Next Article in Journal
In Situ X-ray Diffraction as a Basic Tool to Study Oxide and Metal Oxide Catalysts
Next Article in Special Issue
Water Oxidation over Au-Pd/TiO2 as a Substitute for Iridium-Based Catalysts
Previous Article in Journal
Propane Dehydrogenation over Cobalt Aluminates: Evaluation of Potential Catalytic Active Sites
Previous Article in Special Issue
Sn-Doped Hematite Films as Photoanodes for Photoelectrochemical Alcohol Oxidation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

The Effectiveness of Ni-Based Bimetallic Catalysts Supported by MgO-Modified Alumina in Dry Methane Reforming

by
Ahmed A. Ibrahim
1,*,
Anis H. Fakeeha
1,
Ahmed E. Abasaeed
1,
Irfan Wazeer
1,
Abdulaziz Bentalib
1,
Nadavala Siva Kumar
1,
Jehad K. Abu-Dahrieh
2,* and
Ahmed S. Al-Fatesh
1,*
1
Chemical Engineering Department, College of Engineering, King Saud University, Riyadh 11421, Saudi Arabia
2
School of Chemistry and Chemical Engineering, Queen’s University Belfast, Belfast BT9 5AG, Northern Ireland, UK
*
Authors to whom correspondence should be addressed.
Catalysts 2023, 13(11), 1420; https://doi.org/10.3390/catal13111420
Submission received: 9 September 2023 / Revised: 26 October 2023 / Accepted: 31 October 2023 / Published: 7 November 2023
(This article belongs to the Special Issue Feature Papers in Catalytic Materials)

Abstract

:
Syngas is produced through the carbon dioxide reforming of methane. The traditional nickel-based catalysts are substantially destroyed by carbon deposition. The reforming reaction was conducted in a tubular microreactor at 700 °C using bimetallic Ni catalysts supported over 37% Al2O3 and 63% MgO mixtures. The impregnation process formed the catalysts, which were subsequently examined by N2-physisorption, XRD, H2-TPR, TGA, and Raman spectroscopy. The 2.5Ni+2.5Co/37%Al2O3+63%MgO bimetallic catalyst, which displayed 72% and 76% conversions of CH4 and CO2 over the course of a seven-hour procedure, was discovered to be the most active in DRM. The bimetallic catalyst with the largest weight loss in TGA, 2.5Ni+2.5Fe-MG63, had a loss of 61.3%, a difference of 26% and 21% in the activity performance of CH4 and CO2, respectively, of the tested bimetallic Ni catalysts was recorded. The long-time of 30 h on-stream CH4 and CO2 conversion reactions for 2.5Ni+2.5Co-MG63 and 2.5Ni+2.5Ce-MG63 catalysts showed the catalysts’ high stability. The TPO analysis for the 2.5Ni+2.5Cs-MG63 catalyst showed a peak at 650 °C, attributed to the oxidation of the filamentous carbon, whereas the TPO analysis for the 2.5Ni+2.5Co-MG63 catalyst depicted a peak at 540 °C, ascribed to the presence of amorphous/graphite carbon.

1. Introduction

Massive interest in the dry reforming of methane (DRM) has been sparked by the depletion of fossil fuels and the chemical recycling of CH4 and CO2 into fuels or chemicals [1,2,3]. This is because DRM has many benefits, including directly utilizing natural-gas resources [4] and converting greenhouse gases CH4 and CO2 into synthesis gas [4,5,6,7]. The greenhouse gases are responsible for global warming a major factor in climate change [8]. The DRM reaction Equation (1) generates an H2/CO ratio of unity, while the reforming of methane Equation (2) produces an H2/CO ratio of 3. The partial oxidation of methane Equation (3) generates an H2/CO ratio of 2.
C H 4 + C O 2 2 H 2 + 2 C O       Δ H 298 = 247   k J / m o l
  C H 4 + H 2 O 3 H 2 + C O       Δ H 298 = 206   k J / m o l
    C H 4 + 0.5 O 2 2 H 2 + CO       Δ H 298 = 36   k J / m o l
For the Fischer–Tropsch process, which produces long-chain hydrocarbons and oxygenated molecules, a unity H2/CO ratio is optimal [9,10]. The DRM has advantages over other reforming reactions, but it also has several disadvantages, which limit its use on a broad scale. Due to CO2’s thermodynamic stability, it requires considerable energy to oxidize it and create a thermodynamic driving force that produces high yields of H2 and CO [11]. The H2/CO ratio falls below unity as a result of the reverse water–gas shift reaction in Equation (4), which also consumes the hydrogen created [12]. In addition, there are side reactions that cause the catalysts to lose their effectiveness. For instance, the Boudouard reaction Equation (5) reduces the CO to carbon, while the CH4 cracking reaction Equation (6) creates carbon [13,14]. Carbon is produced in both processes, which is detrimental to the efficiency of the dry-reforming catalysts. Both reactions result in the creation of carbon, which is harmful to the effectiveness of the dry-reforming catalysts.
H 2 + C O 2 H 2 O + C O       Δ H 298 = 41   k J / m o l
2 C O C O 2 + C       Δ H 298 = 131   k J / m o l
C H 4 2 H 2 + C       Δ H 298 = 75   k J / m o l
The Ni-based catalysts are found to be promising because they are affordable and readily available [15,16]; however, sintering and coke deposition caused by side reactions, such as CO disproportionation and CH4 cracking, are drawbacks for Ni catalysts [17]. The catalyst structure can be changed to prevent the creation of carbon. Utilizing supports with strong mechanical properties like alumina, which remain stable at higher temperatures, can endure sintering and suppression [18]. A mixture of Al2O3 and MgO supported by Ni, according to a study, inhibits the deposition of carbon [19]. It was also well known that utilizing MgO with basic properties as a support modifier would decrease carbon deposition and increase catalytic stability. The combination of an element with Ni forms bimetallic features, and it was observed that this often acts positively and has therefore attracted wide attention in many eco-friendly and energy-status catalytic applications [20].
Bimetallic catalysts have thus far attracted significant attention, due to the influence of the electronic synergy effect. Because of the increased CO2 dissociation and metal reducibility, a Ni-Zn/ZrO2 catalyst is much more active and stable at different temperatures in the dry reforming of the methane reaction [21]. The carbon deposition in the DRM reaction is improved by the common bimetallic catalysts, such as Ni-Co [17,18], Ni-Fe [19,20,21,22], Ni-Cu [23], Ni-In [24], Ni-Mo [25], and Ni-Sn [26]. The main strategy to suppress carbon deposition is to increase the concentration of lattice O* on the catalyst surface, which removes filamentous and amorphous carbon. In order to avoid further graphitic carbon production, it is necessary to increase the loading of doped metals, thereby increasing the energy barrier of C–H bond breaking and slowing down the reaction rate. However, a decrease in reaction rate makes the work less efficient, even to stagnate. Various studies indicate that Co can show a nickel-like activity, forming a Ni-Co bimetallic catalyst that exhibits great stability and reduction and effectively suppresses the carbon formation compared to the Ni and Co monometallic catalysts [27,28]. By adjusting the Ni-Co molar ratio, the catalyst has been shown to have reduced carbon accumulation deactivation and enhanced catalytic activity, particularly when the ratio is around 3:1. For instance, the addition of Mo with Ni improves catalytic conversion yields of DRM [22]. Therefore, it could be a good strategy to design bimetallic systems by adding two active metal phases to eliminate significantly the coking and sintering drawbacks that were encountered when a single-element DRM catalyst is employed [20,23]. The synergistic properties of bimetallic systems play a critical part in the catalytic performance and attainment of energy storage of these hybrid materials [23]. Numerous types of research on the DRM reaction have stated that Ni forms an alloy with highly active and carbon-resistant metals, among which Co is an effective option. Indeed, doping Co in the Ni-Co bimetallic catalyst displays high catalytic performance while reducing carbon formation, owing to its high oxygen affinity [24]. Alloying Ni with a second metal was naturally assumed to stabilize metal nanoparticles. Bimetallic Ni-Fe systems have been broadly investigated for DRM [26]. Theofanidis et al. investigated the DRM stability and impact of Fe in alloying with Ni, and confirmed that a high selectivity toward CO and a CO/H2 ratio close to unity can be achieved [27].
Commonly, cerium oxide advances catalyst stability, enhances metal dispersion, and upgrades CO2 activation. Cerium element is a superb promoter for catalytic applications, because of the interconversion capacity of Ce3+/Ce4+, which permits an enhanced oxygen mobility [28]. Some researchers have noted that perovskite oxides containing Sr are resistant to coke deposition in dry reforming; the Sr-doped catalysts exhibit upgraded catalytic performance owing to high oxygen vacancy and activation of the C–H bond [29]. For the purpose of evaluating the activities of the catalysts supported over 63%MgO+Al2O3 in the DRM reaction, this paper will study catalysts’ chemical, physical, and structural features. For the bimetallic catalysts, a weight loading of 2.5 wt.% was selected for each element. The potential for coking resistance in DRM, the function of bimetallic in optimizing catalytic efficiency, and the production of the synthesis gas are investigated. Various characterization techniques, including N2 adsorption–desorption, TPR, XRD, Raman, and TGA, were employed to analyze the catalysts.

2. Results

2.1. Surface Description

Table 1 provides a summary of the findings from the BET analysis of the bimetallic samples. The bimetallic samples exhibited similar surface areas and pore volumes, except for the 2.5Ni+2.5Sr-MG63 sample. There is not much variation in the pore size. Figure 1A displays the N2 adsorption–desorption isotherms. The N2 physisorption plots in Figure 1A were classified as type IV isotherms by the IUPAC, indicating mesoporous materials with pores that range between 2 and 50 nm diameters, and they have an H3-type adsorption hysteresis loop as a result of slit-shaped pores, which allowed the multilayer of adsorbed N2 molecules to grow and pack more tightly together [30]. Figure 1B shows mono-modal pore width distribution by BJH Adsorption dV/dlog (D) pore volume method. It confirmed the mesoporosity of the catalysts. The peak maxima lay between 3.7 and 4.0 nm. Table 1 shows the surface area, pore volume, and pore diameters of the bimetallic sample. The 2.5Ni+2.5Sr-MG63 sample assumed the lowest surface area and the lowest pore volume, while the remaining samples displayed similar values.

2.2. TPR

Temperature programmed reduction (H2-TPR) was used to test the reducibility of metals to disclose the interaction between the metals and the impact of the reduction process on the structure and composition of the catalysts. Figure 2 displays the H2-TPR of the bimetallic catalysts. All the catalysts show two prominent peaks. These reduction peaks at different temperatures indicate the extent of the interaction of the active metals with the support. All the catalysts show peaks at about 945 °C, which can be associated with the metallic exsolution from the stable phases such as NiAl2O4. The 2.5Ni+2.5Fe-MG63 catalyst has a broad peak at 535 °C, which can be ascribed to the strong interaction of oxides of Ni and Fe with the support. Similarly, in the same region at 686 °C, the 2.5Ni+2.5Ce-MG63 and 2.5Ni+2.5Sr-MG63 show small peaks. On the other hand, the 2.5Ni+2.5Co-MG63 displays a peak at 342 °C, which is characterized by free oxides of Ni and Co on the surface of the support. Because of this type of interaction, the catalyst depicted the best activity. A negative TPR peak at around 200 °C was noticed, and it was determined that it was the result of the metal hydride, which had been created at ambient temperature as a result of the atomic hydrogen being absorbed inside the structure of massive metallic particles. Table 2 shows the quantity of H2 consumption during the TPR analysis. The 2.5Ni+2.5Co-MG63 sample assumed the highest hydrogen consumption after the 2.5Ni+2.5Fe-MG63 sample. This made the 2.5Ni+2.5Co-MG63 sample the optimum. Because of the presence of several oxidation states of iron, the 2.5Ni+2.5Fe-MG63 sample consumed the highest amount of H2 to accomplish the stepwise reduction of Fe2O3 to Fe through Fe3O4 and FeO.

2.3. XRD

To examine the crystalline phases in the bimetallic samples, powder X-ray diffraction was employed. Figure 3 shows the diffraction angles for 2.5Ni+2.5Co-MG63, 2.5Ni+2.5Ce-MG63, 2.5Ni+2.5Fe-MG63, and 2.5Ni+2.5Sr-MG63. The 2.5Ni+2.5Co-MG63 catalyst had diffraction peaks at 2θ = 36.5°, 43.3°, and 63.0°. The peak at 2θ = 36.5° corresponded to the (111) plane of CoO (JCPDS card, No. 01-075-0419), while the 2.5Ni+2.5Ce-MG63 catalyst had diffraction peaks at 2θ = 28.6°, 33.0°, 47.4°, 56.1°, and 63.0°. The diffraction peaks at 2θ = 28.6°, 33.0°, 47.4°, and 56.1° were ascribed to CeO2 (JCPDS 34-0394), while the 2.5Ni+2.5Fe-MG63 and 2.5Ni+2.5Sr-MG63 catalysts had peaks at 2θ = 35.5°, 43.3°, and 63.0°. The peak at 2θ = 35.5° corresponds to pure magnetite (Fe3O4) (JCPDS card NO. 01-075-0449). In all the catalysts, two diffraction peaks appearing at 2θ = 43.3°and 63.0°could be attributed, respectively, to (113), and (214) crystallographic planes of the cubic γ-alumina phase (JCPDS No. 42-1468).

2.4. TPD

The CO2-TPD profile presented in Figure 4 may be used to understand the catalyst’s fundamental profile as well as its capacity for CO2 adsorption. Three zones are identified in the CO2-TPD profile of a catalyst: the low-temperature region (100–150 °C) for weak Brønsted basic sites such as the surface OH group and the intermediate temperature region (200–450 °C) for the medium-strength Lewis base sites, respectively [31]. The CO2-TPD of the bimetallic catalysts displayed broad and intense peaks in the low to intermediate temperature region. This indicated that the bimetallic catalysts possessed a number of weak- and moderate-strength basic sites. The Sr and Fe bimetallic Ni catalysts have a higher number of basic sites compared to the Co and Ce bimetallic Ni catalysts. The lesser activity performance observed with the Fe and Sr bimetallic Ni catalysts was due to their excessive basicity, which increased the coke deposition. The amount of CO2 desorbed during the TPD analysis is shown in Table 3. The results show that the Sr and Fe bimetallic Ni catalysts, which exhibit lower activity than the Co and Ce bimetallic Ni catalysts, were more abundant.
Raman spectroscopy was used to characterize the structure of the samples. All the Raman spectra displayed the absence of peaks related to carbon, which appears beyond 1000 cm−1. The Raman shift analysis of the fresh samples of bimetallic catalysts was displayed in Figure 5. The peaks that appeared around the Raman shift value of 500 cm−1 were attributable to NiO [32], while the peaks at around 650–700 and 1000 cm−1 were ascribed to Co, Fe and Sr respectively. The peak at at 283 cm−1 for the 2.5Ni+2.5Fe-MG63 catalyst was attributable to the Fe3O4 [33], while for the 2.5Ni+2.5Co-MG63 catalyst the peak at 282 cm−1 was ascribed to the F2g vibrational modes of the crystalline phase of CO3O4 [34]. The high-intensity peak of 2.5Ni+2Ce-MG63 sample was due to the overlapping effect of Ce and NiO. Ni sites are diluted by MgO or another metal oxide, and the Raman peaks are shifted toward a lower frequency [35,36]. So, the shift of the Raman band of NiO was due to the contribution of nearby metal oxides. For Ni-Ce, the appearance of the peak in the range 550–560 cm−1 is easily observable. However, the Raman band is shifted to 476 cm−1 upon incorporation of Fe, suggesting that Fe was inserted into the lattice of NiO. On the other hand, for the Ni-Co, the added Co could form a solid solution as Ni(Co)O. In this case, a broad band around 580 cm−1 assigned to the Co2+−O stretching mode appeared, but this signal was overlapped by those of NiO; similarly, the interaction between the NiO and that of the SrO resulted in the appearance of a peak at 559 cm−1.

2.5. SEM

SEM analysis was used to evaluate the morphological aspects of the samples. The SEM images of the 2.5Ni+2.5Ce-MG63, 2.5Ni+2.5Co-MG63, and 2.5Ni+2.5Sr-MG63 catalysts are presented in Figure 6. From the figures, quite homogeneous distributions of oxide particles were observed in fresh catalysts, while for the used catalysts, aggregation of clusters due to coke and sintering after the reaction (7 h) was displayed.

2.6. TEM

Figure 7 shows the fresh and the spent TEM images of the 2.5Ni+2.5Co-MG63 catalyst. The image of the fresh catalyst displays the good distribution of the active metal particles, while the used TEM shows the formation of slight graphitic carbon, as depicted by the TPO characterization analysis.

2.7. Catalytic Activity

Before evaluating commercially produced catalysts, a blank experiment to test catalytic performance was conducted using an empty stainless-steel reactor without catalysts under the same feed ratio and reaction temperature of 700 °C. The CH4 and CO2 conversions of 1.5% and 0.3%, respectively, were obtained; therefore, the metallic reactor had an insignificant effect. The CH4 and CO2 conversions are illustrated in Figure 8. The 2.5Ni+2.5Co-MG63 sample gave a conversion of CH4 about 74% higher than the other bimetallic samples. The 2.5Ni+2.5Co-MG63 sample also gave the highest CO2 conversion, of about 78%, whereas the 2.5Ni+2.5Ce-MG63 displayed higher CO2 conversion values, of about 74%, than 2.5Ni+2.5Fe-MG63 and 2.5Ni+2.5Sr-MG63. The lowest activity performance was recorded by the 2.5Ni+2.5Sr-MG63. In all the catalyst samples, the conversion values of CO2 are higher than the corresponding CH4 value. The side reaction of the reverse water–gas shift may have contributed to this outcome by reacting with the produced H2, as depicted in Equation (4) [37]. The active site of DRM is metallic Ni. 2.5Ni2.5Fe/Mg63 catalyst has the highest consumption of H2 and the highest interaction of CO2 but the DRM activity of 2.5Ni2.5Fe/Mg63 is less than the 2.5Ni2.5Co/Mg63. This indicates that, despite good interaction with CO2, the catalytic active sites are not much exposed for excelling the DRM due to the shading of Ni by Fe. 2.5Ni2.5Co/Mg63 has relatively less H2-consumption and CO2 consumption (less reducibility and less extent of interaction with CO2) than 2.5Ni2.5Fe/Mg63 but the DRM activity of the catalyst is highest. This indicates that, with a combination of Co, the catalytic active sites Ni are well exposed, and it excels in the DRM reaction even though it has lower number of active sites than 2.5Ni2.5Fe/Mg63. The reducible species over 2.5Ni2.5Ce/Mg63 and 2.5Ni2.5Sr/Mg63 catalysts are much lower than 2.5Ni2.5Fe/Mg63 and 2.5Ni2.5Co/Mg63 catalysts. So, the catalytic activities of these catalysts toward DRM are inferior.

2.8. TGA

The amount of carbon that had been deposited was calculated using the thermogravimetric analysis. Figure 9 displays the weight loss of the carbon deposits on the catalyst samples as a fraction of the total weight loss. The catalyst sample 2.5Ni+2.5Fe-MG63 indicated a maximum weight loss of 61.3% at a reaction temperature of 700 F and a reaction duration of 7 h, followed by 2.5Ni+2.5Co-MG63, 2.5Ni+2.5Sr-MG63, and 2.5Ni+2.5Ce-MG63 catalysts, which produced 58%, 27.2%, and 23.7% weight loss, respectively. The 2.5Ni+2.5Co-MG63 catalyst had a significant weight loss of carbon deposited on it, which was in perfect agreement with the catalyst’s high activity.
For verification of stability and suitability for industrial operation, the long-time on-streams of CH4 and CO2 conversion reactions for 2.5Ni+2.5Co-MG63 and 2.5Ni+2.5Ce-MG63 catalysts were performed for 30 h, as shown in Figure 10. The result displays the catalysts maintained their activities with negligible deactivation and hence established the good stability of the catalysts. The TPO analysis for the same catalysts after the 30 h reaction at 700 °C temperature was carried out and depicted in Figure 11. The TPO was done to ascertain the overall quantity and reactivity of the carbon deposit on the surface of catalysts. The 2.5Ni+2.5Co-MG63 catalyst’s TPO curve showed the existence of a peak with a maximum temperature of 540 °C, assigned to the presence of amorphous/graphite carbon [38], while the 2.5Ni+2.5Cs-MG63 catalyst showed the presence of a peak with a maximum temperature of 650 °C, attributed to the oxidation of the filamentous carbon. Table 4 compares the effectiveness of the present work to those in the past. The result show that a good catalyst is developed.

3. Materials

All the chemicals were purchased from commercial sources and used without further treatment. Nickel nitrate hexahydrate Ni(NO3)2·6H2O, cobalt nitrate hexahydrate Co(NO3)2·6H2O, iron nitrate nonahydrate Fe(NO3)3·9H2O, cerium nitrate hexahydrate Ce(NO3)3·6H2O, and strontium nitrate hexahydrate Sr(NO3)2·6H2O were purchased from Alfa Aesar (Heysham, Britain). 63%MgO+37%Al2O3 support was obtained as a gift from SASOL Anckelmannsplatz 1 (Hamburg, Germany). deionized water, crucible, and crusher.

3.1. Preparation of Catalysts

2.5 wt.% of Ni and 2.5 wt.% Co, 2.5 wt.% Ce, 2.5 wt.% Fe, and 2.5 wt.% Sr were synthesized to form bimetallic Ni by a simple wet impregnation method. 0.95 g of support was dispersed in 100 mL of deionized water at room temperature for 20 min. Then, appropriate amounts of nickel nitrate hexahydrate, cobalt nitrate hexahydrate, iron nonahydrate, cerium nitrate hexahydrate, or strontium nitrate hexahydrate were added to form bimetallic compounds of Ni, and the output product was stirred at 80 °C until dry. The obtained solid sample was first calcined in air at 600 °C for 3 h.

3.2. Catalytic Evaluation

In a continuous-flow fixed-bed reactor (30 cm long and 0.94 cm in diameter) operating at 1 atm, the catalyst’s activity (0.1 g) was measured. A thermocouple was fastened to the catalyst’s central bed in order to gauge the reaction’s temperature. Before the response began, H2 was added to the catalysts at a flow rate of 30 mL/min for an hour at 700 °C. The reactor was then purged with flowing N2 at a rate of 20 mL/min for 20 min at 700 °C to eliminate any leftover H2. Before the N2 purge was halted, the reactor temperature was allowed to rise to 800 °C. CH4/CO2/N2 was the input gas. Throughout the process, the flow was maintained at 70 mL/min. with a space velocity of 42,000 mL/hgcat. A 30/30/10 volume ratio was used. To evaluate the concentration of the reactor’s products and unconverted feed gases, a thermal conductivity detector and an online GC (GC-Shimadzu 2014, Kyoto, Japan) outfitted with two columns (Porapak Q and Molecular Sieve 5A) were coupled to the reactor. The effectiveness of the activity was assessed using Equations (7) and (8).
  M e t h a n e   c o n v e r s i o n % = C H 4 , i n C H 4 , o u t C H 4 , i n × 100
C a r b o n   d i o x i d e   c o n v e r s i o n % = C O 2 , i n C O 2 , o u t C O 2 , i n × 100

3.3. Catalyst Characterization

The specific surface areas of the catalysts were measured using nitrogen physisorption at 196 °C. Using a Micromeritics Tristar II 3020 instrument (Micromeritics, Atlanta, GA, USA), the surface area (BET) was calculated using the Brunauer–Emmett–Teller technique (Micromeritics, Atlanta, GA, USA). The quantity of carbon deposited on spent was evaluated using a Shimadzu thermal gravimetric analysis (TGA) system in air, catalysts Shimadzu Corporation’s TGA-51 (Shimadzu Corporation, Kyoto, OP, Japan). Using a laser Raman (NMR-4500-JASCO, Tokyo, Japan) spectrometer, the level of graphitization and the kind of carbon over the catalysts were assessed. The excitation laser utilized had a wavelength of 532 nm. The structure of the employed materials was documented using a TEM (transmission electron microscope) (120 kV JEOL JEM-2100F-Akishima, Tokyo, Japan). TEM micrographs were captured at 120 kV. To investigate the crystal phases and the structure of the fresh catalysts, an X-ray diffractometer was employed. A Miniflex Rigaku diffractometer (Rigaku, Bahrain, Saudi Arabia) was used for the experiment, producing Cu Ka X-ray radiation while operating at 40 kV and 40 mA. A 2 h angle span of 10–85 and a step magnitude of 0.01 were used to collect the data. Micromeritics Auto Chem II 2920 (H2-TPR) equipment (Micromeritics, Atlanta, GA, USA) was used to do the temperature-programmed reduction. The catalysts were heated for an hour at 200 °C in an argon environment before the tests, and they were then cooled to room temperature. For H2-TPR, 0.07 g of the sample was heated to 1000 °C and passed through an H2/Ar (v/v, 10/90) gas mixture at a rate of 40 mL/min. The structure of the produced catalysts was investigated via SEM analysis utilizing a Philips XL-30-FEG equipment (Eindhoven, The Netherlands). Utilizing the Micromeritics Auto Chem II 2920 (Atlanta, GA, USA), temperature-programmed oxidation (TPO) analysis was carried out. Prior to cooling to room temperature, the used catalyst was pre-treated with argon at 150 °C for 30 min. The catalyst’s temperature was then increased from 50 to 1000 °C while being exposed to a 10% oxidizing gas/He combination (flow rate of 40 mL/min).

4. Conclusions

Syngas was produced through the CO2 reforming of methane (DRM) using bimetallic Ni catalysts supported on 63% MgO and 37% alumina. Co, Ce, Fe, and Sr are the bimetallic Ni catalysts tested. The catalyst 2.5Ni+2.5Co-MG63 produced the maximum catalytic activity for DRM. Over seven hours, there were 73% and 76%, respectively, CH4 and CO2 conversions. The catalyst gave the highest weight loss, of 58%, after the 2.5Ni+2.5Fe-MG63. This could be related to the high activity of the sample. The difference of activity performance of CH4 and CO2 of the tested bimetallic Ni catalysts generated values of 26% and 21% respectively. The long-time on stream of CH4 and CO2 conversion reactions for 2.5Ni+2.5Co-MG63 and 2.5Ni+2.5Ce-MG63 catalysts displayed that the catalysts upheld their activities with negligible deactivation and hence established he good stability of the catalysts. The TPO analysis of spent 2.5Ni+2.5Co-MG63 and 2.5Ni+2.5Ce-MG63 catalysts after 30 h reaction at 700 °C temperature determined the overall quantity and reactivity of the carbon deposit on the surface of catalysts: amorphous/graphite carbon for the 2.5Ni+2.5Co-MG63 catalyst and filamentous carbon for 2.5 Ni + 2.5 Cs-MG63 catalyst.

Author Contributions

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

Funding

Deputyship for Research and Innovation, “Ministry of Education” in Saudi Arabia for funding this research (IFKSUOR3-544-1), King Saud University, Riyadh, Saudi Arabia.

Data Availability Statement

Not applicable.

Acknowledgments

The authors extend their appreciation to the Deputyship for Research and Innovation, “Ministry of Education” in Saudi Arabia for funding this research (IFKSUOR3-544-1).

Conflicts of Interest

The authors declare that they have no competing interests.

References

  1. Yan, X.; Hu, T.; Liu, P.; Li, S.; Zhao, B.; Zhang, Q.; Jiao, W.; Chen, S.; Wang, P.; Lu, J.; et al. Highly Efficient and Stable Ni/CeO2-SiO2 Catalyst for Dry Reforming of Methane: Effect of Interfacial Structure of Ni/CeO2 on SiO2. Appl. Catal. B Environ. 2019, 246, 221–231. [Google Scholar] [CrossRef]
  2. Cao, P.; Adegbite, S.; Zhao, H.; Lester, E.; Wu, T. Tuning Dry Reforming of Methane for F-T Syntheses: A Thermodynamic Approach. Appl. Energy 2018, 227, 190–197. [Google Scholar] [CrossRef]
  3. Usman, M.; Daud, W.M.A.W.; Abbas, H.F. Dry Reforming of Methane: In Fl Uence of Process Parameters—A Review. Renew. Sustain. Energy Rev. 2015, 45, 710–744. [Google Scholar] [CrossRef]
  4. Amin, R.; Liu, B.; Huang, Z.B.; Zhao, Y.C. Hydrogen and Syn Gas Production via CO2 Dry Reforming of Methane over Mg/La Promoted Co-Ni/MSU-S Catalyst. Int. J. Hydrogen Energy 2016, 41, 807–819. [Google Scholar] [CrossRef]
  5. Al-Fatesh, A.S.; Atia, H.; Ibrahim, A.A.; Fakeeha, A.H.; Singh, S.K.; Labhsetwar, N.K.; Shaikh, H.; Qasim, S.O. CO2 Reforming of CH4: Effect of Gd as Promoter for Ni Supported over MCM-41 as Catalyst. Renew. Energy 2019, 140, 658–667. [Google Scholar] [CrossRef]
  6. Furkan, H.B.; Rakibul Hasan, K.M.; Uddin, M.J. Greenhouse Gas Emission, GDP, Tertiary Education, and Rule of Law: A Comparative Study between High-Income and Lower-Middle Income Countries. Heliyon 2023, 9, e16265. [Google Scholar] [CrossRef] [PubMed]
  7. Hamza Fakeeha, A.; Sadeq Al-Fatesh, A.; Aidid Ibrahim, A.; Elhag Abasaeed, A. CO2 Reforming of CH4 over Ni-Catalyst Supported on Yttria Stabilized Zirconia. J. Saudi Chem. Soc. 2021, 25, 101244. [Google Scholar] [CrossRef]
  8. Yentekakis, I.V.; Dong, F. Grand Challenges for Catalytic Remediation in Environmental and Energy Applications Toward a Cleaner and Sustainable Future. Front. Environ. Chem. 2020, 1, 1–14. [Google Scholar] [CrossRef]
  9. Zhang, M.; Liu, J.; Zhang, Y.H.; Wang, L.; Li, J.L.; Hong, J.P. Preparation of Highly Dispersed Silicon Spheres Supported Cobalt-Based Catalysts and Their Catalytic Performance for Fischer-Tropsch Synthesis. J. Fuel Chem. Technol. 2023, 51, 608–615. [Google Scholar] [CrossRef]
  10. Pratschner, S.; Hammerschmid, M.; Müller, S.; Winter, F. Evaluation of CO2 Sources for Power-to-Liquid Plants Producing Fischer-Tropsch Products. J. CO2 Util. 2023, 72, 102508. [Google Scholar] [CrossRef]
  11. Aramouni, N.A.K.; Touma, J.G.; Tarboush, B.A.; Zeaiter, J.; Ahmad, M.N. Catalyst Design for Dry Reforming of Methane: Analysis Review. Renew. Sustain. Energy Rev. 2018, 82, 2570–2585. [Google Scholar] [CrossRef]
  12. Nikoo, M.K.; Amin, N.A.S. Thermodynamic Analysis of Carbon Dioxide Reforming of Methane in View of Solid Carbon Formation. Fuel Process. Technol. 2011, 92, 678–691. [Google Scholar] [CrossRef]
  13. Al-Zahrani, S.A.; Al-Fatesh, A.S.; Kaydouh, M.N.; Al Otaibi, A.; Francesco, F.; Fakeeha, A.H.; El Hassan, N. High Carbon-Resistant Nickel Supported on Yttria–Zirconia Catalysts for Syngas Production by Dry Reforming of Methane: The Promoting Effect of Cesium. Alexandria Eng. J. 2023, 74, 371–386. [Google Scholar] [CrossRef]
  14. Abasaeed, A.E.; Al-Fatesh, A.S.; Naeem, M.A.; Ibrahim, A.A.; Fakeeha, A.H. Catalytic Performance of CeO2 and ZrO2 Supported Co Catalysts for Hydrogen Production via Dry Reforming of Methane. Int. J. Hydrogen Energy 2015, 40, 6818–6826. [Google Scholar] [CrossRef]
  15. Muraleedharan Nair, M.; Kaliaguine, S. Structured Catalysts for Dry Reforming of Methane. New J. Chem. 2016, 40, 4049–4060. [Google Scholar] [CrossRef]
  16. Abdullah, B.; Abd Ghani, N.A.; Vo, D.V.N. Recent Advances in Dry Reforming of Methane over Ni-Based Catalysts. J. Clean. Prod. 2017, 162, 170–185. [Google Scholar] [CrossRef]
  17. Palanichamy, K.; Umasankar, S.; Ganesh, S.; Sasirekha, N. Highly Coke Resistant Ni-Co/KCC-1 Catalysts for Dry Reforming of Methane. Int. J. Hydrogen Energy 2023, 48, 11727–11745. [Google Scholar] [CrossRef]
  18. Almeida, C.M.R.; Ghica, M.E.; Durães, L. An Overview on Alumina-Silica-Based Aerogels. Adv. Colloid Interface Sci. 2020, 282, 102189. [Google Scholar] [CrossRef]
  19. Wang, S.; Lu, G.Q.M. CO2 Reforming of Methane on Ni Catalysts: Effects of the Support Phase and Preparation Technique. Appl. Catal. B Environ. 1998, 16, 269–277. [Google Scholar] [CrossRef]
  20. Bian, Z.; Das, S.; Wai, M.H.; Hongmanorom, P.; Kawi, S. A Review on Bimetallic Nickel-Based Catalysts for CO2 Reforming of Methane. ChemPhysChem 2017, 18, 3117–3134. [Google Scholar] [CrossRef]
  21. Velisoju, V.K.; Virpurwala, Q.J.S.; Attada, Y.; Bai, X.Q.; Davaasuren, B.; Hassine, M.B.; Yao, X.L.; Lezcano, G.; Kulkarni, S.R.; Castano, P. Overcoming the kinetic and deactivation limitations of Ni catalyst by alloying it with Zn for the dry reforming of methane. J. CO2 Util. 2023, 75, 102573. [Google Scholar] [CrossRef]
  22. Gaillard, M.; Virginie, M.; Khodakov, A.Y. New Molybdenum-Based Catalysts for Dry Reforming of Methane in Presence of Sulfur: A Promising Way for Biogas Valorization. Catal. Today 2017, 289, 143–150. [Google Scholar] [CrossRef]
  23. De, S.; Zhang, J.; Luque, R.; Yan, N. Ni-Based Bimetallic Heterogeneous Catalysts for Energy and Environmental Applications. Energy Environ. Sci. 2016, 9, 3314–3347. [Google Scholar] [CrossRef]
  24. Hu, Y.H.; Ruckenstein, E. Comment on “Dry Reforming of Methane by Stable Ni-Mo Nanocatalysts on Single-Crystalline MgO”. Science 2020, 368, 777–781. [Google Scholar] [CrossRef] [PubMed]
  25. Luisetto, I.; Tuti, S.; Di Bartolomeo, E. Co and Ni Supported on CeO2 as Selective Bimetallic Catalyst for Dry Reforming of Methane. Int. J. Hydrogen Energy 2012, 37, 15992–15999. [Google Scholar] [CrossRef]
  26. Zhang, T.; Liu, Z.; Zhu, Y.A.; Liu, Z.; Sui, Z.; Zhu, K.; Zhou, X. Dry Reforming of Methane on Ni-Fe-MgO Catalysts: Influence of Fe on Carbon-Resistant Property and Kinetics. Appl. Catal. B Environ. 2020, 264, 118497. [Google Scholar] [CrossRef]
  27. Theofanidis, S.A.; Galvita, V.V.; Poelman, H.; Marin, G.B. Enhanced Carbon-Resistant Dry Reforming Fe-Ni Catalyst: Role of Fe. ACS Catal. 2015, 5, 3028–3039. [Google Scholar] [CrossRef]
  28. Papavasiliou, J.; Rawski, M.; Vakros, J.; Avgouropoulos, G. A Novel Post-Synthesis Modification of CuO-CeO2 Catalysts: Effect on Their Activity for Selective CO Oxidation. ChemCatChem 2018, 10, 2096–2106. [Google Scholar] [CrossRef]
  29. Yang, E.H.; Noh, Y.S.; 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]
  30. Thommes, M.; Kaneko, K.; Neimark, A.V.; Olivier, J.P.; Rodriguez-Reinoso, F.; Rouquerol, J.; Sing, K.S.W. Physisorption of Gases, with Special Reference to the Evaluation of Surface Area and Pore Size Distribution (IUPAC Technical Report). Pure Appl. Chem. 2015, 87, 1051–1069. [Google Scholar] [CrossRef]
  31. Al-Fatesh, A.S.; Kumar, R.; Kasim, S.O.; Aidid, A.; Anis, I.; Fakeeha, H.; Abasaeed, A.E.; Atia, H.; Armbruster, U.; Kreyenschulte, C.; et al. Effect of Cerium Promoters on an MCM-41-Supported Nickel Catalyst in Dry Reforming of Methane. Cite This Ind. Eng. Chem. Res 2021, 2022, 174. [Google Scholar] [CrossRef]
  32. Mironova-Ulmane, N.; Kuzmin, A.; Sildos, I.; Puust, L.; Grabis, J. Magnon and Phonon Excitations in Nanosized NiO. Latv. J. Phys. Tech. Sci. 2019, 56, 61–72. [Google Scholar] [CrossRef]
  33. Shebanova, O.N.; Peter Lazor, P. Raman study of magnetite (Fe3O4): Laser-induced thermal effects and oxidation. J. Raman Spectrosc. 2003, 34, 845–852. [Google Scholar] [CrossRef]
  34. Tang, C.-W.; Wang, C.-B.; Chien, S.-H. Characterization of cobalt oxides studied by FT-IR, Raman, TPR and TG-MS. Thermochim. Acta 2008, 473, 68–73. [Google Scholar] [CrossRef]
  35. Cazzanelli, E.; Kuzmin, A.; Mariotto, G.; Mironova-Ulmane, N. Study of Vibrational and Magnetic Excitations in NicMg1-CO Solid Solutions by Raman Spectroscopy. J. Phys. Condens. Matter 2003, 15, 2045–2052. [Google Scholar] [CrossRef]
  36. Funkenbusch, E.F.; Cornilsen, B.C. Two-Magnon Raman Scattering in Calcium Doped NiO. Solid State Commun. 1981, 40, 707–710. [Google Scholar] [CrossRef]
  37. González-Castaño, M.; Dorneanu, B.; Arellano-García, H. The Reverse Water Gas Shift Reaction: A Process Systems Engineering Perspective. React. Chem. Eng. 2021, 6, 954–976. [Google Scholar] [CrossRef]
  38. Stroud, T.; Smith, T.J.; Le Saché, E.; Santos, J.L.; Centeno, M.A.; Arellano-Garcia, H.; Odriozola, J.A.; Reina, T.R. Chemical CO2 Recycling via Dry and Bi Reforming of Methane Using Ni-Sn/Al2O3 and Ni-Sn/CeO2-Al2O3 Catalysts. Appl. Catal. B Environ. 2018, 224, 125–135. [Google Scholar] [CrossRef]
  39. Zou, Z.; Zhang, T.; Lv, L.; Tang, W.; Zhang, G.; Gupta, R.K.; Wang, Y.; Tang, S. Preparation adjacent Ni-Co bimetallic nano catalyst for dry reforming of methane. Fuel 2023, 343, 128013. [Google Scholar] [CrossRef]
  40. Palanichamy, K.; Sasirekha, N. Dry reforming of methane using bimetallic Pt–Ni/TiSBA-15: Effect of Si/Ti ratio on catalytic activity and stability. Int. J. Hydrogen Energy 2023, 48, 31126–31141. [Google Scholar] [CrossRef]
  41. Mohandessi, M.; Tavakolian, M.; Rahimpour, M.R. Cadmium as a robust and novel promoter over Ni@γ-Al2O3 catalysts in dry methane reforming. ACS Appl. Energy Mater. 2023, 6, 8209–8220. [Google Scholar] [CrossRef]
  42. Yao, L.; Galvez, M.E.; Hu, C.; Da Costa, P. Mo-Promoted Ni/Al2O3 Catalyst for Dry Reforming of Methane. Int. J. Hydrogen Energy 2017, 42, 23500–23507. [Google Scholar] [CrossRef]
Figure 1. N2 physical adsorption–desorption isotherm (A) and the pore width distribution via BJH (B).
Figure 1. N2 physical adsorption–desorption isotherm (A) and the pore width distribution via BJH (B).
Catalysts 13 01420 g001
Figure 2. H2-Temperature Programmed Reduction spectra of bimetallic catalysts.
Figure 2. H2-Temperature Programmed Reduction spectra of bimetallic catalysts.
Catalysts 13 01420 g002
Figure 3. X-ray diffraction.
Figure 3. X-ray diffraction.
Catalysts 13 01420 g003
Figure 4. CO2-Temperature Programmed Desorption profile of Bimetallic catalysts.
Figure 4. CO2-Temperature Programmed Desorption profile of Bimetallic catalysts.
Catalysts 13 01420 g004
Figure 5. Raman spectroscopy of fresh bimetallic catalysts.
Figure 5. Raman spectroscopy of fresh bimetallic catalysts.
Catalysts 13 01420 g005
Figure 6. SEM images of 2.5Ni+2.5Ce-MG63 (A,B), 2.5Ni+2.5Co-MG63 (C,D), 2.5Ni+2.5Sr-MG63 (E,F), where A, C, E (fresh) and B, D, F (used) catalysts.
Figure 6. SEM images of 2.5Ni+2.5Ce-MG63 (A,B), 2.5Ni+2.5Co-MG63 (C,D), 2.5Ni+2.5Sr-MG63 (E,F), where A, C, E (fresh) and B, D, F (used) catalysts.
Catalysts 13 01420 g006
Figure 7. TEM image of 2.5Ni+2.5Co-MG63 (A) Fresh, (B) Used.
Figure 7. TEM image of 2.5Ni+2.5Co-MG63 (A) Fresh, (B) Used.
Catalysts 13 01420 g007
Figure 8. Conversions of (a) CH4 and (b) CO2 as a result of TOS (reaction circumstances: CH4/CO2/N2 = 3/3/1 (v/v/v); GHSV = 42,000 mL/gcat/h; Mcat = 0.1g; t = 700 °C).
Figure 8. Conversions of (a) CH4 and (b) CO2 as a result of TOS (reaction circumstances: CH4/CO2/N2 = 3/3/1 (v/v/v); GHSV = 42,000 mL/gcat/h; Mcat = 0.1g; t = 700 °C).
Catalysts 13 01420 g008
Figure 9. TGA curve for temperature gravimetric analysis of the used catalyst after 7 h time on-stream.
Figure 9. TGA curve for temperature gravimetric analysis of the used catalyst after 7 h time on-stream.
Catalysts 13 01420 g009
Figure 10. Conversions of CH4 and CO2 as a result of long-time on-stream (reaction circumstances: CH4/CO2/N2 = 3/3/1 (v/v/v); GHSV = 42,000 mL/gcat/h; Mcat = 0.1 g; t = 700 °C).
Figure 10. Conversions of CH4 and CO2 as a result of long-time on-stream (reaction circumstances: CH4/CO2/N2 = 3/3/1 (v/v/v); GHSV = 42,000 mL/gcat/h; Mcat = 0.1 g; t = 700 °C).
Catalysts 13 01420 g010
Figure 11. TPO profile of 2Ni+2.5Ce-MG63 and 2Ni+2.5Ce-MG63 catalysts after 30 h of reaction.
Figure 11. TPO profile of 2Ni+2.5Ce-MG63 and 2Ni+2.5Ce-MG63 catalysts after 30 h of reaction.
Catalysts 13 01420 g011
Table 1. The textural characteristics of the catalysts.
Table 1. The textural characteristics of the catalysts.
SampleBET (m2/g)P.V. (cm3/g)Pore Size (Å)
2.5Ni+2.5Co-MG631990.30958.7
2.5Ni+2.5Ce-MG632010.31256.7
2.5Ni+2.5Fe-MG632000.32959.8
2.5Ni+2.5Sr-MG631740.28358.2
Table 2. The amount of H2 used for the TPR analysis.
Table 2. The amount of H2 used for the TPR analysis.
SampleMaximum
Temperature (°C)
Quantity
(cm3/g.STP)
Total Quantity
(cm3/g.STP)
2.5Ni+2.5Co-MG63343.4
923.6
0.36
19.18
19.55
2.5Ni+2.5Ce-MG63684.2
944.7
0.34
12.13
12.47
2.5Ni+2.5Fe-MG63543.2
924.3
17.46
14.22
31.68
2.5Ni+2.5Sr-MG63636.4
705.8
957.9
0.10
1.24
13.12
14.57
Table 3. The quantity of CO2 desorbed in the TPD.
Table 3. The quantity of CO2 desorbed in the TPD.
SampleTemperature (°C)Quantity
(cm3/g.STP)
Total Quantity (cm3/g.STP)
2.5Ni+2.5Co-MG63251.2
696.5
10.04
0.39
10.43
2.5Ni+2.5Ce-MG63252.8
729.1
9.66
0.36
10.02
2.5Ni+2.5Fe-MG63273. 628.3828.38
2.5Ni+2.5Sr-MG63271.0
831.6
28.32
0.50
28.82
Table 4. Performance assessment of the current study in comparison to past studies.
Table 4. Performance assessment of the current study in comparison to past studies.
Cat-Name
(°C)
CH4/CO2GHSV
(mL/hgcat)
RT (°C)CH4-Conv. (%)CO2-Conv. (%) REF.
Ni-Mn/CeO2-ZrO21:120,0006003940[34]
Na-Ni/ZrO268:3112,0006752555[37]
Ni-Cu/Mg(Al)O1:160,0006005057[38]
Ni-Co/SiO21:170024,0003040[39]
Ni-Pt/Ti-SBA-151:170036,0005760[40]
Ni-Cd1:170012,0005060[41]
Ni-Mo/Al2O31:120,0009006070[42]
2.5Ni+2.5Co-MG63 25:2536,0007007276This work
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

Ibrahim, A.A.; Fakeeha, A.H.; Abasaeed, A.E.; Wazeer, I.; Bentalib, A.; Siva Kumar, N.; Abu-Dahrieh, J.K.; Al-Fatesh, A.S. The Effectiveness of Ni-Based Bimetallic Catalysts Supported by MgO-Modified Alumina in Dry Methane Reforming. Catalysts 2023, 13, 1420. https://doi.org/10.3390/catal13111420

AMA Style

Ibrahim AA, Fakeeha AH, Abasaeed AE, Wazeer I, Bentalib A, Siva Kumar N, Abu-Dahrieh JK, Al-Fatesh AS. The Effectiveness of Ni-Based Bimetallic Catalysts Supported by MgO-Modified Alumina in Dry Methane Reforming. Catalysts. 2023; 13(11):1420. https://doi.org/10.3390/catal13111420

Chicago/Turabian Style

Ibrahim, Ahmed A., Anis H. Fakeeha, Ahmed E. Abasaeed, Irfan Wazeer, Abdulaziz Bentalib, Nadavala Siva Kumar, Jehad K. Abu-Dahrieh, and Ahmed S. Al-Fatesh. 2023. "The Effectiveness of Ni-Based Bimetallic Catalysts Supported by MgO-Modified Alumina in Dry Methane Reforming" Catalysts 13, no. 11: 1420. https://doi.org/10.3390/catal13111420

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