Next Article in Journal
The Influence of Cerium to Manganese Ratio and Preparation Method on the Activity of Ceria-Manganese Mixed Metal Oxide Catalysts for VOC Total Oxidation
Next Article in Special Issue
Fe-Promoted Alumina-Supported Ni Catalyst Stabilized by Zirconia for Methane Dry Reforming
Previous Article in Journal
Biocatalysis as a Green Approach for Synthesis of Iron Nanoparticles—Batch and Microflow Process Comparison
Previous Article in Special Issue
Effect of Modified Alumina Support on the Performance of Ni-Based Catalysts for CO2 Reforming of Methane
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Carbon Dioxide Valorization into Methane Using Samarium Oxide-Supported Monometallic and Bimetallic Catalysts

by
Radwa A. El-Salamony
1,
Ahmed S. Al-Fatesh
2,*,
Kenit Acharya
3,
Abdulaziz A. M. Abahussain
2,
Abdulaziz Bagabas
4,
Nadavala Siva Kumar
2,
Ahmed A. Ibrahim
2,
Wasim Ullah Khan
5 and
Rawesh Kumar
3,*
1
Process Development Department, Egyptian Petroleum Research Institute (EPRI), Cairo 11727, Egypt
2
Chemical Engineering Department, College of Engineering, King Saud University, P.O. Box 800, Riyadh 11421, Saudi Arabia
3
Department of Chemistry, Indus University, Ahmedabad 382115, Gujarat, India
4
President Office, King Abdulaziz City for Science and Technology (KACST), P.O. Box 6086, Riyadh 11442, Saudi Arabia
5
Chemical and Process Engineering, College of Engineering, University of Canterbury, Christchurch 8041, New Zealand
*
Authors to whom correspondence should be addressed.
Submission received: 6 November 2022 / Revised: 21 December 2022 / Accepted: 28 December 2022 / Published: 4 January 2023
(This article belongs to the Special Issue Catalytic Reforming of Light Hydrocarbons)

Abstract

:
Samarium oxide (Sm2O3) is a versatile surface for CO2 and H2 interaction and conversion. Samarium oxide-supported Ni, samarium oxide-supported Co-Ni, and samarium oxide-supported Ru-Ni catalysts were tested for CO2 methanation and were characterized by X-ray diffraction, nitrogen physisorption, infrared spectroscopy, H2-temperature programmed reduction, and X-ray photoelectron spectroscopy. Limited H2 dissociation and widely available surface carbonate and formate species over 20 wt.% Ni, dispersed over Sm2O3, resulted in ~98% CH4 selectivity. The low selectivity for CO could be due to the reforming reaction between CH4 (methanation product) and CO2. Co-impregnation of cobalt with nickel over Sm2O3 had high surface adsorbed oxygen and higher CO selectivity. On the other hand, co-impregnation of ruthenium and nickel over Sm2O3 led to more than one catalytic active site, carbonate species, lack of formate species, and 94% CH4 selectivity. It indicated the following route of CH4 synthesis over Ru-Ni/Sm2O3; carbonate → unstable formate → CO → CH4.

1. Introduction

The global concentration of carbon dioxide (CO2) has already reached dangerous levels, where island countries and temperature-sensitive animals are on the borderline of extinction due to global warming. It has affected the seasonal cycles and has brought natural disasters across the globe. If the scientific community did not find a way of decreasing CO2 concentration in the nearby future, the worst days would be faced. Besides the strategy of reducing CO2 emissions, scientific findings for the conversion of CO2 into useful chemicals are also encouraged worldwide [1,2].
The catalytic conversion of CO2 over metal oxide, doped metal oxide, and supported metal oxide has received the highest attention due to its catalytic potential, heterogeneity, and high-temperature sustainability. CO2 and H2 interaction over catalytic surface generates variety of products from hydrocarbon (olefin and paraffin) to oxygenates (CH3OH, CH3COOH) [3,4,5,6,7]. Formate is a surface-bonded carbon dioxide species, whereas hydrogen is easily dissociated over a metal surface such as iron, copper, silver, gold, platinum, palladium, rhodium, molybdenum, rhenium, or nickel [6,8,9,10]. Controlling catalytic surfaces becomes crucial to understanding product distribution. Formate species is not stable, except over Ni, and it is decomposed into CO under H2 flow through reverse water–gas shift reaction. Copper oxide (CuO) showed just 13% CO2 conversion with 10.4% CO yield and iron (III) oxide (Fe2O3) showed 8% CO2 conversion with 6.3% CO yield [11], indicating that unsupported CuO and Fe2O3 were not effective catalysts. Thus, the concept of promoted and supported metal catalysts has come into the picture, where the catalytic efficiency is augmented in terms of the dispersion of catalytic active sites over proper support and the role of promoters in modifying the physiochemical properties of supported catalyst systems. The comprehensive reaction scheme over the supported metal catalyst is shown in Figure 1.
When a “supported active metal iron” catalyst is utilized, carbon monoxide is converted into metal carbide, switches to the chain growth mechanism, and gives C1 to higher hydrocarbon (concerned with olefin and paraffin; Route 1) [12]. For route 1, the potassium- and copper-promoted, Al2O3-supported iron (Fe-Cu-K-Al) catalyst has been studied extensively [8,13,14]. Over the Cu-supported catalyst, CO hydrogenation takes place in two ways: the formic acid route (Route 2) and the methanol route through formaldehyde and methoxide intermediate (Route 3) [15]. The zirconium-promoted CuO-ZnO-Al2O3 (1:1:1) catalyst has received wide attention for high CO2 conversion and high CH3OH selectivity [9]. The supported Ag and Au catalysts prefer route 3 [10]. Supported 5.0 wt.% Ru metal induces an eight electron reduction in intermediate CO into CH4 (Route 4) [6]. The Al2O3-supported Ru catalyst had an atomic assembly with high dispersion of three dimensional Ru cluster. It showed about 55% CO2 conversion with 55% CH4 yield. Pd, Pt, Rh, Mo, and Re metals direct CO2 and H2 through routes 3 and 4, which lead to high selectivity to methanol and methane, respectively [10]. On the other hand, the supported Ni catalyst carries CO2 in the COx-free catalytic route (CO2 methanation route): CO2→ carbonate → formate → –CH2OH → –CH3 → CH4 (Route 5) [16]. This route is useful in generating COx-free CH4. Incorporating Co in Ni led to a smaller lattice constant and higher hydrogenation activity of CO2 [17,18]. Cobalt-based catalysts were also investigated greatly for CO2 conversion into light hydrocarbons [19].
Al2O3-supported Ni catalysts showed 50% CO2 conversion with 50% CH4 yield [20], while 3.0 wt.% Zr-promotional addition inhibited migration of Ni into Al2O3 matrix or inhibited the formation of NiAl2O4, caused >60% CO2 conversion and 70% CH4 yield [21]. The activity of La2O3-supported Ni catalyst toward CO2 methanation was worse because of the formation of stable lanthanum carbonate [La2(CO3)3], which did not decompose into formate [20]. In addition, the SiO2-supported Ni catalyst was not good for CO2 methanation because of the collapse of the Si-O-Si structure by hydrolysis of water, sintering of Ni particles, and rapid coke deposition and deactivation of the catalyst [22]. Among supported catalyst systems, the Y2O3-supported Ni catalyst displayed the maximum CO2 conversion (75%) with 75% CH4 yield due to the presence of more basic sites and easily reducible NiO species. However, in terms of CH4 selectivity, the ZrO2-supported Ni catalyst resulted in the best (60% CO2 conversion with 100% CH4 selectivity) [23].
For such an important reaction, other supports should also be investigated continuously. For instance, Sm2O3 had various favorable surface properties for CO2 and H2 interaction. It had Ni-size optimizing capacity like Y2O3, higher Ni dispersion and a stronger metal-support interaction than Y2O3 [24,25]. Electrochemically reduced Sm(II) compound was found to be a powerful catalyst for CO2 activation [26]. Sm2O3 was utilized in quantitative detection of CO2 and CO [27]. Moreover, a 5 atom% Sm addition over the Ni-Zr catalyst induced a higher H2 uptake [28]. Muroyama et al. showed that upon reduction, the Ni particle size distribution was the narrowest and their sizes were the smallest over Sm2O3 compared to other supports (Y2O3, Al2O3, La2O3, and ZrO2) [20]. A 5.0 wt.% Ni loading over Sm2O3 achieved 100% CH4 selectivity, but with ~20% CO2 conversion at 300 °C [29]. Upon incorporation of a 5.0 wt.% Ba into the Sm2O3-supported Ni catalyst, the CH4 selectivity remained 100% and CO2 conversion was maintained at 40% for 28-h TOS at 300 °C. The CO-free, high-performance catalytic route of the 5 wt.%Ni-5 wt.%Ba/Sm2O3 catalyst was due to the enhanced oxygen mobility (causing more oxygen vacancy), increased CO2 adsorption (causing the formation of more surface carbonate species), and enhanced hydrogenation of carbonate into formate species.
Inspired by its properties in favor of CO2 and H2 interaction, samarium oxide-supported Ni, samarium oxide-supported Co-Ni, and samarium oxide-supported Ru-Ni catalysts were prepared by the impregnation and co-impregnation method and were tested for CO2 methanation. Catalysts were characterized by X-ray diffraction (XRD), surface area and porosity, infrared spectroscopy, H2-temperature programmed reduction (H2-TPR), and X-ray photoelectron spectroscopy (XPS). Fine-tuning between surface properties and characterization results was found. This thorough study would help to understand the surface behavior of Sm2O3-supported metal catalysts toward CO2 methanation reaction.

2. Results

2.1. Characterization Results

Energy Dispersive X-ray (EDX) elemental analysis of 10Ni/Sm2O3, 20Ni/Sm2O3, 40Ni/Sm2O3, 10Ni-10Co/Sm2O3 and 10Ni-10Ru/Sm2O3 catalysts is shown in Figure S1. EDX elemental analysis showed that with increasing 10 wt.% to 40 wt.% Ni loading, surface atomic composition of Ni was increasing. The 10Ni-10Co/Sm2O3 catalyst showed that the surface composition of Ni over the 10Ni-10Co/Sm2O3 catalyst was minimum, whereas surface composition of Ni was maximum at 10Ni-10Ru/Sm2O3 catalyst. The X-ray diffraction results of the catalyst samples are shown in Figure 2 and diffraction peaks are matched with Joint Committee on Powder Diffraction Standards (JCPDS) reference card number. All catalysts showed low intensity diffraction peaks. The support had diffraction peaks for the Sm2O3 phases at 2θ = 27.68°, 28.11°, 29.04°, 29.74°, 30.64°, 31.09°, 31.86°, 32.61°, 38.59°, 40.51°, 46.73°, 50.13°, 51.44°, 53.24°, 54.70°, 55.72° (JCPDS reference card number 00-013-0244 and 00-042-1461). Upon the 10.0 wt.% Ni addition, the intensity for the diffraction peaks for the Sm2O3 phases decreased significantly. However, the peaks of the cubic NiO phase were not detected, indicating that up to 10 wt.% Ni loading over the Sm2O3, NiO was well-dispersed (Figure 2A). NiO dispersion and the reduction in the peak intensity of Sm2O3 in the 10Ni/Sm2O3 catalyst suggested the close interaction of NiO with Sm2O3 support. The same type of finding was observed over the Y2O3-supported Ni catalyst [30]. Upon increasing the NiO loading to 20.0 wt.%, the cubic NiO phase appeared at 2θ = 37.24°, 43.28°, 62.75° (JCPDS reference card number 01-073-1523) (Figure 2B). Furthermore, increasing the NiO loading to 40 wt.% led only to the observation of the diffraction peaks of Sm2O3 without detecting those diffraction peaks of the cubic NiO phase (Figure 2C). The diffraction patterns of 40Ni/Sm2O3 and 10Ni/Sm2O3 were comparable. However, the bimetallic catalyst of 10 wt.%Co-10 wt.%Ni, dispersed over Sm2O3 support, had a very similar diffraction pattern to that of 20Ni/Sm2O3 (Figure 2D). It indicated that in presence of CoO, the NiO crystalline phase was relatively more exposed (like 20Ni/Sm2O3). The bimetallic catalyst of 10 wt.%Ru-10 wt.%Ni, dispersed over Sm2O3 support, had no crystalline peaks for NiO, but it had diffraction peaks for the Sm2O3 phase and for the tetragonal RuO2 phase (at 2θ = 35.02°; JCPDS reference card number 00-040-1290) (Figure 2E).
The N2-adsorption isotherms, porosity distribution, BET surface area, BJH pore volume, and BJH pore diameter are depicted in Figure 3A,B and Figure S2. All samples showed type-IV isotherm with H3 hysteresis loop, which indicated non-rigid aggregates like-pore with unlimited adsorption at high p/p0 [31]. The Sm2O3-supported 10 wt.% Ni catalyst had the least surface area, and pore volume among the remaining supported catalysts. However, upon 20.0 wt.% Ni loading, the height of the hysteresis loop was amplified and the surface area, pore volume, and pore diameter were increased to four times (25.70l m2/g), six times (0.175 cc/g) and seven times (15.797 nm), respectively, implying that Ni did not only disperse over the catalyst surface, but also it was incorporated into the Sm2O3 framework and caused expansion of framework [29]. Upon 40.0 wt.% Ni impregnation, the height of the hysteresis loop was decreased sharply, and the surface area and pore volume were decreased sharply (to 12.7 m2/g and 0.061 cc/g, respectively) without affecting the pore diameter. It indicated the pore blocking by inclusions of NiO nanoparticles [32] upon higher loading. Co-impregnation of 10.0 wt.% Co and 10.0 wt.% Ni over Sm2O3 support was found to maximize the surface area and pore volume up to the highest limit (with respect to the rest catalysts). The surface area and pore volume of Co-Ni/Sm2O3 were enhanced more than 10 times without affecting the pore diameter (with respect to 10Ni/Sm2O3). The height of the hysteresis loop in the isotherm plot of the Co-Ni/Sm2O3 catalyst was also enhanced up to the highest extent. However, with a co-impregnation of 10.0 wt.% Ru and 10.0 wt.% Ni over Sm2O3 support, the surface parameters were dropped. In XRD, the tetragonal RuO2 was detected over the Ru-Ni/Sm2O3 catalyst, indicating that a sharp drop of the surface parameters was due to the large deposition of tetragonal RuO2 crystals into the pore of the catalyst. These phenomena are related mainly to the ionic radii of the employed metals, where in the case of Co ions, the ionic radius was equal to that of Ni ions, but in the case of Ru, it was bigger. These findings are coincident with our previous study on Y2O3 support [33]. It appeared that in the case of co-impregnation, the pore diameter of the catalyst was not modified and the pore diameter of Co-Ni/Sm2O3 and Ru-Ni/Sm2O3 and 10Ni/Sm2Os were within the range from 2.0 to 2.4 nm. These observations were quite informative and indicated the uptake capacity of the heteroatom in the Sm2O3 framework. If 10 wt.% Ni loading was taken as a standard, then further 10 wt.% loading of Ni on Sm2O3 caused the expansion of the entire framework. However, when 10 wt.% Co along with 10 wt.% Ni (total 10 wt.%Co-10 wt.% Ni) was loaded over Sm2O3, Co was incorporated in the Sm2O3 framework except on the pore opening of the surface. Thus, pore volume and surface area were increased, but not the pore size. During loading of 10 wt.% Ru along with 10 wt.% Ni over Sm2O3, RuO2 particles were largely deposited in the pore, causing decrements in all surface parameters.
Infrared (IR) spectra of the catalyst samples are shown in Figure 3C and Figure S3. The IR band at ~1620 cm−1 and 3429 cm−1 could be attributed to the bending and stretching vibrations of the hydroxyl groups over the catalyst surface, respectively [34]. These two bands were present even in the bare support of Sm2O3 (Figure S3). Upon Ni loading over Sm2O3, various other prominent bands appeared. The transmittance peak was at 852 cm−1 for CO32− symmetric stretching, 1044 cm−1 for the vibration of bidentate carbonate species, 1381 cm−1 and 1500 cm−1 for the vibration of unidentate carbonate species [35,36,37], 2850 cm−1 for the stretching vibration of C-H (of coordinated formate species), and 2925 cm−1 for the combination of the asymmetric stretching of –COO and the bending vibration of C–H bond (of bi-coordinated formate species) [38]. Interaction of CO2 with samarium (through different CO2 intermediates) was previously reported [39]. The IR spectra of the 20Ni/Sm2O3 catalyst could be specified by the intense spectral vibration of the unidentate carbonate species at 1500 cm−1, whereas the Co-Ni/Sm2O3 catalyst had the most intense vibration for the bicarbonate species at 1044 cm−1, and the unidentate carbonate species at 1381 cm−1. It is noticeable that the Ru-Ni/Sm2O3 catalyst had no vibration bands for the formate species (~2850 cm−1 and 2925 cm−1).
The Sm2O3-supported 10 wt.% Ni catalyst had reduction peaks at 360 °C and 475 °C with total adsorption of 2.197 mmol/g of H2. The earlier one was attributed to the reduction in NiO species, which interacted weakly with support, while the latter peak was attributed to the reduction in NiO species, which interacted strongly with the support [40]. For higher Ni loading (20.0 wt.%), the 20Ni/Sm2O3 catalyst had the highest H2 consumption (7.28 mmol/g) among the other Ni-loaded Sm2O3 catalysts with a merge reduction peak in the temperature range 327–615 °C (the peak maximum ~400 °C). It indicated the presence of a high concentration of reducible NiO species, interacting with the support up to different extents. For 40.0 wt.% Ni loading, the 40Ni/Sm2O3 catalyst had the worst reducible profile. It had a negligible amount of reducible NiO species (0.244 mmol/g H2 consumption). Accordingly, 40Ni/Sm2O3 had minimum reducibility. The reducibility profile of the bimetallic Co-Ni, dispersed over Sm2O3, had the maximum amount of H2 consumption (8.827 mmol/g) among all catalysts. The Co-Ni/Sm2O3 catalyst had the highest intensity of reducible peak about the peak maxima at 400 °C for the reduction in NiO-interacted species, low-temperature overlapping peaks ~300 °C for the reduction in Co3O4 to CoO, and a high-temperature reduction peak ~500 °C for the reduction in CoO to Co [41,42]. However, the high-temperature reduction peak ~500 °C in the Co-Ni/Sm2O3 catalyst might be related to the combining contributions of reducible NiO and CoO species. The H2-consumption profile of the bimetallic Ru-Ni, dispersed over Sm2O3, showed an additional reduction peak at a relatively lower temperature (~330 °C) for a partial reduction in dispersed RuOx species on the support [43]. The total hydrogen consumption of the Ru-Ni/Sm2O3 catalyst in the H2-temperature programmed reduction study was found 6.25 mmol/g.
There is a marked difference in the surface characteristics of both supported bimetallic catalysts. To understand the valance state and bonding character: Ni (2p), Sm (3d) and O (1s) XPS spectra of the Co-Ni/Sm2O3 and Ru-Ni/Sm2O3 catalysts were depicted in Figure 4A–D. The Ni2p spectra of both Sm2O3-supported bimetallic catalysts had mostly similar profiles. The absence of metallic Ni peak (about binding energy <854 eV and 869–870 eV) in Ni2p XPS spectra was confirmed in the Co-Ni/Sm2O3 and Ru-Ni/Sm2O3 catalysts [44,45]. The presence of Ni+2 oxidation state in the form of NiO compound was confirmed by Ni2p1/2 spectra ~871.8–873.8 eV and 879 eV, and Ni2p3/2 spectra ~854.3–855.7 eV [46,47,48,49] (Figure 4A). Sm3d3/2 and Sm3d5/2 spectra at ~1110.25 eV and 1083.24 eV, respectively, indicated the presence of Sm+3 oxidation state (Sm2O3 compound) [50,51] (Figure 4C). The binding energy values of the O1s profiles indicated the overlapping contributions of oxygen from various oxide species [52]. The O1s spectra with binding energy at 531.2 eV and 530 eV were noticed in both the Co-Ni/Sm2O3 and Ru-Ni/Sm2O3 catalysts (Figure 4D). The binding energy at 531.2 eV was attributed to adsorbed oxygen [53], whereas the binding energy at 530 eV signified the oxygen in samarium oxide [51]. Both XPS peak intensities were much less in the Ru-Ni/Sm2O3 catalyst than to their corresponding ones in the Co-Ni/Sm2O3 catalyst, indicating the availability of “surface adsorbed oxygen” and oxygen in samarium oxide in the Co-Ni/Sm2O3 catalyst. To bring more focus on the valance state and bonding character of the second metal over the Sm2O3-supported Ni catalyst, the Co (2p) spectrum of the Co-Ni/Sm2O3 catalyst and the Ru (3d) spectrum of the Ru-Ni/Sm2O3 catalyst were also recorded, as depicted in Figure S4, respectively. Co2p XPS spectra did not show a peak for metallic Co at binding energy ~778 eV, but it showed a peak with a binding energy centered at 780 eV for 2p3/2) for Co3O4 [42]. The Ru3d spectrum of the Ru-Ni/Sm2O3 catalyst showed a broad overlapping peak in the binding energy range 281.5–285 eV. It indicated the presence of Ru oxide species [54,55].

2.2. Catalytic Activity Result and Discussion

The reaction temperature suitable for CO2 methanation was tested over the 10Ni/Sm2O3 catalyst and was depicted in Figure S5. It was found that the mole% of CH4 formation from CO2 and H2 was negligible at 250 °C, notable at 300 °C, and prominent at 350 °C. At 350 °C temperature, the mole% of CO formation was null, which indicated that CH4 formation was carried out without including CO in the intermediate reaction step. In infrared spectra, a peak for formate was observed in all catalysts except 10Ru-10Ni/Sm2O3. The literature report suggested stable formate formation over the Ni-based catalyst system, which further underwent successive hydrogenation to methane through the methoxy intermediate route [16]. Thus, 350 °C was kept standard for CO2 methanation over our the Sm2O3-supported Ni catalyst system. Other research groups also found 350 °C as the suitable temperature for CO2 methanation [56,57,58]. The catalytic activity of the xNi/Sm2O3 (x = 10, 20, 40) and M-Ni/Sm2O3 (M = Co, Ru) catalysts in the term of CO2 conversion, CH4 selectivity, and CO selectivity are shown in Figure 5. XRD did not detect metallic Ni or NiO phase over the Sm2O3-supported 10.0 wt.% Ni catalyst and H2-TPR indicated the reduction peaks of NiO-interacted species. IR confirmed the presence of surface interacted CO2-species such carbonate and formate over 10Ni/Sm2O3 surface. Overall, 10Ni/Sm2O3 had widely available surface-interacted CO2-species and reducible NiO species (inducing H2 dissociation and spillover). The CO2 conversion remained constant at ~35% for 13 h, but constant selectivities for CO and CH4 were attained after 100 min of reaction. Initially, the CO and CH4 selectivities were comparative (47% and 52%), but after 100 min, the CO selectivity dropped to 5–6%, whereas the CH4 selectivity jumped to 94–95%. Further, the constant CH4 and CO selectivities were maintained for 13 h time-on-stream. This finding indicated that the 10Ni/Sm2O3 catalyst system needed some time for constant product selectivity.
For 20.0 wt.% loading of Ni over Sm2O3, the NiO phase appeared in the X-ray diffraction pattern. Assembly of NiO particles might sign toward less dispersion of catalytic active species and it might provide limitation in higher activity toward CO2 methanation. The surface area of the 20Ni/Sm2O3 catalyst was four times that of 10Ni/Sm2O3 and unidentate carbonate intensity in infrared spectra was also amplified. It had formate species and the highest number of reducible NiO species among the rest of the Ni-loaded Sm2O3 catalysts. However, the largely extendible surface and high amount of reducible NiO species over 20Ni/Sm2O3 did not turn into higher CO2 conversion, but it resulted in the highest CH4 selectivity. The 20Ni/Sm2O3 catalyst showed 24–26% CO2 conversion, ~97–98% CH4 selectivity, and 3–2% CO selectivity over 13 h of reaction. It indicated that a large number of reducible NiO species over large catalytic surfaces did not grant the higher CO2 conversion by H2. Hydrogen dissociation and its spillover concisely depend on the size of Ni particles. The higher NiO crystallinity over the 20Ni/Sm2O3 catalyst than 10Ni/Sm2O3 caused the formation of a relatively larger size of metallic Ni upon reduction, which limited H2 dissociation and thereby CO2 conversion. However, the relatively higher density of unidentate carbonate in the infrared spectrum of the 20Ni/Sm2O3 catalyst indicated that limited H2 dissociation, followed by CO2 interaction over a largely extended surface, led to the formation of unidentate carbonate to methane via a stable formate pathway. So, 20Ni/Sm2O3 conveyed highest ~98% CH4 selectivity. The 10Ni/Sm2O3 catalyst needed some time for constant product selectivity, but the 20Ni/Sm2O3 catalyst gave constant product selectivity from the beginning. The surface mechanism over the samarium oxide-supported Ni system could be outlined at this stage. CO2 was initially adsorbed as carbonate/bicarbonate species over the support of Sm2O3. The carbonate/bicarbonate species were hydrogenated to formate species. The formate species over Ni was stable and did not decompose into CO and H2. Strong binding between Ni and the formate species guided the formate species to further hydrogenation to methane through methoxy intermediate (Route 5) [16]. The presence of a very small amount of CO over 20Ni/Sm2O3 indicated the initiation of further reaction of CH4 with CO2 (dry reforming of methane) or “CH4 with H2O” (stream reforming of methane) over the surface as metallic Ni also catalyzes CH4 decomposition. For 40.0 wt.% impregnation of Ni over Sm2O3, the catalyst crystallinity was comparable to 10Ni/Sm2O3, but the surface area was decreased to the half and the amount of reducible NiO species over the catalyst surface was fallen suddenly. The shortage of reducible NiO species turned into a lower number of active sites (metallic Ni) after reduction. It caused just 8% CO2 conversion. It appeared that, at such a high Ni loading, the formate species were majorly destabilized and decomposed into CO. Furthermore, the catalyst surface, bearing fewer active sites, was not so capable to push CO toward methanation. Overall, due to the destabilization of the formate-hydrogenation pathway over 40Ni/Sm2O3, two third of CO2 was converted to CO and the remaining one third to CH4 (75% CO selectivity and 25% CH4 selectivity) during 180 min time-on-stream. To bring more focus to the reaction mechanism, Co-impregnation of Co-Ni and Ru-Ni over Sm2O3 support was carried out and was tested for CO2 methanation reaction. When 10.0 wt.% Co and 10.0 wt.% Ni were co-impregnated over Sm2O3, the crystallinity of the catalyst was comparable to that of the 20Ni/Sm2O3, whereas it had the highest surface area and pore volume among all catalysts. Including reducible nickel oxide and cobalt oxide, the 10Co-10Ni/Sm2O3 catalyst had the highest intensity of reducible oxide over the catalyst surface. In addition, it had an adequate amount of formate species and the highest intensity of bicarbonate/carbonate species over the catalyst surface. Altogether, the 10Co-10Ni/Sm2O3 catalyst showed initially high CO2 conversion (41%), which slowly and steadily dropped to 22% at the end of 13 h. Thus, diluting the active Ni sites by Co caused inferior catalytic activity in terms of CO2 conversion. In EDX elemental analysis also, minimum Ni distribution over the surface was found in the case of the 10Co-10Ni/Sm2O3 catalyst (than the remaining catalysts). The product selectivity pattern over the 10Co-10Ni/Sm2O3 was very similar to the 10Ni/Sm2O3, namely, the 10Co-10Ni/Sm2O3 catalyst also needed some time for constant product selectivity. After 100 min, the 10Co-10Ni/Sm2O3 catalyst had attained 88–89% CH4 selectivity and 12–11% CO selectivity. The relatively higher selectivity of CO over the 10Co-10Ni/Sm2O3 catalyst could be due to persuading reverse water–gas shift reaction over the cobalt-based catalyst [59]. When 10.0 wt.% Ru and 10.0 wt.% Ni were co-impregnated over Sm2O3, the cubic NiO phase was not found, but the tetragonal RuO2 phase appeared. It had a good amount of reducible NiO and RuOx species. It had an easier reducibility profile than other catalysts. The 10Ru-10Ni/Sm2O3 had five times less surface area and less surface adsorbed oxygen than the 10Co-10Ni/Sm2O3. Maximum surface distribution of Ni was confirmed over the 10Ru-10Ni/Sm2O3 catalyst by EDX elemental analysis (than rest catalyst). Among all catalyst systems, 10Ru-10Ni/Sm2O3 produced peaks for the carbonate species, but no peaks for the the formate species, indicating that the unstable formate species decomposed into CO. Further CO methanation might lead to methane formation over the 10Ru-10Ni/Sm2O3 catalyst. The previous literature also argued the mechanism of CO2 methanation in the same line [6]. The 10Ru-10Ni/Sm2O3 catalyst had a higher catalytic activity (31–33% CO2 conversion, 93–94% CH4 selectivity, and ~6% CO selectivity constantly from the beginning to 13 h) than 10Co-10Ni/Sm2O3 due to the presence of more than one active site as Ni and Ru metals over the catalyst surface.

3. Materials and Methods

3.1. Materials

All chemicals were purchased and were used as received: Sm2O3 (99.9% AR from Loba Chemie, Mumbai, India), Ni(NO3)2·6H2O (Acros Organics-Thermo Fisher Scientific, Geel, Belgium 99.9%), Co(NO3)2·6H2O (Acros Organics-Thermo Fisher Scientific, 99.9%), and RuCl3·xH2O (Acros Organics-Thermo Fisher Scientific, 99.9%).

3.2. Catalyst Preparation

Samarium oxide-supported Ni, samarium oxide-supported Co-Ni, and samarium oxide-supported Ru-Ni catalysts were prepared by wet impregnation and co-impregnation methods. For supported Ni catalysts, an appropriate amount of Ni(NO3)2·6H2O (equivalent to 10, 20, or 40 wt.% of Ni metal) was dissolved in water. For supported M-Ni (M = Co or Ru) bimetallic catalysts, an appropriate amount of metal precursor (equivalent to 10.0 wt.% Co or Ru) and Ni (NO3)2·6H2O (equivalent to 10.0 wt.% Ni) were dissolved in water. In all, 3.0 g pre-calcined Sm2O3 (at 500 °C for five hours) was added as support to these aqueous solutions at 80 °C under slow stirring until complete evaporation for four hours. The slurries were kept overnight at 110 °C in the oven and were then calcined at 500 °C under air for three hours. The prepared catalysts were abbreviated as: 10Ni/Sm2O3, 20Ni/Sm2O3, 40Ni/Sm2O3, Co-Ni/Sm2O3, and Ru-Ni/Sm2O3.

3.3. Catalysts Characterization

The surface area and porosity of catalyst samples were analyzed by NOVA 3200 apparatus (Quantachrome Corporation, Boynton Beach, FL, USA). Crystallinity and phase identification of the catalysts were studied by X-ray diffraction analysis (XRD) by using a Shimadzu XD-1 diffractometer (Kyoto, Japan). The presence of functional groups over the catalytic surface was studied by infrared spectroscopy by using a Nicolet Is-10 model, USA. The surface reducibility of the catalyst was studied under temperature programmed reduction under H2-stream (H2-TPR) by using a Chem BET 3000 (Quantachrome equipment, USA). The valance states of the elements and the binding energy of the bond electron were determined by X-ray photoelectron spectroscopy (XPS) by Themo Fisher Scientific, USA. The elemental mapping of catalyst samples was carried out by Energy Dispersive X-ray Spectroscopy instrument (Zeiss Smart EDX, Jena, Germany) equipped with a field-emission scanning electron microscope (FE-SEM; FEI Quanta FEG 250, Hilsboro, OR, USA). The detailed description of instrument and characterization procedure is given in the Supporting Information S6.

3.4. Catalyst Activity Test

Carbon dioxide methanation was carried out in a fixed-bed, tubular quartz silica reactor (length = 50 cm, Inner diameter = 13 mm), equipped with an electric furnace, having a programmable temperature controller. A total of 2.0 g of fresh catalyst was diluted with silicon carbide (to obtain a 5.0 cm bed height and 5 cm3 volume of catalyst bed) and was packed in the middle of the reactor. The temperature of the catalyst bed was monitored by a K-type thermocouple, placed in the center of the catalyst bed. Prior to the reaction, the catalyst sample was reduced in situ with 30 mL/min H2 flow at atmospheric pressure for two hours at 600 °C. After the reduction, the catalysts were cooled down, and CO2:H2:Ar (1:4:5 volume ratio) gas feed was passed through the activated catalyst at GHSV of 6000 ccg−1h−1. Afterwards, the temperature was increased to 350 °C and the methanation reaction was carried out for 13 h. Gaseous reaction products were analyzed online by a quantitative gas analysis system (HIDEN ANALYTICAL QGA, England, UK). The mol % of H2, CO2, CH4, and CO was detected by gas analyzer by using a matrix equation to correct the overlapped detection values from the m/z of 28 (CO), 44 (CO2), 2 (H2), and 16 (CH4). The carbon dioxide conversion ( X CO 2 ), methane selectivity ( S CH 4 ), and carbon monoxide selectivity ( S CO ) were calculated using Equations (1)–(3) [21]:
X CO 2 = CO 2 , in CO 2 , out CO 2 , in × 100
S CH 4 = CH 4 , out CH 4 , out + CO out × 100
S CO = CO out CH 4 , out + CO out × 100

4. Conclusions

A Sm2O3-based Ni-supported catalyst was found excellent for CO2 methanation at 350 °C reaction temperature. The 10Ni/Sm2O3 had the highest CO2 conversion (35%) but needed some time (100 min) for constant high CH4 selectivity (94–95%) over the 13 h reaction period. The 20Ni/Sm2O3 had larger NiO crystallite, which limited the H2 dissociation and thereafter CO2 conversion to 24–26%. However, the slower H2 dissociation and interaction of CO2 over a largely expanded surface showed the highest ~98% CH4 selectivity from the beginning of the reaction. The formate species were stable at the Ni surface, as observed in the infrared spectra, which underwent further hydrogenation to methane. Despite the presence of the largest surface area and highest reducible metal oxide over the surface, 10Co-10Ni/Sm2O3 showed inferior catalytic performance because of diluting Ni sites by Co. On the other hand, diluting the Ni sites by Ru over Sm2O3 (10Ru-10Ni/Sm2O3 catalyst) caused again good catalytic activity. It had RuO2 crystallite, very less surface parameters but easy reducibility. The absence of format species over the 10Ru-10Ni/Sm2O3 catalyst surface indicated an eight-electron reduction of intermediate CO into CH4. In the mean of constant catalytic activity, the 10Ru-10Ni/Sm2O3 catalyst was found the best, as it showed 31–33% CO2 conversion, and 93–94% CH4 selectivity constantly from the beginning to the 13 h time-on-stream.

Supplementary Materials

The following supporting information can be downloaded at: https://0-www-mdpi-com.brum.beds.ac.uk/article/10.3390/catal13010113/s1, Figure S1. EDX elemental analysis of (A) 10Ni/Sm2O3 (B) 20Ni/Sm2O3 (C) 40Ni/Sm2O3 (D) 10Co-10Ni/Sm2O3 (E) 10Ni-10Ru/Sm2O3 catalysts; Figure S2. N2 adsorption isotherm of Sm2O3 and 10Ni/Sm2O3 catalysts; Figure S3. Infrared transmittance spectra of Sm2O3 and 10Ni/Sm2O3 catalysts; Figure S4. Ru3d XPS spectrum of Ru-Ni/Sm2O3 catalyst. Figure S5. The methanation reaction over 10Ni/Sm2O3 at 3-different temperature for 1h at each 250 °C, 300 °C and 350 °C. Supporting Information S6. The detail description of instrument and characterization procedure.

Author Contributions

R.A.E.-S.: Project administration, investigation, methodology, conceptualization, formal analysis; A.S.A.-F.: supervision, data curation, methodology, validation, review, funding acquisition; K.A., W.U.K. and N.S.K.: formal analysis, writing—original draft; A.B. and A.A.I.: investigation, conceptualization, data curation; R.K. and A.A.M.A.: review, editing and supervision. All authors have read and agreed to the published version of the manuscript.

Funding

This research work was funded by Researchers Supporting Project number (RSP2023R368), King Saud University, Riyadh, Saudi Arabia.

Data Availability Statement

No data were used for the research described in the article.

Acknowledgments

The authors would like to extend their sincere appreciation to the Researchers Supporting Project number (RSP2023R368), King Saud University, Riyadh, Saudi Arabia.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Aziz, M.A.A.; Jalil, A.A.; Triwahyono, S.; Ahmad, A. CO2 Methanation over Heterogeneous Catalysts: Recent Progress and Future Prospects. Green Chem. 2015, 17, 2647–2663. [Google Scholar] [CrossRef]
  2. Navarro, J.C.; Centeno, M.A.; Laguna, O.H.; Odriozola, J.A. Policies and Motivations for the CO2 Valorization through the Sabatier Reaction Using Structured Catalysts. A Review of the Most Recent Advances. Catalysts 2018, 8, 578. [Google Scholar] [CrossRef] [Green Version]
  3. Saeidi, S.; Najari, S.; Fazlollahi, F.; Nikoo, M.K.; Sefidkon, F.; Klemeš, J.J.; Baxter, L.L. Mechanisms and Kinetics of CO2 Hydrogenation to Value-Added Products: A Detailed Review on Current Status and Future Trends. Renew. Sustain. Energy Rev. 2017, 80, 1292–1311. [Google Scholar] [CrossRef]
  4. 4Mandal, S.C.; Das, A.; Roy, D.; Das, S.; Nair, A.S.; Pathak, B. Developments of the Heterogeneous and Homogeneous CO2 Hydrogenation to Value-Added C2+-Based Hydrocarbons and Oxygenated Products. Coord. Chem. Rev. 2022, 471, 214737. [Google Scholar] [CrossRef]
  5. Chiang, C.L.; Lin, K.S.; Chuang, H.W. Direct Synthesis of Formic Acid via CO2 Hydrogenation over Cu/ZnO/Al2O3 Catalyst. J. Clean. Prod. 2018, 172, 1957–1977. [Google Scholar] [CrossRef]
  6. Kwak, J.H.; Kovarik, L.; Szanyi, J. CO2 Reduction on Supported Ru/Al2O3 Catalysts: Cluster Size Dependence of Product Selectivity. ACS Catal. 2013, 3, 2449–2455. [Google Scholar] [CrossRef]
  7. Yu, Y.; Chan, Y.M.; Bian, Z.; Song, F.; Wang, J.; Zhong, Q.; Kawi, S. Enhanced Performance and Selectivity of CO2 Methanation over G-C3N4 Assisted Synthesis of Ni–CeO2 Catalyst: Kinetics and DRIFTS Studies. Int. J. Hydrogen Energy 2018, 43, 15191–15204. [Google Scholar] [CrossRef]
  8. 8Hong, J.S.; Hwang, J.S.; Jun, K.W.; Sur, J.C.; Lee, K.W. Deactivation Study on a Coprecipitated Fe-Cu-K-Al Catalyst in CO2 Hydrogenation. Appl. Catal. A Gen. 2001, 218, 53–59. [Google Scholar] [CrossRef]
  9. Bansode, A.; Urakawa, A. Towards Full One-Pass Conversion of Carbon Dioxide to Methanol and Methanol-Derived Products. J. Catal. 2014, 309, 66–70. [Google Scholar] [CrossRef]
  10. Wambach, J.; Baiker, A.; Wokaun, A. CO2 Hydrogenation over Metal/Zirconia Catalysts CO CO Hydrogenation over Metal / Zirconia Catalysts. Phys. Chem. Chem. Phys. 1999, 1, 5071–5080. [Google Scholar] [CrossRef]
  11. 1Ando, H.; Xu, Q.; Fujiwara, M.; Matsumura, Y.; Tanaka, M.; Souma, Y. Hydrocarbon Synthesis from CO2 over Fe-Cu Catalysts. Catal. Today 1998, 45, 229–234. [Google Scholar] [CrossRef]
  12. Kishan, G.; Lee, M.; Nam, S.; Choi, M.; Lee, K. The catalytic conversion of CO2 to hydrocarbons over Fe–K supported on Al2O3–MgO mixed oxides. Catal. Lett. 1998, 56, 215–219. [Google Scholar]
  13. Kim, J.; Lee, S.; Kang, M.; Lee, K.; Choi, M.; Kang, Y. Promotion of CO2 hydrogenation to hydrocarbons in three-phase catalytic (Fe-Cu-K-Al) slurry reactors. Korean J. Chem. Eng. 2003, 20, 967–972. [Google Scholar]
  14. Kim, J.S.; Lee, S.; Lee, S.B.; Choi, M.J.; Lee, K.W. Performance of Catalytic Reactors for the Hydrogenation of CO2 to Hydrocarbons. Catal. Today 2006, 115, 228–234. [Google Scholar] [CrossRef]
  15. Guo, M.; Lu, G. The Regulating Effects of Cobalt Addition on the Catalytic Properties of Silica-Supported Ni–Co Bimetallic Catalysts for CO2 Methanation. React. Kinet. Mech. Catal. 2014, 113, 101–113. [Google Scholar] [CrossRef]
  16. Deng, L.; Liu, X.; Wang, R.; Wang, C.; Zhou, G. Unsupported Ni-Co Alloy as Efficient Catalysts for CO2 Methanation. J. Alloys Compd. 2022, 918, 165472. [Google Scholar] [CrossRef]
  17. Owen, R.E.; O’Byrne, J.P.; Mattia, D.; Plucinski, P.; Pascu, S.I.; Jones, M.D. Cobalt Catalysts for the Conversion of CO2 to Light Hydrocarbons at Atmospheric Pressure. Chem. Commun. 2013, 49, 11683–11685. [Google Scholar] [CrossRef] [Green Version]
  18. Muroyama, H.; Tsuda, Y.; Asakoshi, T.; Masitah, H.; Okanishi, T.; Matsui, T.; Eguchi, K. Carbon Dioxide Methanation over Ni Catalysts Supported on Various Metal Oxides. J. Catal. 2016, 343, 178–184. [Google Scholar] [CrossRef] [Green Version]
  19. Cai, M.; Wen, J.; Chu, W.; Cheng, X.; Li, Z. Methanation of Carbon Dioxide on Ni/ZrO2-Al2O3 Catalysts: Effects of ZrO2 Promoter and Preparation Method of Novel ZrO2-Al2O3 Carrier. J. Nat. Gas Chem. 2011, 20, 318–324. [Google Scholar] [CrossRef]
  20. Aziz, M.A.A.; Jalil, A.A.; Triwahyono, S.; Saad, M.W.A. CO2 methanation over Ni-Promoted Mesostructured Silica Nanoparticles: Influence of Ni Loading and Water Vapor on Activity and Response Surface Methodology Studies. Chem. Eng. J. 2015, 260, 757–764. [Google Scholar] [CrossRef]
  21. Zhao, K.; Wang, W.; Li, Z. Highly Efficient Ni/ZrO2 Catalysts Prepared via Combustion Method for CO2 Methanation. J. CO2 Util. 2016, 16, 236–244. [Google Scholar] [CrossRef]
  22. Taherian, Z.; Yousefpour, M.; Tajally, M.; Khoshandam, B. A Comparative Study of ZrO2, Y2O3 and Sm2O3 Promoted Ni/SBA-15 Catalysts for Evaluation of CO2/Methane Reforming Performance. Int. J. Hydrogen Energy 2017, 42, 16408–16420. [Google Scholar] [CrossRef]
  23. Vahid Shahed, G.; Taherian, Z.; Khataee, A.; Meshkani, F.; Orooji, Y. Samarium-Impregnated Nickel Catalysts over SBA-15 in Steam Reforming of CH4 Process. J. Ind. Eng. Chem. 2020, 86, 73–80. [Google Scholar] [CrossRef]
  24. Bazzi, S.; Schulz, E.; Mellah, M. Electrogenerated Sm(II)-Catalyzed CO2 Activation for Carboxylation of Benzyl Halides. Org. Lett. 2019, 21, 10033–10037. [Google Scholar] [CrossRef]
  25. Michel, C.R.; Martínez-Preciado, A.H.; Parra, R.; Aldao, C.M.; Ponce, M.A. Novel CO2 and CO Gas Sensor Based on Nanostructured Sm2O3 Hollow Microspheres. Sens. Actuators B Chem. 2014, 202, 1220–1228. [Google Scholar] [CrossRef]
  26. Yamasaki, M.; Komori, M.; Akiyama, E.; Habazaki, H.; Kawashima, A.; Asami, K.; Hashimoto, K. CO2 Methanation Catalysts Prepared from Amorphous Ni-Zr-Sm and Ni-Zr-Misch Metal Alloy Precursors. Mater. Sci. Eng. A 1999, 267, 220–226. [Google Scholar] [CrossRef] [Green Version]
  27. Ayub, N.A.; Bahruji, H.; Mahadi, A.H. Barium Promoted Ni/Sm2O3 Catalysts for Enhanced CO2 Methanation. RSC Adv. 2021, 11, 31807–31816. [Google Scholar] [CrossRef]
  28. El-Salamony, R.A.; El-Sharaky, S.A.; Al-Temtamy, S.A.; Al-Sabagh, A.M.; Killa, H.M. CO2 Valorization into Synthetic Natural Gas (SNG) Using a Co–Ni Bimetallic Y2O3 Based Catalysts. Int. J. Chem. React. Eng. 2021, 19, 571–583. [Google Scholar] [CrossRef]
  29. Suslick, K.S.; Skrabalak, S.E. Handbook of Heterogeneous Catalysis; Ertl, G., Knozinger, H., Weitkamp, J., Eds.; Wiley-VCH: Weinheim, Germany, 1997; Volume 4. [Google Scholar]
  30. Yu, Q.; Yao, X.; Zhang, H.; Gao, F.; Dong, L. Effect of ZrO2 Addition Method on the Activity of Al2O3-Supported CuO for NO Reduction with CO: Impregnation vs. Coprecipitation. Appl. Catal. A Gen. 2012, 423–424, 42–51. [Google Scholar] [CrossRef]
  31. El-Salamony, R.A.; El-Sharaky, S.; El-Temtamy, S.; Al-Sabagh, A.; Killa, H. Effect of Ruthenium Promotor Ratio on Ni/Y2O3 Based Catalysts for CO2 Methanation Reaction. Egypt. J. Chem. 2021, 64, 3–4. [Google Scholar] [CrossRef]
  32. Sapkal, P.S.; Kuthe, A.M.; Kashyap, R.S.; Nayak, A.R.; Kuthe, S.A.; Kawle, A.P. Indirect Casting of Patient-Specific Tricalcium Phosphate Zirconia Scaffolds for Bone Tissue Regeneration Using Rapid Prototyping Methodology. J. Porous Mater. 2017, 24, 1013–1023. [Google Scholar] [CrossRef]
  33. Vlasenko, N.V.; Kyriienko, P.I.; Valihura, K.V.; Kosmambetova, G.R.; Soloviev, S.O.; Strizhak, P.E. Yttria-Stabilized Zirconia as a High-Performance Catalyst for Ethanol to n-Butanol Guerbet Coupling. ACS Omega 2019, 4, 21469–21476. [Google Scholar] [CrossRef]
  34. Salinas, D.; Escalona, N.; Pecchi, G.; Fierro, J.L.G. Lanthanum Oxide Behavior in La2O3-Al2O3 and La2O3-ZrO2 Catalysts with Application in FAME Production. Fuel 2019, 253, 400–408. [Google Scholar] [CrossRef]
  35. Lee, S.M.; Lee, Y.H.; Moon, D.H.; Ahn, J.Y.; Nguyen, D.D.; Chang, S.W.; Kim, S.S. Reaction Mechanism and Catalytic Impact of Ni/CeO2- XCatalyst for Low-Temperature CO2 Methanation. Ind. Eng. Chem. Res. 2019, 58, 8656–8662. [Google Scholar] [CrossRef]
  36. Calatayud, M.; Collins, S.E.; Baltanás, M.A.; Bonivardi, A.L. Stability of Formate Species on β-Ga2O3. Phys. Chem. Chem. Phys. 2009, 11, 1397–1405. [Google Scholar] [CrossRef]
  37. Xémard, M.; Goudy, V.; Braun, A.; Tricoire, M.; Cordier, M.; Ricard, L.; Castro, L.; Louyriac, E.; Kefalidis, C.E.; Clavaguéra, C.; et al. Reductive Disproportionation of CO2 with Bulky Divalent Samarium Complexes. Organometallics 2017, 36, 4660–4668. [Google Scholar] [CrossRef]
  38. Al-Fatesh, A.S.; Kumar, R.; Kasim, S.O.; Ibrahim, A.A.; Fakeeha, A.H.; Abasaeed, A.E.; Atia, H.; Armbruster, U.; Kreyenschulte, C.; Lund, H.; et al. Effect of Cerium Promoters on an MCM-41-Supported Nickel Catalyst in Dry Reforming of Methane. Ind. Eng. Chem. Res. 2022, 61, 164–174. [Google Scholar] [CrossRef]
  39. Rahman, S.; Santra, C.; Kumar, R.; Bahadur, J.; Sultana, A.; Schweins, R.; Sen, D.; Maity, S.; Mazumdar, S.; Chowdhury, B. Highly Active Ga Promoted Co-HMS-X Catalyst towards Styrene Epoxidation Reaction Using Molecular O2. Appl. Catal. A Gen. 2014, 482, 61–68. [Google Scholar] [CrossRef]
  40. Hassan, S.; Kumar, R.; Tiwari, A.; Song, W.; van Haandel, L.; Pandey, J.K.; Hensen, E.; Chowdhury, B. Role of Oxygen Vacancy in Cobalt Doped Ceria Catalyst for Styrene Epoxidation Using Molecular Oxygen. Mol. Catal. 2018, 451, 238–246. [Google Scholar] [CrossRef]
  41. Liu, N.; Du, R.J.; Li, W. Hydrogenation CO2 to Formic Acid over Ru Supported on γ-Al2O3 Nanorods. Adv. Mater. Res. 2013, 821–822, 1330–1335. [Google Scholar] [CrossRef]
  42. Löchel, B.P.; Strehblow, H.-H. Breakdown of Passivity of Nickel by Fluoride: II. Surface Analytical Studies. J. Electrochem. Soc. 1984, 131, 713–723. [Google Scholar] [CrossRef]
  43. Mansour, A.N. Nickel Mg Kα XPS Spectra from the Physical Electronics Model 5400 Spectrometer. Surf. Sci. Spectra 1994, 3, 211–220. [Google Scholar] [CrossRef]
  44. Mansour, A.N. Characterization of NiO by XPS. Surf. Sci. Spectra 1994, 3, 231–238. [Google Scholar] [CrossRef]
  45. Khawaja, E.E.; Salim, M.A.; Khan, M.A.; Al-Adel, F.F.; Khattak, G.D.; Hussain, Z. XPS, Auger, Electrical and Optical Studies of Vanadium Phosphate Glasses Doped with Nickel Oxide. J. Non-Cryst. Solids 1989, 110, 33–43. [Google Scholar] [CrossRef]
  46. Li, C.P.; Proctor, A.; Hercules, D.M. Curve fitting analysis of esca Ni 2p spectra of nickel-oxygen compounds and Ni/Al2O3 catalysts. Appl. Spectrosc. 1984, 38, 880–886. [Google Scholar] [CrossRef]
  47. Klein, J.C.; Hercules, D.M. Surface Characterization of Model Urushibara Catalysts. J. Catal. 1983, 82, 424–441. [Google Scholar] [CrossRef]
  48. Mandal, S.; Bando, K.K.; Santra, C.; Maity, S.; James, O.O.; Mehta, D.; Chowdhury, B. Sm-CeO2 Supported Gold Nanoparticle Catalyst for Benzyl Alcohol Oxidation Using Molecular O2. Appl. Catal. A Gen. 2013, 452, 94–104. [Google Scholar] [CrossRef]
  49. Phenomena, R.; V, E.S.P.B.; Uwamino, Y.; Ishizuka, T.; Yamatera, H. X-ray photoelectron spectroscopy of rare-earth compounds. J. Electron Spectrosc. Relat. Phenom. 1984, 34, 67–78. [Google Scholar]
  50. Reddy, B.M.; Chowdhury, B.; Smirniotis, P.G. XPS Study of the Dispersion of MoO3 on TiO2-ZrO2, TiO2-SiO2, TiO2-Al2O3, SiO2-ZrO2, and SiO2-TiO2-ZrO2 Mixed Oxides. Appl. Catal. A Gen. 2001, 211, 19–30. [Google Scholar] [CrossRef]
  51. Zhang, W.D.; Liu, B.S.; Zhan, Y.P.; Tian, Y.L. Syngas Production via CO2 Reforming of Methane over Sm2O3-La2O3-Supported Ni Catalyst. Ind. Eng. Chem. Res. 2009, 48, 7498–7504. [Google Scholar] [CrossRef]
  52. Lucentini, I.; Casanovas, A.; Llorca, J. Catalytic Ammonia Decomposition for Hydrogen Production on Ni, Ru and Ni[Sbnd]Ru Supported on CeO2. Int. J. Hydrogen Energy 2019, 44, 12693–12707. [Google Scholar] [CrossRef]
  53. Wang, Y.; Wang, J.; Xie, T.; Zhu, Q.; Zeng, D.; Li, R.; Zhang, X.; Liu, S. Ru Doping in Ni(OH)2 to Accelerate Water Reduction Kinetics for Efficient Hydrogen Evolution Reaction. Appl. Surf. Sci. 2019, 485, 506–512. [Google Scholar] [CrossRef]
  54. Italiano, C.; Llorca, J.; Pino, L.; Ferraro, M.; Antonucci, V.; Vita, A. CO and CO2 Methanation over Ni Catalysts Supported on CeO2, Al2O3 and Y2O3 Oxides. Appl. Catal. B Environ. 2020, 264, 118494. [Google Scholar] [CrossRef]
  55. Yan, Y.; Dai, Y.; Yang, Y.; Lapkin, A.A. Improved Stability of Y2O3 Supported Ni Catalysts for CO2 Methanation by Precursor-Determined Metal-Support Interaction. Appl. Catal. B Environ. 2018, 237, 504–512. [Google Scholar] [CrossRef]
  56. Falbo, L.; Visconti, C.G.; Lietti, L.; Szanyi, J. The Effect of CO on CO2 Methanation over Ru/Al2O3 Catalysts: A Combined Steady-State Reactivity and Transient DRIFT Spectroscopy Study. Appl. Catal. B Environ. 2019, 256, 117791. [Google Scholar] [CrossRef]
  57. Yin, G.; Yuan, X.; Du, X.; Zhao, W.; Bi, Q.; Huang, F. Efficient Reduction of CO2 to CO Using Cobalt–Cobalt Oxide Core–Shell Catalysts. Chem.-A Eur. J. 2018, 24, 2157–2163. [Google Scholar] [CrossRef]
  58. Guilera, J.; Del Valle, J.; Alarcón, A.; Díaz, J.A.; Andreu, T. Metal-Oxide Promoted Ni/Al2O3 as CO2 Methanation Micro-Size Catalysts. J. CO2 Util. 2019, 30, 11–17. [Google Scholar] [CrossRef]
  59. Kumar, R.; Shah, S.; Bahadur, J.; Melnichenko, Y.B.; Sen, D.; Mazumder, S.; Vinod, C.P.; Chowdhury, B. Highly Stable In-SBA-15 Catalyst for Vapor Phase Beckmann Rearrangement Reaction. Microporous Mesoporous Mater. 2016, 234, 293–302. [Google Scholar] [CrossRef]
Figure 1. Reaction scheme of CO2 and H2 interaction over supported metal catalyst.
Figure 1. Reaction scheme of CO2 and H2 interaction over supported metal catalyst.
Catalysts 13 00113 g001
Figure 2. X-ray diffraction profiles of (A) Sm2O3 and 10Ni/Sm2O3 (B) 10Ni/Sm2O3 and 20Ni/Sm2O3 (C) 20Ni/Sm2O3 and 40Ni/Sm2O3 (D) 20Ni/Sm2O3 and 10Co-10Ni/Sm2O3 (E) 10Ni/Sm2O3 and 10Ru-10Ni/Sm2O3.
Figure 2. X-ray diffraction profiles of (A) Sm2O3 and 10Ni/Sm2O3 (B) 10Ni/Sm2O3 and 20Ni/Sm2O3 (C) 20Ni/Sm2O3 and 40Ni/Sm2O3 (D) 20Ni/Sm2O3 and 10Co-10Ni/Sm2O3 (E) 10Ni/Sm2O3 and 10Ru-10Ni/Sm2O3.
Catalysts 13 00113 g002
Figure 3. (A) N2 adsorption isotherms of the prepared catalysts (B) Table indicating the surface area, pore volume, and pore diameter of the different catalysts (C) Infrared transmittance spectra of the different catalysts (D) H2-temperature programmed reduction profiles of the different catalysts.
Figure 3. (A) N2 adsorption isotherms of the prepared catalysts (B) Table indicating the surface area, pore volume, and pore diameter of the different catalysts (C) Infrared transmittance spectra of the different catalysts (D) H2-temperature programmed reduction profiles of the different catalysts.
Catalysts 13 00113 g003
Figure 4. (A) Ni (2p) core-level XPS spectrum of Co-Ni/Sm2O3 (B) Ni (2p) core-level XPS spectrum of Ru-Ni/Sm2O3 catalyst (C) Sm (3d) XPS spectra of Co-Ni/Sm2O3 and Ru-Ni/Sm2O3 catalyst (D) O (1s) core-level XPS spectra of Co-Ni/Sm2O3 and Ru-Ni/Sm2O3 catalyst.
Figure 4. (A) Ni (2p) core-level XPS spectrum of Co-Ni/Sm2O3 (B) Ni (2p) core-level XPS spectrum of Ru-Ni/Sm2O3 catalyst (C) Sm (3d) XPS spectra of Co-Ni/Sm2O3 and Ru-Ni/Sm2O3 catalyst (D) O (1s) core-level XPS spectra of Co-Ni/Sm2O3 and Ru-Ni/Sm2O3 catalyst.
Catalysts 13 00113 g004
Figure 5. The catalytic activity vs. Time on stream of xNi/Sm2O3 (x = 10, 20, 40) and M-Ni/Sm2O3 (M = Co, Ru) catalysts at 350 °C reaction temperature over CO2:H2:Ar (1:4:5 volume ratio) feed gas: (A) CO2 conversion vs. Time on stream (B) CH4 selectivity vs. Time on stream (C) CO selectivity vs. Time on stream.
Figure 5. The catalytic activity vs. Time on stream of xNi/Sm2O3 (x = 10, 20, 40) and M-Ni/Sm2O3 (M = Co, Ru) catalysts at 350 °C reaction temperature over CO2:H2:Ar (1:4:5 volume ratio) feed gas: (A) CO2 conversion vs. Time on stream (B) CH4 selectivity vs. Time on stream (C) CO selectivity vs. Time on stream.
Catalysts 13 00113 g005
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

El-Salamony, R.A.; Al-Fatesh, A.S.; Acharya, K.; Abahussain, A.A.M.; Bagabas, A.; Kumar, N.S.; Ibrahim, A.A.; Khan, W.U.; Kumar, R. Carbon Dioxide Valorization into Methane Using Samarium Oxide-Supported Monometallic and Bimetallic Catalysts. Catalysts 2023, 13, 113. https://0-doi-org.brum.beds.ac.uk/10.3390/catal13010113

AMA Style

El-Salamony RA, Al-Fatesh AS, Acharya K, Abahussain AAM, Bagabas A, Kumar NS, Ibrahim AA, Khan WU, Kumar R. Carbon Dioxide Valorization into Methane Using Samarium Oxide-Supported Monometallic and Bimetallic Catalysts. Catalysts. 2023; 13(1):113. https://0-doi-org.brum.beds.ac.uk/10.3390/catal13010113

Chicago/Turabian Style

El-Salamony, Radwa A., Ahmed S. Al-Fatesh, Kenit Acharya, Abdulaziz A. M. Abahussain, Abdulaziz Bagabas, Nadavala Siva Kumar, Ahmed A. Ibrahim, Wasim Ullah Khan, and Rawesh Kumar. 2023. "Carbon Dioxide Valorization into Methane Using Samarium Oxide-Supported Monometallic and Bimetallic Catalysts" Catalysts 13, no. 1: 113. https://0-doi-org.brum.beds.ac.uk/10.3390/catal13010113

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