Δ H ° 298 = 222.4   kJmol 1 )
C 3 H 8 + 3 CO 2 = 6 C O + 4 H 2 ( Δ H ° 298 = 644.1   kJmol 1 )
However, the steam and dry are an endothermic reaction, which requires high energy to occur reaction. Numerous catalysts have been described for each of the processes mentioned above for various hydrocarbons [10,11,12]. Noble metal catalysts show superior performance regarding the activity and the durability than non-noble metal catalysts. The Ni-based catalysts are attractive and promising due to their high activity, low cost, and abundant availability [13]. It has been widely studied with different support materials such as Al2O3, SiO2, CeO2, and SiC, and among them, Al2O3 is the most widely used as catalytic support in dry reforming [14,15]. However, the main limitations of Ni catalyst are particle sintering and coke formation. Therefore, the development of a suitable catalyst with high activity and anti-coking capability is essential for dry reforming and the selection of suitable support may be a pivotal step to prepare the catalyst with high performance. The typical perovskites are ABO3 type crystal structured materials with rare earth or alkaline earth metal at the A-site and a transition metal at the B-site [16,17]. These possess some exciting chemical and physical properties, e.g., high thermal stability, the excellent reactivity of lattice oxygen, low cost and abundant resources [16]. A-site replacement with alkaline earth metal is expected to enhance the carbon susceptible and thermal stability, and the B-site alteration is expected to increase the activity of the catalyst [18,19]. In last decade, many researchers have been widely reported LaNiO3 [20,21,22], LaFeO3 [23], SmCoO3 [24] and also partial substitution of B- site perovskite catalysts show the higher activity, stability, and resistance to coke formation even at elevated temperature. The perovskite catalysts increase the syngas production in dry reforming reaction [25,26,27]. Metal foams are highly porous with open-celled materials, which pack in a three-dimensional network. Metal foams have promising support due to their low density, mechanical strength, and high thermal conductivity. It has been reported in heterogeneous catalysis include dry reforming and catalytic combustion of CH4 [28,29].
Hence, this study has focused on low-cost and high basicity material with key features such as high catalytic activity and coke resistance. The objectives of this study are to synthesize of novel SrNiO3 perovskite as a catalyst to produce syngas via dry reforming of propane. The physicochemical properties of catalysts characterized with various techniques and the catalytic activity investigated in typical experimental gas reactor setup. The catalytic conversion of reactants and enhancement of production of syngas was carried out with the effects of reaction temperature. The effects of support materials were examined with SrNiO3 perovskite by a different form of support materials such as γ-Al2O3 and nickel foam. Also, the effect of H2 reduction was studied by comparing the conversion of the reduced and unreduced SrNiO3 perovskite catalyst before the propane dry reforming reaction.

2. Result and Discussion

2.1. Materials Characterization

XRD patterns of the calcined perovskite and reduced perovskite are shown in Figure 1. The sample was calcined at 900 °C and reduced by hydrogen at 700 °C. The fresh catalyst shows the real pattern of SrNiO3 perovskite state and the somewhat slight secondary peak of NiO also found. As seen in Figure 1, the highest sharp peaks found at 2θ with a value of 32.52°. It shows the formation of the SrNiO3 crystalline phase. The other peaks at 2θ values of 19.24°, 29.36°, 39.04°, 41.08°, 44.04°, 49.84°, 56.24°, 58.24°, 68.4°, and 76.24° also correspond to SrNiO3 species. Beside SrNiO3 phase, some peaks were evolved at 37.04°, 43.08°, and 62.76°, which belong to the NiO crystalline phase. All Ni precursors were not used to form SrNiO3. The desired SrNiO3 catalyst was synthesis successfully, and the diffraction pattern confirms formation of the hexagonal perovskite structure, matched with standard pattern (JCPDS card no: 33-1347) [30]. The post-reduction catalyst clearly shows the diffraction peaks corresponding to Ni0 and SrO. The peaks found at 2θ values of 44.4°, 51.8°, and 76.3° correspond to metallic Ni and peaks were evolved at 31.4°, 36.3°, and 62.2° correspond to the SrO crystalline phase. Those peaks were matched with the standard patterns (JCPDS card no: 65-2865 and JCPDS card no: 27-1304), respectively [31,32]. Generally, the high-temperature reduction of perovskites leads to the formation of nanoparticles of B site metal (e.g., Ni) dispersed on the oxide formed by the A site metal (e.g., Sr) [33,34]. The XRD diffraction of γ-Al2O3 and NiF supported catalysts are shown in Figure 1b,c. The results indicate the existence of SrNiO3, NiO, Ni, SrO peaks in both supported catalysts. The peaks were observed at 2θ values of 45.3°, and 66.7° belong to the γ-Al2O3 (JCPDS card no: 02-1420) [35] crystalline phase and peaks were found at 2θ values of 44.4°, 51.8°, and 76.3° correspond to metallic Ni [36].
The N2 physisorption isotherm of the SrNiO3 perovskite is shown in Figure 2. The Brunauer–Emmett–Teller (BET) surface area is 3.3 m2 g−1, which is agreed with perovskite materials as reported in the literature [34]. The BET surface area of the catalysts with different supports reported in Table 1, the surface area of reduced catalyst was increased in both support materials compared to the fresh catalyst SrNiO3. It implies that the breakdown of the perovskite structure, leading to formation of Ni/SrO, resulted in the generation of porosity to some extent, which also explains the slight enhancement of surface area of this bulk perovskite after reduction. The surface morphology of the SrNiO3 perovskite catalyst is shown in Figure 3. The high-resolution SEM image reveals the porous and flake structure of the material. From low resolution and high-resolution SEM images, it shows the less impact of particle agglomeration due to the citrate sol-gel method [37].
Temperature programmed reduction (H2-TPR) analysis was determined by the reduction behavior of the fresh catalyst and reduced catalyst. Figure 4 shows the H2-TPR profile of the fresh catalyst and reduced SrNiO3 perovskite catalyst. The two intense peaks were observed at lower and higher temperatures (600 K and 987 K, respectively) of the SrNiO3 fresh catalyst. The reduced SrNiO3 exhibited one broad peak at higher temperature 1096 K. The fresh calcined SrNiO3 catalyst reduction process occurred by the following steps. The first step at lower temperature 600 K is attributed to Ni3+ reduction to Ni2+ and, the second peak at 987 K corresponds to the complete reduction of the perovskite to form Ni0 (Equation (3)) [27].
SrNiO 3 + H 2 SrO + Ni 0 + 2 H 2 O
The reduced SrNiO3 perovskite shows the single reduction peak at high temperature, which means the complete reduction of SrNiO3 phase to SrO. The perovskite compound’s reduction occurred at high temperature, which agreed with the previous report [38]. The reducibility of fresh and reduced SrNiO3 perovskite supported with γ-Al2O3 and nickel foam is presented in Figure 4b. The reduction peaks were observed in the supported catalysts significantly matched with unsupported catalysts. The peaks were evolved at lower region 624 °K, 585 °K, and at higher region 948 °K, 970 °K of γ-Al2O3 and NiF supported catalysts confirmed the reduction properties of fresh SrNiO3 perovskite. Those peaks were shifted to the lower temperature than an unsupported catalyst, which shows the interaction of support. The peak was found at around 1054 °K of both supported catalysts, which matched with the reduced SrNiO3 perovskite. The reducibility of the both supported catalysts showed the metal support interaction enhance the accessibility of active species to the reactant.
Figure 5 shows the chemisorption characters of SrNiO3 perovskite by H2 temperature programmed desorption method. As seen in Figure 5, the two distinct peaks were observed at 373 °C, and 426 °C. The first peaks attributed to the H2 molecules desorbed from the metal particles, and they represent the outer positioned of Ni atoms in the perovskite lattice [39,40]. The second sharp peak attributed to the H2 dissociated in the subsurface layers of the perovskite lattice. This phenomenon is known as H2 spillover [41,42,43]. It reveals the high energy requires to unbind the H2 molecules and strong enough of chemisorption. From the results, the amount of H2-chemisorbed molecules was estimated under the curve. The H2 consumption amount of the SrNiO3 catalyst was 1.89 × 1020 g c a t 1 .

2.2. Catalytic Activity

The activity of perovskite catalysts has been studied in two forms without reduction and reduced with H2 before reaction. The reduced catalysts have shown the significant activity than a fresh catalyst. The reduction process happened to reduce the transition metal phase of the catalyst to a metallic phase, which is located at the active site of the catalyst [44]. It has been reported that reduction enhances the metal dispersion, thereby providing an adequate platform for propane dry reforming reaction to occur. Reduction step has been reported to improve the reactant conversion as well as syngas formation [45]. In this experiment, catalytic activity was examined to determine reactants conversion for the reduced and unreduced catalyst with different supports. As seen in Figure 6, the catalytic activity of SrNiO3/γ-Al2O3 (F and R) and SrNiO3/NiF (F and R) catalysts showed that increases conversion of C3H8 and CO2 with increasing reaction temperature ranged from 550 °C to 700 °C. In terms of C3H8 conversion, the SrNiO3/γ-Al2O3(R) showed that the higher conversion among all other catalysts. It showed a maximum 94% of C3H8 conversion at 700 °C and minimum 24% conversion at 550 °C, which is higher than the SrNiO3/γ-Al2O3(F) catalyst. The C3H8 conversion of the Ni supported catalysts shows the increasing conversion with increases reaction temperature. The C3H8 conversion of SrNiO3/NiF (F and R) catalyst is 81% and 87% at 700 °C, respectively. The unreduced catalysts showed the lower C3H8 conversion with compare to the reduced catalysts. These indicated that in-situ catalyst reduction occurred by H2 arising from the C3H8 cracking to produce H2 and carbon [46].
Interestingly, the CO2 conversion of the SrNiO3/NiF (F and R) catalysts showed the significant conversion of CO2 than the SrNiO3/γ-Al2O3 (F and R) catalyst. Among all the catalysts, SrNiO3/NiF(R) showed the 69% of CO2 conversion, which was 13% higher than the SrNiO3/γ-Al2O3(R). It suggested that the SrNiO3/NiF(R) leads to the dry reforming reaction even at increasing temperature while the SrNiO3/γ-Al2O3 (F and R) has no significant ability to convert CO2 to CO since it leads the C3H8 cracking reaction and produces more carbon on the surface.
The catalytic activity of bare supports is shown in Figure 7. As seen in Figure 7, both bare supports showed low catalytic activities even at 700 °C compared to SrNiO3 loaded catalysts. The conversion of C3H8 of bare γ-Al2O3 shows slightly higher conversion at low temperature than bare NiF and the C3H8 conversion is almost the same (16%) at a higher temperature in both supports. The supports majorly contribute the C3H8 cracking reaction instead of DRP. Moreover, the CO2 conversion of both supports shows the poor activity, which did not reduce to CO. The CO2 conversion of γ-Al2O3 support showed the higher conversion (5%) at low temperatures than C3H8 conversion due to carbon formation from the C3H8 cracking reaction. Hence, the SrNiO3 loaded catalysts show excellent activity compared to the bare supports, which indicates that the supports improve the dispersivity of active catalyst to enhance the interaction of the reactants with active sites.
The outlet concentration of H2 and CO of the catalysts is shown in Figure 8. As seen in Figure 8, the H2 production increases with increasing temperature for all catalyst. The SrNiO3/γ-Al2O3 (F and R) catalysts showed that the higher H2 production compares with SrNiO3/NiF (F and R) catalyst. The maximum H2 and CO production were observed by SrNiO3/γ-Al2O3 (R) is 22% and 20% at 700 °C respectively. It suggested that the γ-Al2O3 supported catalyst primarily enhanced the C3H8 cracking reaction instead of DRP, which leads to the coke generation. Notably, the SrNiO3/NiF (F and R) catalysts have shown 20%, 21%, and 27%, 29% of H2 and CO production, respectively. The CO2 conversion always showed lower than the C3H8 conversion, and the H2/CO ratio was significantly 0.7 for the SrNiO3/NiF (F and R). This observation could be explained by a low amount of coke formation on the surface of the catalyst, and the DRP reaction took place in thermodynamic equilibrium.
As shown in Figure 9, we could compare the SrNiO3 activity in two different supports materials at the isothermal condition. From the results, the NiF foam-supported catalyst showed superior catalytic activity. It could be active in two forms, and the results were almost the same. These suggested that the NiF could be facilitated the significant interaction of the catalytic surface and efficiently enhance the syngas production, which closes to the stoichiometric reaction. As a result of Figure 9, the SrNiO3/NiF(R) revealed the significant conversion of C3H8 and CO2 among all. Regarding CO and H2 selectivity were 96% and 64%, respectively. Moreover, the H2/CO ratio of the catalyst revealed that reaction mechanism, the SrNiO3/γ-Al2O3(F and R) catalysts showed higher values 1.8 and 1.0, respectively, which suggested that the CO selectivity was less than the H2 selectivity. The CO2 conversion was lower than C3H8 conversion due to the additional CO2 produced by water gas shift reaction (WGSR) (Equation (4)). The WGSR was confirmed by the value of H2/CO ratio higher than 1 [34]. The CO selectivity severely affected due to the predominantly produced CO molecules could react with H2O to produce CO2 and H2. Notably, the SrNiO3/NiF (F and R) catalysts have shown the significant conversion of C3H8, CO2 and higher selectivity of CO than the selectivity of H2. The H2/CO ratio is 0.7, which closes to the stoichiometric reaction value of DRP (Equation (2)). These indicated that the NiF supported catalysts performed excellently in terms of activity, which aid the DRP reaction and also significantly favor the side reaction reverse water gas shift reaction (Equation (5)). The H2/CO ratio would be higher than 1 since this reaction did not occur [47,48].
CO + H 2 O CO 2 + H 2
CO 2 + H 2 H 2 O + CO
The catalyst stability of the SrNiO3 perovskite catalysts was examined throughout 50 h at 700 °C shown in Figure 10. It indicated that the SrNiO3/NiF(R) catalyst shows significantly stable catalytic activity during the DRP, among other catalysts. For instance, the SrNiO3/NiF(R) showed no significant conversion loss of C3H8 and CO2 was 85% and 66% from the begin 87% and 69%, respectively, after 50 h of the DRP. The significant activity and long-term stability of SrNiO3/NiF(R) are exhibited for its high thermal stability, high dispersion, and gas permeability of support [30]. In notably, the unreduced SrNiO3 perovskite catalyst showed a gradual decrease the catalytic activity even in two support materials for the period. The conversion of C3H8 and CO2 over the SrNiO3/γ-Al2O3(F) was 87% and 53% until 2 h of the reaction, which gradually decreased to 82% and 48%, respectively, after 50 h due to the amphoteric property of γ-Al2O3.
The catalytic stability results are shown in Table 2, the conversion of C3H8, CO2 and the H2/CO ratio in the steady state at 700 °C after 50 h. The C3H8 and CO2 conversion over SrNiO3/NiF(R) were superior to other catalysts. Besides, the H2/CO ratio was 0.7 for SrNiO3/NiF(R) and (F), which is closed to the stoichiometric value of DRP. Moreover, the results claimed that the higher selectivity of nickel foam-supported catalysts resist the carbon formation compared to SrNiO3/γ-Al2O3(F) and (R). According to the results, The SrNiO3 perovskite catalyst is a basic oxide. Due to its strong basicity, low surface area, and supported by NiF, CO2 might interact actively with basic sites favoring the formation of carbonates to coverts CO, which minimize the carbon formation during DRP. Several authors [49,50,51] found that the highly basic catalysts like lanthanum oxide and SmCoO3 are almost present in the carbonate phase on the catalyst surface after reforming reaction. These prevent the sintering of particles and extraction of particles from the surface of carbon filaments during the reaction.
The performance of the dry reforming of methane and propane over the perovskite-type catalysts were compared with the reported results, and the data are presented in Table 3. Strontium substituted perovskite catalyst were performed good catalytic conversion in dry reforming reaction and the H2/CO ratio of all the catalysts was less than the 1, which lower than the stoichiometry value of reforming reaction. The comparison suggested the excellent catalytic activity of SrNiO3 supported catalysts compared to the reported materials.

2.3. Catalytic Characterization of Spent Catalyst

The post characterization of catalyst was examined by several techniques such as FE-SEM, Raman, temperature-programmed oxidation (TPO), and XRD to understand the catalyst after DRP at elevated temperature. Figure 11 showed the FE-SEM of all examined catalysts after DRP over time on stream (TOS) 50 h at 1μm magnification. As seen in Figure 11, the morphology of carbon formation on the catalyst was observed whisker/filament form. It is agreed with many previous reports because most of Ni present catalysts are the response to the carbon growth like tubular at high temperature [56,57].
As shown in Figure 11, the SrNiO3/NiF (R) catalyst showed significantly low carbon on the surface among others, and the form of carbon also was a filament in nature. Figure 12 exhibited the carbon species of the all catalysts after DRP by Raman spectra. From the results, the carbon species were majorly in the form of graphitic. Three active peaks were observed in all catalysts. The first peak at 1337–1342 cm−1 belongs to the D-band of Raman active mode of C—C bond stretching. The second peak at 1572–1580 cm−1 corresponds to the G-band, which attributed to graphitized carbon form and also the third peak at 2678–2691 cm−1 is attributed to the 2D-band of carbon nanotubes or filaments. The intensity of D-band (ID) and intensity of G-band (IG) could explain the graphitic disorder [58,59]. In this case of SrNiO3/γ-Al2O3 (F and R) catalyst, the ID was always higher than the IG, which indicates deposited carbon could be in the form of graphite. The SrNiO3/NiF (F and R) catalyst showed that the different trend of ID and IG was almost same, which suggested the existence of crystalline graphite [60].
TPO studies were done to understand the carbon species formation during DRP and its shown in Figure 13. The three primary types of carbon occurred on the surface of the catalyst after DRP. These types were distinguished by oxidation temperature of solid carbon with O2. These are polymeric amorphous films or filaments (Cβ) at 523–773 K, the vermicular carbon filaments/fibers/tubes (Cν) at 573–1273 K and the crystalline graphite (CC) at 773–823 K [61,62,63].
The results of the TPO profile, the reduced catalysts showed the three kinds of Cβ, CC, and Cν on the surface, interestingly the both of reduced catalysts have vermicular carbon filaments. The un-reduce catalyst showed the polymeric amorphous films or filaments and crystalline graphite majorly [64]. These suggested that all catalyst produce carbon during DRP, even though the fresh catalyst showed a higher amount of carbon than reduced catalyst. The γ-Al2O3 supported catalysts produced more carbon instead of NiF, which suggested that the NiF allows increasing the molecule interaction due to their porous properties.
XRD analysis of the used catalysts was examined after DRP stability test is shown in Figure 14. The patterns show most of the diffraction peaks were matched with the fresh catalysts patterns (Figure 1b,c). The patterns show the new three peaks at 25.8°, 46.5°, and 49.9° were identified in all catalysts, which corresponds to the SrCO3 phase (JCPDS card no. 01-0556) [65]. Sutthiumporn et al. [56] reported that the Sr-doped La2O3 catalyst produces bidentate carbonate during the DRM reaction, which majorly reduces the carbon formation by the interaction of CO2. The peak at a 2θ value of 25.8° corresponds to the carbon species of all catalysts, which overlapped with SrCO3 peak. The reduced catalysts show the high intense peak of SrCO3 due to SrO reacts with CO2 to form SrCO3 [66]. This result confirmed that the higher CO2 conversion of reduced catalysts than unreduced catalysts.

3. Experimental Section

3.1. Materials and Methods

Strontium nitrate (Sr(NO3)2, Nickel nitrate hexahydrate (Ni(NO3)2·6H2O), citric acid anhydrous, hydrochloric acid (HCl), and anhydrous ethanol (C2H5OH) were purchased from Daejung Chemicals Ltd., (Gyeonggi-do, South Korea). Gamma alumina (γ-Al2O3) pellets were purchased from Alfa Aesar Co., Inc, (Seoul, South Korea). Nickel foam (NiF) was purchased from MTI Crop., Ltd., (Seoul, South Korea). All the chemicals used in this experiment were of research grade, and de-ionized water (DI) was used in whole experiments.

3.2. Synthesis of SrNiO3 Compound

SrNiO3 based perovskite catalyst was prepared by the citric acid sol-gel method. In a typical preparation method, the precursor of 0.02 M of Sr(NO3)2, and Ni(NO3)2·6H2O was dissolved in DI water in two separate beakers. These two solutions were mixed until to get a clear solution. Citric acid (0.06 m) in water was added by dropwise in the above nitrate solution to ensure miscibility. The solution was kept at 80 °C hot plate and stirred gently with Teflon-coated magnetic stir bar until the solution becomes viscous liquid. The resulting solution was kept at 110 °C for 12 h to dried off the excess of water. The dried powder was treated at 450 °C for 5 h to allow combustion reaction the resulting product was foamy. The final powder was ground well then calcined at 900 °C for ten hours, resulting in the synthesis of SrNiO3 perovskite compound.

3.3. Preparation of Catalyst

The 10 wt.% of catalyst was prepared with different support materials (γ-Al2O3 and NiF) as follow procedure. In order to remove undesired materials on the surface of support materials, both supports were pretreated to remove the moisture in γ-Al2O3 and to remove oxide layers on NiF. The γ-Al2O3 pellets were washed several times with DI water to swipe out of loss bound particles on the surface final rinse with ethanol and then kept at 150 °C for 2 h to remove moistures. The purchased NiF was soaked in 1.0 M HCl solution for 30 min to remove the surface oxide layer and then washed several times with DI water to remove excess of acid and finally rinsed with ethanol and kept 110 °C for 1 h. The 16.0 mm diameter disc of pretreated NiF was cut by hole puncher to attain a disc shape. The desired amount (0.2 gm) of SrNiO3 was ground with polyvinylidene difluoride (catalyst:PVDF ratio: 95:5) using N-methyl pyrrolidone (NMP) as a solvent to form a slurry. The 1.8 gm of the NiF discs (16 mm; OD) support was then coated with the prepared slurry, and the coated support was allowed to dry in an air oven at 85 °C for 12 h. The pretreated γ-Al2O3 pellets were ground well and made a fine powder, the 1.8 gm of fine γ-Al2O3 powder was mixed with 0.2 gm of SrNiO3 powder to ensure the blending between γ-Al2O3 and SrNiO3. A hydraulic press was used to prepare the pressured pellet, and the pellet was broken down into a small size by manually and sieved with 2 mm of mesh to remove the fine powder.

3.4. Material Characterization

X-ray diffractograms (XRD) were recorded an X-ray diffractometer (D/MAX 2200H, Bede 200, Rigaku Instruments C, Tokyo, Japan) with a horizontal goniometer performed by a fine focus copper X-ray tube (40 kV, 40 mA). BET specific surface area was measured by an Autosorb-1-MP instrument (Boynton Beach, FL, USA) at liquid nitrogen temperature. Prior to the analysis, the sample was degassed at 150 °C for 3 h, and the BET multi-point method was applied to estimate the surface area. The surface morphology of the synthesized materials was analyzed by Field-emission scanning electron microscopy (FE-SEM, JSM-6700F, JEOL, Tokyo, Japan). The Raman spectroscopy was evaluated on the samples using Raman HR Evolution Raman Spectrometer (LabRAM Horiba, Longjumeau, France) used at Ar+ ion laser operating at 10 mW and a wavelength of 514 nm. Temperature-programmed desorption/reduction (TPD/TPR) experiments were carried out on the gas-chromatography (DS science, DS-6500, Gyeonggi-Do, South Korea) equipped with TCD detector. In case of TPR, the 100 mg of the catalyst sample was loaded in a U-shaped quartz tube and placed on the quartz wool. The temperature of the sample was measured using a thermocouple fixed with quartz tube near the sample. The sample was degassed at 250 °C for 30 min in a flow of Ar (45 mL/min) and to remove the moisture and other adsorbed gases from surface and pores of the sample and then the reactor cooled down to room temperature. The hydrogen was measured with the ramping of the sample temperature to 900 °C with a heating rate of 5 °C/min. Hydrogen consumption was measured by the difference in the thermal conductivity of the gas mixture., Aforementioned in TPR analysis until the degassed was same as the TPD analysis. Carry out H2 adsorption at room temperature for 30 min by passing 5%H2/Ar mixed gas with a flow rate of 50 mL/min. H2 gas cut off and purge with pure Ar (45 mL/min) at room temperature for 30 min to remove physically adsorbed H2. The H2 desorption signal (as a function of time) was recorded by using a GC-TCD with a linear temperature increase from 25 °C to 800 °C with a heating rate of 5°C/min under Ar flow (45 mL/min). The amount of desorped H2 (mol) was measured the area under the curve. The following equation was used to calculate the amount of chemisorped H2.
Area ( mol ) = 0 t [ H 2 ] Q total dt
where [H2] is a concentration of H2 in mol/L and Qtotal is total flow rate L/min. Temperature-programmed oxidation (TPO) experiments of spent catalyst were carried out on the Fourier transform infrared spectroscopy (FTIR) (FTIR-7600 spectrometer, Lambda, Edwardstown, Australia) to understand the nature of carbon species. The 100 mg of spent catalyst was placed in a U-shaped quartz tube and flush with 50 mL/min of N2 at 250 °C for 30 min prior to remove surface oxygen and to attain inert atmosphere and reactor cool down to room temperature with continues flow of N2 gas. The 10%O2/N2 (50 mL/min) of oxidant gas feed was changed over, and outlet of gas (CO2) was monitor through online FTIR at a ramped temperature from 25 °C to 900 °C at the rate of 5 °C/min.

3.5. Catalytic Activity and Selectivity

The catalytic activity of prepared catalysts was measured in fixed-bed quartz reactor (16 mm ID and 600 mm length). A 2.0 g of the catalyst was placed in the reactor between a sandwich of quartz wool. A tubular furnace has maintained the reaction temperature of the catalyst with an external temperature controller equipped with a K-type thermocouple. The reaction temperature of the catalyst was measured by a K-type thermocouple which fixed on it. The ratio of feed gases of carbon dioxide (CO2) and propane (C3H8) (CPR) was 3, the total composition of feed gases are C3H8:CO2:Ar in percentage (10:30:60) and controlled by a mass flow controller (MFC-500, Atovac Co., Yongin, Korea). The total flow rate of the reactant gas is 200 mL/min for all catalytic studies. The concentration of C3H8 and CO2 were measured by online gas chromatography (GC- Micro-GCCP-4900, 10m PPQ column, Palo Alto, CA, USA) and the concentration of H2 and CO were measured by GC (DS-Science, 20m- HayeSep-Q column, Gyeonggi-Do, South Korea) equipped with a thermal conductive detector (TCD). The catalytic activity of all catalysts was examined at every 50 °C interval of the temperatures from 550 °C to 700 °C. The conversion of reactants (XA) and selectivity (S) of various products were calculated by the following equations [67].
X A ( % ) = C A 0 C A C A 0 + ε A C A × 100
S H 2 = [ H 2 ] out F out ( mL min ) 4 ( [ C 3 H 8 ] in F in [ C 3 H 8 ] out F out ) ( mL min )
S CO = [ CO ] out F out ( mL min ) 3 ( [ C 3 H 8 ] in F in [ C 3 H 8 ] out F out ) + ( [ CO 2 ] in F in [ CO 2 ] out F out ) ( mL min )
H 2 CO = F H 2 F CO
where, XA is the conversion of C3H8 and CO2, CA0, and CA are inlet and outlet concentration of reactants C3H8 and CO2, respectively, Fin and Fout are inlet and outlet flow rates (mL/min). [C3H8]in, [CO2]in, [C3H8]out, [CO2]out, [H2]out, and [CO]out were inlet and outlet concentration of C3H8, CO2, H2, and CO, respectively. The FH2 and FCO molar flow rate of H2 and CO in the outlet.

4. Conclusions

SrNiO3 perovskite catalyst successfully synthesized by the citrate sol-gel method and used as a catalyst with γ-Al2O3 and NiF supports for the first time in propane dry reforming and found to have propane and CO2 activity for syngas production efficiently. The effect of reduced SrNiO3 perovskite catalyst on DRP has been investigated. For comparison study, the reduced and unreduced SrNiO3 catalysts showed significant improvement in conversion of C3H8 and CO2. The SrNiO3/NiF(R) showed excellent activity regarding syngas production, the syngas produced with a significant selectivity of H2 and CO and H2/CO ratio was maintained close to the stoichiometric value. These measurements of catalytic activity with two different support materials have significantly affected the production of syngas. The results comprise to support perovskites on porous material like metallic foam and high surface area material like γ-Al2O3 to increase the number of exposed perovskite active sites. The strong basicity of strontium metal with NiF could aid the CO production and reduce carbon formation. The finding of this study has presented SrNiO3 perovskite as a suitable catalyst in the DRP experiment and also could be the replacement of rare earth perovskite in reforming reaction and cost effective with the less health-hazardous catalyst.

Author Contributions

S.M.S.P., M.M.H., G.G., performed the experimental work and analyzed the data; S.M.S.P. wrote the paper; Y.S.M. examined the experimental data, manuscript and supervised all the study.

Funding

This research received no external funding.

Acknowledgments

This work was supported by the Basic Science Research Program through the National Research Foundation funded by the Ministry of Science, ICT and Future Planning, Korea (Grant Nos. 2016R1A2A2A05920703 and 2018R1A4A1025998).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Xu, J.; Yeung, C.M.Y.; Ni, J.; Meunier, F.; Acerbi, N.; Fowles, M.; Tsang, S.C. Methane steam reforming for hydrogen production using low water-ratios without carbon formation over ceria coated Ni catalysts. Appl. Catal. A Gen. 2008, 345, 119–127. [Google Scholar] [CrossRef]
  2. Sutthiumporn, K.; Maneerung, T.; Kathiraser, Y.; Kawi, S. CO2 dry-reforming of methane over La0.8Sr0.2Ni0.8M0.2O3 perovskite (M = Bi, Co, Cr, Cu, Fe): Roles of lattice oxygen on C–H activation and carbon suppression. Int. J. Hydrogen Energy 2012, 37, 11195–11207. [Google Scholar] [CrossRef]
  3. Sudhakaran, M.S.P.; Sultana, L.; Hossain, M.M.; Pawlat, J.; Diatczyk, J.; Brüser, V.; Reuter, S.; Mok, Y.S. Iron–ceria spinel (FeCe2O4) catalyst for dry reforming of propane to inhibit carbon formation. J. Ind. Eng. Chem. 2018, 61, 142–151. [Google Scholar] [CrossRef]
  4. Todic, B.; Nowicki, L.; Nikacevic, N.; Bukur, D.B. Fischer–Tropsch synthesis product selectivity over an industrial iron-based catalyst: Effect of process conditions. Catal. Today 2016, 261, 28–39. [Google Scholar] [CrossRef] [Green Version]
  5. Siahvashi, A.; Adesina, A.A. Synthesis gas production via propane dry (CO2) reforming: Influence of potassium promotion on bimetallic Mo-Ni/Al2O3. Catal. Today 2013, 214, 30–41. [Google Scholar] [CrossRef]
  6. Staniforth, J.; Evans, S.E.; Good, O.J.; Darton, R.J.; Ormerod, R.M. A novel perovskite based catalyst with high selectivity and activity for partial oxidation of methane for fuel cell applications. Dalton Trans. 2014, 43, 15022–15027. [Google Scholar] [CrossRef] [PubMed]
  7. Nojoumi, H.; Dincer, I.; Naterer, G.F. Greenhouse gas emissions assessment of hydrogen and kerosene-fueled aircraft propulsion. Int. J. Hydrogen Energy 2009, 34, 1363–1369. [Google Scholar] [CrossRef]
  8. Gnanamani, M.K.; Jacobs, G.; Shafer, W.D.; Davis, B.H. Fischer–Tropsch synthesis: Activity of metallic phases of cobalt supported on silica. Catal. Today 2013, 215, 13–17. [Google Scholar] [CrossRef]
  9. Pashchenko, D. Thermodynamic equilibrium analysis of combined dry and steam reforming of propane for thermochemical waste-heat recuperation. Int. J. Hydrogen Energy 2017, 42, 14926–14935. [Google Scholar] [CrossRef]
  10. 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]
  11. Karuppiah, J.; Mok, Y.S. Plasma-reduced Ni/γ–Al2O3 and CeO2–Ni/γ–Al2O3 catalysts for improving dry reforming of propane. Int. J. Hydrogen Energy 2014, 39, 16329–16338. [Google Scholar] [CrossRef]
  12. Malaibari, Z.O.; Amin, A.; Croiset, E.; Epling, W. Performance characteristics of Mo–Ni/Al2O3 catalysts in LPG oxidative steam reforming for hydrogen production. Int. J. Hydrogen Energy 2014, 39, 10061–10073. [Google Scholar] [CrossRef]
  13. Barroso-Quiroga, M.M.; Castro-Luna, A.E. Catalytic activity and effect of modifiers on Ni-based catalysts for the dry reforming of methane. Int. J. Hydrogen Energy 2010, 35, 6052–6056. [Google Scholar] [CrossRef]
  14. Foo, S.Y.; Cheng, C.K.; Nguyen, T.-H.; Adesina, A.A. Kinetic study of methane CO2 reforming on Co–Ni/Al2O3 and Ce–Co–Ni/Al2O3 catalysts. Catal. Today 2011, 164, 221–226. [Google Scholar] [CrossRef]
  15. Corthals, S.; Van Nederkassel, J.; Geboers, J.; De Winne, H.; Van Noyen, J.; Moens, B.; Sels, B.; Jacobs, P. Influence of composition of MgAl2O4 supported NiCeO2ZrO2 catalysts on coke formation and catalyst stability for dry reforming of methane. Catal. Today 2008, 138, 28–32. [Google Scholar] [CrossRef]
  16. Yamazoe, N.; Teraoka, Y. Oxidation catalysis of perovskites—Relationships to bulk structure and composition (valency, defect, etc.). Catal. Today 1990, 8, 175–199. [Google Scholar] [CrossRef]
  17. Rynkowski, J.; Samulkiewicz, P.; Ladavos, A.K.; Pomonis, P.J. Catalytic performance of reduced La2−xSrxNiO4 perovskite-like oxides for CO2 reforming of CH4. Appl. Catal. A Gen. 2004, 263, 1–9. [Google Scholar] [CrossRef]
  18. Mawdsley, J.R.; Krause, T.R. Rare earth-first-row transition metal perovskites as catalysts for the autothermal reforming of hydrocarbon fuels to generate hydrogen. Appl. Catal. A Gen. 2008, 334, 311–320. [Google Scholar] [CrossRef]
  19. Dinka, P.; Mukasyan, A.S. Perovskite catalysts for the auto-reforming of sulfur containing fuels. J. Power Sources 2007, 167, 472–481. [Google Scholar] [CrossRef]
  20. Gallego, G.S.; Mondragón, F.; Barrault, J.; Tatibouët, J.-M.; Batiot-Dupeyrat, C. CO2 reforming of CH4 over La–Ni based perovskite precursors. Appl. Catal. A Gen. 2006, 311, 164–171. [Google Scholar] [CrossRef]
  21. Chawl, S.K.; George, M.; Patel, F.; Patel, S. Production of Synthesis Gas by Carbon Dioxide Reforming of Methane over Nickel based and Perovskite Catalysts. Procedia Eng. 2013, 51, 461–466. [Google Scholar] [CrossRef]
  22. Nair, M.M.; Kaliaguine, S.; Kleitz, F. Nanocast LaNiO3 Perovskites as Precursors for the Preparation of Coke-Resistant Dry Reforming Catalysts. ACS Catal. 2014, 4, 3837–3846. [Google Scholar] [CrossRef]
  23. Kondakindi, R.R.; Kundu, A.; Karan, K.; Peppley, B.A.; Qi, A.; Thurgood, C.; Schurer, P. Characterization and activity of perovskite catalysts for autothermal reforming of dodecane. Appl. Catal. A Gen. 2010, 390, 271–280. [Google Scholar] [CrossRef]
  24. Osazuwa, O.U.; Setiabudi, H.D.; Rasid, R.A.; Cheng, C.K. Syngas production via methane dry reforming: A novel application of SmCoO3 perovskite catalyst. J. Nat. Gas Sci. Eng. 2017, 37, 435–448. [Google Scholar] [CrossRef]
  25. Goldwasser, M.R.; Rivas, M.E.; Pietri, E.; Pérez-Zurita, M.J.; Cubeiro, M.L.; Gingembre, L.; Leclercq, L.; Leclercq, G. Perovskites as catalysts precursors: CO2 reforming of CH4 on Ln1−xCaxRu0.8Ni0.2O3 (Ln = La, Sm, Nd). Appl. Catal. A Gen. 2003, 255, 45–57. [Google Scholar] [CrossRef]
  26. Pichas, C.; Pomonis, P.; Petrakis, D.; Ladavos, A. Kinetic study of the catalytic dry reforming of CH4 with CO2 over La2−xSrxNiO4 perovskite-type oxides. Appl. Catal. A Gen. 2010, 386, 116–123. [Google Scholar] [CrossRef]
  27. Valderrama, G.; Kiennemann, A.; Goldwasser, M.R. La-Sr-Ni-Co-O based perovskite-type solid solutions as catalyst precursors in the CO2 reforming of methane. J. Power Sources 2010, 195, 1765–1771. [Google Scholar] [CrossRef]
  28. Smorygo, O.; Mikutski, V.; Marukovich, A.; Vialiuha, Y.; Ilyushchanka, A.; Mezentseva, N.; Alikina, G.; Vostrikov, Z.; Fedorova, Y.; Pelipenko, V.; et al. Structured catalyst supports and catalysts for the methane indirect internal steam reforming in the intermediate temperature SOFC. Int. J. Hydrogen Energy 2009, 34, 9505–9514. [Google Scholar] [CrossRef]
  29. Podyacheva, O.Y.; Ketov, A.A.; Ismagilov, Z.R.; Ushakov, V.A.; Bos, A.; Veringa, H.J. Metal foam supported perovskite catalysts. React. Kinet. Catal. Lett. 1997, 60, 243–250. [Google Scholar] [CrossRef]
  30. Takeda, Y.; Hashino, T.; Miyamoto, H.; Kanamaru, F.; Kume, S.; Koizumi, M. Synthesis of SrNiO3 and related compound, Sr2Ni2O5. J. Inorg. Nucl. Chem. 1972, 34, 1599–1601. [Google Scholar] [CrossRef]
  31. De Abreu, W.C.; De Moura, C.V.R.; Costa, J.C.S.; De Moura, E.M. Strontium and nickel heterogeneous catalysts for biodiesel production from macaw oil. J. Braz. Chem. Soc. 2017, 28, 319–327. [Google Scholar] [CrossRef]
  32. Luo, N.; Liu, K.X.; Li, X.; Shen, H.; Wu, S.; Fu, Z. Systematic study of detonation synthesis of Ni-based nanoparticles. Chem. Eng. J. 2012, 210, 114–119. [Google Scholar] [CrossRef]
  33. Rivas, M.E.; Fierro, J.L.G.; Goldwasser, M.R.; Pietri, E.; Pérez-Zurita, M.J.; Griboval-Constant, A.; Leclercq, G. Structural features and performance of LaNi1−xRhxO3 system for the dry reforming of methane. Appl. Catal. A Gen. 2008, 344, 10–19. [Google Scholar] [CrossRef]
  34. Pereñíguez, R.; González-DelaCruz, V.M.; Holgado, J.P.; Caballero, A. Synthesis and characterization of a LaNiO3 perovskite as precursor for methane reforming reactions catalysts. Appl. Catal. B Environ. 2010, 93, 346–353. [Google Scholar] [CrossRef]
  35. Sudhakaran, M.S.P.; Trinh, H.Q.; Karuppiah, J.; Hossain, M.M.; Mok, Y.S. Plasma Catalytic Removal of p-Xylene from Air Stream Using γ-Al2O3 Supported Manganese Catalyst. Top. Catal. 2017, 60, 944–954. [Google Scholar] [CrossRef]
  36. Tang, C.; Pu, Z.; Liu, Q.; Asiri, A.M.; Sun, X.; Luo, Y.; He, Y. In Situ Growth of NiSe Nanowire Film on Nickel Foam as an Electrode for High-Performance Supercapacitors. ChemElectroChem 2015, 2, 1903–1907. [Google Scholar] [CrossRef]
  37. Kumar, M.; Srikanth, S.; Ravikumar, B.; Alex, T.C.; Das, S.K. Synthesis of pure and Sr-doped LaGaO3, LaFeO3 and LaCoO3 and Sr,Mg-doped LaGaO3 for ITSOFC application using different wet chemical routes. Mater. Chem. Phys. 2009, 113, 803–815. [Google Scholar] [CrossRef]
  38. Valderrama, G.; Goldwasser, M.R.; de Navarro, C.U.; Tatibouët, J.M.; Barrault, J.; Batiot-Dupeyrat, C.; Martínez, F. Dry reforming of methane over Ni perovskite type oxides. Catal. Today 2005, 107–108, 785–791. [Google Scholar] [CrossRef]
  39. Chen, H.-Y.T.; Tosoni, S.; Pacchioni, G. Hydrogen Adsorption, Dissociation, and Spillover on Ru10 Clusters Supported on Anatase TiO2 and Tetragonal ZrO2 (101) Surfaces. ACS Catal. 2015, 5, 5486–5495. [Google Scholar] [CrossRef]
  40. Boudjahem, A.-G.; Monteverdi, S.; Mercy, M.; Bettahar, M.M. Nanonickel Particles Supported on Silica. Morphology Effects on Their Surface and Hydrogenating Properties. Catal. Lett. 2004, 97, 177–183. [Google Scholar] [CrossRef]
  41. Zhan, H.; Li, F.; Xin, C.; Zhao, N.; Xiao, F.; Wei, W.; Sun, Y. Performance of the La–Mn–Zn–Cu–O Based Perovskite Precursors for Methanol Synthesis from CO2 Hydrogenation. Catal. Lett. 2015, 145, 1177–1185. [Google Scholar] [CrossRef]
  42. Dong, X.; Zhang, H.; Lin, G.; Yuan, Y. Highly Active CNT-Promoted Cu–ZnO–Al2 O3 Catalyst for Methanol Synthesis from H2/CO/CO2. Catal. Lett. 2003, 85, 237–246. [Google Scholar] [CrossRef]
  43. Lu, X.; Gu, F.; Liu, Q.; Gao, J.; Liu, Y.; Li, H.; Jia, L.; Xu, G.; Zhong, Z.; Su, F. VOx promoted Ni catalysts supported on the modified bentonite for CO and CO2 methanation. Fuel Process. Technol. 2015, 135, 34–46. [Google Scholar] [CrossRef]
  44. Al-Doghachi, F.A.J.; Rashid, U.; Taufiq-Yap, Y.H. Investigation of Ce(iii) promoter effects on the tri-metallic Pt, Pd, Ni/MgO catalyst in dry-reforming of methane. RSC Adv. 2016, 6, 10372–10384. [Google Scholar] [CrossRef]
  45. Batiot-Dupeyrat, C.; Gallego, G.A.S.; Mondragon, F.; Barrault, J.; Tatibouët, J.-M. CO2 reforming of methane over LaNiO3 as precursor material. Catal. Today 2005, 107–108, 474–480. [Google Scholar] [CrossRef]
  46. Luo, J.Z.; Yu, Z.L.; Ng, C.F.; Au, C.T. CO2/CH4 reforming over Ni-La2O3/5 A: An investigation on carbon deposition and reaction steps. J. Catal. 2000, 194, 198–210. [Google Scholar] [CrossRef]
  47. Gennequin, C.; Safariamin, M.; Siffert, S.; Aboukaïs, A.; Abi-Aad, E. CO2 reforming of CH4 over Co–Mg–Al mixed oxides prepared via hydrotalcite like precursors. Catal. Today 2011, 176, 139–143. [Google Scholar] [CrossRef]
  48. Djinović, P.; Batista, J.; Pintar, A. Efficient catalytic abatement of greenhouse gases: Methane reforming with CO2 using a novel and thermally stable Rh–CeO2 catalyst. Int. J. Hydrogen Energy 2012, 37, 2699–2707. [Google Scholar] [CrossRef]
  49. Slagtern, A.; Schuurman, Y.; Leclercq, C.; Verykios, X.; Mirodatos, C. Specific Features Concerning the Mechanism of Methane Reforming by Carbon Dioxide over Ni/La2O3 Catalyst. J. Catal. 1997, 172, 118–126. [Google Scholar] [CrossRef]
  50. Olafsen, A.; Slagtern, Å.; Dahl, I.M.; Olsbye, U.; Schuurman, Y.; Mirodatos, C. Mechanistic features for propane reforming by carbon dioxide over a Ni/Mg(Al)O hydrotalcite-derived catalyst. J. Catal. 2005, 229, 163–175. [Google Scholar] [CrossRef]
  51. Pakhare, D.; Schwartz, V.; Abdelsayed, V.; Haynes, D.; Shekhawat, D.; Poston, J.; Spivey, J. Kinetic and mechanistic study of dry (CO2) reforming of methane over Rh-substituted La2Zr2O7 pyrochlores. J. Catal. 2014, 316, 78–92. [Google Scholar] [CrossRef]
  52. Rosen, B.A.; Gileadi, E.; Eliaz, N. Electrodeposited Re-promoted Ni foams as a catalyst for the dry reforming of methane. Catal. Commun. 2016, 76, 23–28. [Google Scholar] [CrossRef]
  53. Yang, E.; Noh, Y.; Ramesh, S.; Lim, S.S.; Moon, D.J. The effect of promoters in La0.9M0.1Ni0.5Fe0.5O3 (M=Sr, Ca) perovskite catalysts on dry reforming of methane. Fuel Process. Technol. 2015, 134, 404–413. [Google Scholar] [CrossRef]
  54. Kwon, B.W.; Oh, J.H.; Kim, G.S.; Yoon, S.P.; Han, J.; Nam, S.W.; Ham, H.C. The novel perovskite-type Ni-doped Sr0.92Y0.08TiO3 as a reforming biogas (CH4+CO2) for H2 production. Appl. Energy 2018, 227, 213–219. [Google Scholar] [CrossRef]
  55. Gurav, H.R.; Bobade, R.; Das, V.L.; Chilukuri, S. Carbon dioxide reforming of methane over ruthenium substituted strontium titanate perovskite catalysts. Indian J. Chem. Sect. A Inorg. Phys. Theor. Anal. Chem. 2012, 51, 1339–1347. [Google Scholar]
  56. Sutthiumporn, K.; Kawi, S. Promotional effect of alkaline earth over Ni-La2O3 catalyst for CO2 reforming of CH4: Role of surface oxygen species on H2 production and carbon suppression. Int. J. Hydrogen Energy 2011, 36, 14435–14446. [Google Scholar] [CrossRef]
  57. de Sousa, F.F.; de Sousa, H.S.A.; Oliveira, A.C.; Junior, M.C.C.; Ayala, A.P.; Barros, E.B.; Viana, B.C.; Filho, J.M.; Oliveira, A.C. Nanostructured Ni-containing spinel oxides for the dry reforming of methane: Effect of the presence of cobalt and nickel on the deactivation behaviour of catalysts. Int. J. Hydrogen Energy 2012, 37, 3201–3212. [Google Scholar] [CrossRef]
  58. de Sousa, H.S.A.; da Silva, A.N.; Castro, A.J.R.; Campos, A.; Filho, J.M.; Oliveira, A.C. Mesoporous catalysts for dry reforming of methane: Correlation between structure and deactivation behaviour of Ni-containing catalysts. Int. J. Hydrogen Energy 2012, 37, 12281–12291. [Google Scholar] [CrossRef]
  59. Pinheiro, A.L.; Pinheiro, A.N.; Valentini, A.; Filho, J.M.; de Sousa, F.F.; de Sousa, J.R.; Maria da Graça, C.R.; Bargiela, P.; Oliveira, A.C. Analysis of coke deposition and study of the structural features of MAl2O4 catalysts for the dry reforming of methane. Catal. Commun. 2009, 11, 11–14. [Google Scholar] [CrossRef]
  60. Pimenta, M.A.; Dresselhaus, G.; Dresselhaus, M.S.; Cançado, L.G.; Jorio, A.; Saito, R. Studying disorder in graphite-based systems by Raman spectroscopy. Phys. Chem. Chem. Phys. 2007, 9, 1276–1291. [Google Scholar] [CrossRef]
  61. Xu, L.; Song, H.; Chou, L. Carbon dioxide reforming of methane over ordered mesoporous NiO–MgO–Al2O3 composite oxides. Appl. Catal. B Environ. 2011, 108–109, 177–190. [Google Scholar] [CrossRef]
  62. Swaan, H.M.; Kroll, V.C.H.; Martin, G.A.; Mirodatos, C. Deactivation of supported nickel catalysts during the reforming of methane by carbon dioxide. Catal. Today 1994, 21, 571–578. [Google Scholar] [CrossRef]
  63. Rezaei, M.; Alavi, S.M.; Sahebdelfar, S.; Yan, Z.-F. Effects of CO2 content on the activity and stability of nickel catalyst supported on mesoporous nanocrystalline zirconia. J. Nat. Gas Chem. 2008, 17, 278–282. [Google Scholar] [CrossRef]
  64. Chesnokov, V.V.; Zaikovskii, V.I.; Buyanov, R.A.; Molchanov, V.V.; Plyasova, L.M. Morphology of carbon from methane on nickel-containing catalysts. Catal. Today 1995, 24, 265–267. [Google Scholar] [CrossRef]
  65. Beglaryan, H.A.; Melikyan, S.A.; Terzyan, A.M.; Isahakyan, A.R.; Zulumyan, N.H. Preparation of Hydrated Strontium Silicates from Silica Hydrogel Isolated from Serpentines and Their Thermal Transformations. Russ. J. Inorg. Chem. 2018, 63, 1395–1402. [Google Scholar] [CrossRef]
  66. Hu, B.; Mahapatra, M.K.; Keane, M.; Zhang, H.; Singh, P. Effect of CO2 on the stability of strontium doped lanthanum manganite cathode. J. Power Sources 2014, 268, 404–413. [Google Scholar] [CrossRef]
  67. Mok, Y.S.; Jwa, E.; Hyun, Y.J. Regeneration of C4H10 dry reforming catalyst by nonthermal plasma. J. Energy Chem. 2013, 22, 394–402. [Google Scholar] [CrossRef]
Figure 1. X-ray diffraction patterns of (a) SrNiO3 unreduced (F) and reduced (R) perovskite catalysts (b) SrNiO3 (F and R) supported in γ-Al2O3 (c) supported in nickel foam.
Figure 1. X-ray diffraction patterns of (a) SrNiO3 unreduced (F) and reduced (R) perovskite catalysts (b) SrNiO3 (F and R) supported in γ-Al2O3 (c) supported in nickel foam.
Catalysts 09 00068 g001
Figure 2. N2-adsorption and desorption isotherm of SrNiO3 perovskite catalyst.
Figure 2. N2-adsorption and desorption isotherm of SrNiO3 perovskite catalyst.
Catalysts 09 00068 g002
Figure 3. FE-SEM images of SrNiO3 perovskite catalyst at low (a) and high magnifications (b).
Figure 3. FE-SEM images of SrNiO3 perovskite catalyst at low (a) and high magnifications (b).
Catalysts 09 00068 g003
Figure 4. H2-TPR of (a) unreduced (F) and reduced (R) SrNiO3 perovskite unsupported catalysts and (b) SrNiO3 (F and R) supported with γ-Al2O3, and nickel foam catalysts.
Figure 4. H2-TPR of (a) unreduced (F) and reduced (R) SrNiO3 perovskite unsupported catalysts and (b) SrNiO3 (F and R) supported with γ-Al2O3, and nickel foam catalysts.
Catalysts 09 00068 g004
Figure 5. H2-TPD of the SrNiO3 perovskite catalyst.
Figure 5. H2-TPD of the SrNiO3 perovskite catalyst.
Catalysts 09 00068 g005
Figure 6. Catalytic activity of dry reforming of propane (DRP) as a function of temperature over SrNiO3/γ-Al2O3 (F and R) and SrNiO3/NiF (F and R) catalysts at 550 °C to 700 °C under atmospheric pressure in CPR = 3 (30:10:60 CO2:C3H8:Ar) and total flow rate = 200 mL/min (a) C3H8 conversion (b) CO2 conversion.
Figure 6. Catalytic activity of dry reforming of propane (DRP) as a function of temperature over SrNiO3/γ-Al2O3 (F and R) and SrNiO3/NiF (F and R) catalysts at 550 °C to 700 °C under atmospheric pressure in CPR = 3 (30:10:60 CO2:C3H8:Ar) and total flow rate = 200 mL/min (a) C3H8 conversion (b) CO2 conversion.
Catalysts 09 00068 g006
Figure 7. Catalytic activity of DRP as a function of temperature over bare supports γ-Al2O3 and NiF at 550 °C to 700 °C under atmospheric pressure in CPR = 3 (30:10:60 CO2:C3H8:Ar), and total flow rate = 200 mL/min, C3H8 and CO2 conversion.
Figure 7. Catalytic activity of DRP as a function of temperature over bare supports γ-Al2O3 and NiF at 550 °C to 700 °C under atmospheric pressure in CPR = 3 (30:10:60 CO2:C3H8:Ar), and total flow rate = 200 mL/min, C3H8 and CO2 conversion.
Catalysts 09 00068 g007
Figure 8. The product of DRP as a function of temperature over SrNiO3/γ-Al2O3 (F and R) and SrNiO3/NiF (F and R) catalysts at 550 °C to 700 °C under atmospheric pressure in CPR = 3 (30:10:60 CO2:C3H8:Ar) and total flow rate = 200 mL/min (a) H2 concentration (vol.%) and (b) CO concentration (vol.%).
Figure 8. The product of DRP as a function of temperature over SrNiO3/γ-Al2O3 (F and R) and SrNiO3/NiF (F and R) catalysts at 550 °C to 700 °C under atmospheric pressure in CPR = 3 (30:10:60 CO2:C3H8:Ar) and total flow rate = 200 mL/min (a) H2 concentration (vol.%) and (b) CO concentration (vol.%).
Catalysts 09 00068 g008
Figure 9. Catalytic performance of DRP over SrNiO3/γ-Al2O3 (F and R) and SrNiO3/NiF (F and R) catalysts at 700 °C, CPR = 3 (30:10:60 CO2:C3H8:Ar) and total flow rate = 200 mL/min.
Figure 9. Catalytic performance of DRP over SrNiO3/γ-Al2O3 (F and R) and SrNiO3/NiF (F and R) catalysts at 700 °C, CPR = 3 (30:10:60 CO2:C3H8:Ar) and total flow rate = 200 mL/min.
Catalysts 09 00068 g009
Figure 10. Stability of DRP over SrNiO3/γ-Al2O3(F and R) and SrNiO3/NiF (F and R) catalysts at 700 °C, CPR = 3 (30:10:60 CO2:C3H8:Ar) and total flow rate = 200 mL/min for 50 h.
Figure 10. Stability of DRP over SrNiO3/γ-Al2O3(F and R) and SrNiO3/NiF (F and R) catalysts at 700 °C, CPR = 3 (30:10:60 CO2:C3H8:Ar) and total flow rate = 200 mL/min for 50 h.
Catalysts 09 00068 g010
Figure 11. FE-SEM images of spent catalysts after DRP over 50 h (a) SrNiO3/NiF(R), (b) SrNiO3/NiF(F), (c) SrNiO3/γ-Al2O3 (R), and (d) SrNiO3/γ-Al2O3 (F).
Figure 11. FE-SEM images of spent catalysts after DRP over 50 h (a) SrNiO3/NiF(R), (b) SrNiO3/NiF(F), (c) SrNiO3/γ-Al2O3 (R), and (d) SrNiO3/γ-Al2O3 (F).
Catalysts 09 00068 g011
Figure 12. Raman spectra of spent catalysts after DRP over 50 h.
Figure 12. Raman spectra of spent catalysts after DRP over 50 h.
Catalysts 09 00068 g012
Figure 13. Temperature-programmed oxidation (TPO) profiles of used catalysts after DRP over 50 h.
Figure 13. Temperature-programmed oxidation (TPO) profiles of used catalysts after DRP over 50 h.
Catalysts 09 00068 g013
Figure 14. X-ray diffractogram of used catalyst (a) SrNiO3/γ-Al2O3 (F and R) and (b) SrNiO3/NiF (F and R) after DRP for 50 h.
Figure 14. X-ray diffractogram of used catalyst (a) SrNiO3/γ-Al2O3 (F and R) and (b) SrNiO3/NiF (F and R) after DRP for 50 h.
Catalysts 09 00068 g014
Table 1. Brunauer–Emmett–Teller (BET) surface area of the different supported catalysts.
Table 1. Brunauer–Emmett–Teller (BET) surface area of the different supported catalysts.
CatalystSBET (m2/g)
SrNiO3/γ-Al2O3(F)205.1
SrNiO3/γ-Al2O3(R)223.5
SrNiO3/NiF(F)6.2
SrNiO3/NiF(R)23.7
Table 2. C3H8 and CO conversion, CO, H2 selectivity, and H2/CO ratio.
Table 2. C3H8 and CO conversion, CO, H2 selectivity, and H2/CO ratio.
Catalyst X C 3 H 8   ( % ) X C O 2   ( % ) S C O   ( % ) S H 2   ( % ) H/CO2
SrNiO3/γ-Al2O3 (F)82.048.020.261.01.9
SrNiO3/γ-Al2O3 (R)92.052.146.263.21.0
SrNiO3/NiF(F)70.259.594.164.30.7
SrNiO3/NiF(R)85.367.097.767.70.7
W = 1.0 g, P = 1 atm, T = 700 °C, TOS = 50 h.
Table 3. Performance of the perovskite catalysts in dry reforming of methane.
Table 3. Performance of the perovskite catalysts in dry reforming of methane.
Composition of CatalystsCatalyst PreparationReforming ConditionsConversion at 700 °C (%)H2/CO RatioReferences
LaNiO3Sol-gel700 °C, CO2:CH4 = 165 (CH4)0.82[20]
Ni-Re/SS 316Electrodeposition700–800 °C, CO2:CH4 = 160 (CH4)0.8[52]
La0.8Sr0.2NiO3Sol-gel600–800°C, CO2:CH4 = 180.1 (CH4)NA[2]
La1−xSrxNi0.4Co0.6O3Sol-gel700 °C, CO2:CH4 = 175 (CH4)1.0[27]
La0.9Sr0.1Ni0.5Fe0.5O3Sol-gel700–900 °C, CO2:CH4 = 147.2 (CH4)0.7[53]
Sr0.92Y0.08TiO3-3%wt NiSol-gel625–745 °C, CO2:CH4 = 160 (CH4)NA[54]
SrTi0.85Ru0.15O3-δSol-gel600–800 °C, CO2:CH4 = 175 (CH4)0.8[55]
SrNiO3/NiF(R)Sol-gel550–700 °C, CO2:C3H8 = 385.3 (C3H8)0.7This work

Share and Cite

MDPI and ACS Style

M.S.P, S.; Hossain, M.M.; Gnanasekaran, G.; Mok, Y.S. Dry Reforming of Propane over γ-Al2O3 and Nickel Foam Supported Novel SrNiO3 Perovskite Catalyst. Catalysts 2019, 9, 68. https://doi.org/10.3390/catal9010068

AMA Style

M.S.P S, Hossain MM, Gnanasekaran G, Mok YS. Dry Reforming of Propane over γ-Al2O3 and Nickel Foam Supported Novel SrNiO3 Perovskite Catalyst. Catalysts. 2019; 9(1):68. https://doi.org/10.3390/catal9010068

Chicago/Turabian Style

M.S.P, Sudhakaran, Md. Mokter Hossain, Gnanaselvan Gnanasekaran, and Young Sun Mok. 2019. "Dry Reforming of Propane over γ-Al2O3 and Nickel Foam Supported Novel SrNiO3 Perovskite Catalyst" Catalysts 9, no. 1: 68. https://doi.org/10.3390/catal9010068

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