Next Article in Journal
Ideal Site Geometry for Heterogeneous Catalytic Reactions: A DFT Study
Previous Article in Journal
Assessment of Reaction Kinetics for the Dehydrogenation of Perhydro-Dibenzyltoluene Using Mg- and Zn-Modified Pt/Al2O3 Catalysts
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Tailored Ni-MgO Catalysts: Unveiling Temperature-Driven Synergy in CH4-CO2 Reforming

by
Ahmad M. Alghamdi
1,
Ahmed A. Ibrahim
2,
Fekri Abdulraqeb Ahmed Ali
1,
Nouf A. Bamatraf
3,
Anis H. Fakeeha
2,
Ahmed I. Osman
4,*,
Salwa B. Alreshaidan
3,
Farid Fadhillah
1,
Salma A. Al-Zahrani
5 and
Ahmed S. Al-Fatesh
2,*
1
Chemical Engineering Department, Imam Mohammad Ibn Saud Islamic University (IMSIU), Riyadh 11432, Saudi Arabia
2
Chemical Engineering Department, College of Engineering, King Saud University, Riyadh 11421, Saudi Arabia
3
Department of Chemistry, Faculty of Science, King Saud University, P.O. Box 800, Riyadh 11451, Saudi Arabia
4
School of Chemistry and Chemical Engineering, Queen’s University Belfast, Belfast BT9 5AG, UK
5
Chemistry Department, Faculty of Science, University of Ha’il, P.O. Box 2440, Ha’il 81451, Saudi Arabia
*
Authors to whom correspondence should be addressed.
Catalysts 2024, 14(1), 33; https://doi.org/10.3390/catal14010033
Submission received: 9 December 2023 / Revised: 22 December 2023 / Accepted: 27 December 2023 / Published: 30 December 2023
(This article belongs to the Section Catalytic Materials)

Abstract

:
This study examines nickel catalysts on two different supports—magnesium oxide (MgO) and modified MgO (with 10 wt.% MOx; M = Ti, Zr, Al)—for their effectiveness in the dry reforming of methane. The reactions were conducted at 700 °C in a tubular microreactor. The study compares the best-performing catalyst with a reference catalyst (5Ni/MgO) by conducting dry reforming of methane at different reaction temperatures. The catalysts are evaluated using surface area, porosity, X-ray diffraction, infrared spectroscopy, transmission electron microscope, thermogravimeter, and temperature-programmed techniques. The 5Ni/MgO + ZrO2 catalyst demonstrates inferior catalytic activity due to insufficient active sites. On the other hand, the 5Ni/MgO + TiO2 catalyst shows limited catalytic excellence due to excessive coke deposits, which are six times higher than other catalysts. The 5Ni/MgO and 5Ni/MgO + Al2O3 catalysts have the richest basic and acidic profiles, respectively. The 5Ni/MgO + Al2O3 catalyst is superior to other catalysts due to its stronger metal–support interaction on the expanded surface and the efficient diffusion of carbon on its less crystalline surface. At 700 °C, this catalyst achieves 73% CH4 conversion, and at 800 °C, it reaches 83% conversion. This study emphasizes the crucial role of the reaction temperature in reducing carbon deposition and enhancing the efficiency of the reforming process.

Graphical Abstract

1. Introduction

Technologies for carbon capture and utilization are crucial for diminishing CO2 emissions and preventing global warming [1]. The CO2 reformation of methane (DRM), also referred to as “dry”, is an attractive process for producing syngas (CO/H2) that directly uses greenhouse gas emissions of CO2 and CH4 [2]. The produced syngas has multiple uses, for example as feedstock, the Fischer-Tropsch process, fuel synthesis, chemical synthesis, carbonylation, and hydroformylation [2]. DRM has several advantages, including yielding a syngas ratio of unity that allows for selective modification of feed concentrations for additional chemical synthesis [3]. However, it is important to note that DRM is a strong endothermic process, which means that the reaction requires high temperatures. These high temperatures can cause sintering and agglomeration, which can deactivate the catalysts [4,5]. Despite this, numerous studies have found that Ni-based catalysts are effective and promising DRM candidates due to their high availability, low cost, and high activity when compared to noble metals like Pd, Rh, and Ru [6]. Nevertheless, although Ni particles show the greatest activity in the DRM response, they are hampered by sintering Ni and carbon buildup [3,4,5]. There must be a robust metal support relationship to halt carbon formation and Ni agglomeration, particularly over the DRM [6]. For the DRM method to be viable on a wide scale, it is essential to improve catalytic performance with suitable support materials [7,8,9,10]. Numerous attempts have been made [11,12] to reduce the amount of carbon that deposits on Ni catalysts. Reducing the size of the nickel particles is beneficial. It is established that the combination of NiO and support helps shrink the size 1of the nickel particles even at high temperatures [13]. To achieve a strong metal–support contact, it is essential to choose appropriate metal–oxide combinations and synthesis techniques. Since Mg and Ni have analogous atomic sizes, MgO can easily form a solid solution with NiO. Other advantages include good thermal stability and an inexpensive cost. The Ni-Mg interaction is beneficial to improve the nickel’s dispersion and fine-tune its particle size. Many researchers have established that catalysts supported by MgO show outstanding performance [14]. MgO support can be further enhanced by modifying it with other oxides such as Al2O3, TiO2, and ZrO2. Al2O3 modifier support is characterized by a large surface area and excellent stability, providing highly dispersed Ni active sites with a confinement effect [15]. A NiAl2O4 catalyst was developed and tested for DRM reactions at 800 °C by Zhang et al. [16]. The catalyst demonstrated a high level of long-term stability and resistance to coking and sintering. Similarly, Roger et al. [17] also elaborated on nickel aluminate’s stability and activity by examining catalysts (NiAl4O7, NiAl2O4, and Ni2Al2O5) with varying concentrations. Due to their regenerating characteristics, NiAl2O4 and Ni2Al2O5 reductions had high and consistent activity under challenging DRM conditions. The benefits of the ZrO2 support modifier include high acidic/basic, oxygen mobility, oxidizing/reducing, and heat stability [18,19]. Androulakis et al. found that the methane conversion turnover frequency (TOF) of TiO2-supported Rh catalysts gave optimum performance [20]. TiO2-supported catalysts have proven to be good for improving the performance of Rh-based catalysts in the partial oxidation of methane because they help maximize and retain rhodium dispersion by avoiding its re-oxidation [21]. TiO2-supported Rh was among the various supports investigated by Yokota et al. [22]. It exhibited optimum activity contrary to Al2O3-supported Rh, which presented the lowest turnover frequency (TOF). The former catalyst was able to retain Rh in a reduced state. According to the authors, electronic interactions between Rh and support were responsible for maintaining Rh’s most active DRM reaction metallic state. Equations (1)–(4) represent the DRM and the associated side reactions. The main equation for dry reforming is Equation (1), and the reverse water shift process is represented by Equation (2). The creation of coke is caused by Equation (3) of methane decomposition and Equation (4) of the Boudouard reaction.
C H 4 + C O 2 2 C O + 2 H 2 Δ H 298 K = + 247   k J / m o l e
H 2 + C O 2 C O + H 2 O Δ H 298 K = + 41   k J / m o l e
C H 4 C + 2 H 2 Δ H 298 K = + 75   k J / m o l e
2 C O C + C O 2 Δ H 298 K = 127   k J / m o l e
In this study, a Ni catalyst supported on stabilized MgO with Al2O3, ZrO2, and TiO2 was prepared and applied as the catalyst for DRM at 700 °C and 800 °C. The catalytic performance, carbon resistance, and stability of the prepared catalysts with different modified magnesium oxide supports were investigated systematically. To achieve high conversion performance at reaction temperatures of 700 °C and 800 °C and to prevent metal agglomeration and coke formation, it is crucial to develop active as well as sintering and coke-resistant catalysts. This will open up new possibilities for the DRM reaction’s limited commercialization [23]. Various characterization techniques, such as X-ray diffraction, N2 physisorption, temperature-programmed reduction Raman spectroscopy, TGA, etc., were used for both fresh and used samples.

2. Characterization and Activity Results

2.1. Overview of Textural Analysis

N2 adsorption–desorption isotherms and the BET study of catalysts are shown in Figure S1 and Table 1. According to the IUPAC classification, the catalysts showed type-IV isotherms with an H1-type hysteresis loop, which describes a mesoporous structure. Upon incorporating 20 wt.% titania (or zirconia) along with 80 wt.% MgO, the surface area of the catalyst sample (5Ni/MgO + TiO2) is not affected much. However, the incorporation of 20 wt.% alumina along with 80 wt.% MgO results in a 50% rise in the surface area of the catalyst sample (5Ni/MgO + Al2O3). This aspect shows that combining oxides influences the catalysts’ structural characteristics.

2.2. X-ray Diffraction (XRD) Analysis

Figure 1 shows the XRD patterns of NiO, MgO, TiO2, ZrO2, MgNiO2, NiTiO3, and MgTiO3. The diffraction peaks of the cubic phase of MgO are located at 2θ = 36.92°, 42.82°, 62.09°, 74.48°, and 78.43° (reference number 01-075-0447), whereas the cubic NiO phase characteristic is found at 2θ = 37.00° and 43.02° (reference number 00-047-1049). The anatase phase TiO2 diffraction peaks are located at 2θ = 27.61° and 36.25° (reference number 01-073-1764), whereas the rutile phase TiO2 diffraction peaks are located at 2θ = 28.51° and 54.52° (reference number 01-078-1510). The location of ZrO2’s monoclinic phase is at 2θ = 24.04°, 24.53°, 28.24°, 31.54°, 34.19°, 35.39°, 50.16°, 55.44°, and 59.96° (reference number 00-007-0343), while the rhombohedral phase of NiTiO3 is located at 2θ = 24.20°, 33.09°, 35.62°, 40.83°, 49.36°, 53.82°, and 63.88° (reference number 00-033-0960), and the rhombohedral phase of MgTiO3 is located at 2θ = 21.24°, 24.20°, 33.09°, 35.62°, 40.83°, 49.36°, 53.82°, and 63.88° (reference number 00-002-0901). The diffraction peaks for the cubic phase of MgNiO2 are situated at 2θ = 36.92°, 42.82°, 62.09°, and 74.48° (reference numbers 00-024-0712, 00-003-0999). Overall, the 5Ni/MgO and 5Ni/MgO + Al2O3 catalysts are populated with cubic MgO phases as well as cubic MgNiO2 phases. Over 5Ni/MgO + TiO2, the intensity of Mg-related phases is decreased, and Ti-related phases (rhombohedral NiTiO3 phase, rutile TiO2 phase, and anatase TiO2 phase) appear. In the same way, the 5Ni/MgO + ZrO2 catalyst is populated with the monoclinic zirconia phase.

2.3. Temperature-Programmed Reduction and Desorption Study

The temperature-programmed reduction (H2-TPR) profile of catalysts is shown in Figure S2. H2-TPR was used to understand the interaction between the active catalyst and its support. The catalysts showed broad negative peaks between 200 and 300 °C, possibly caused by hydrogen seeping into the mesopores [24]. The catalysts exhibit modest interactions between the NiO and the support matrix, with a peak temperature of around 500 °C. The strong interaction between NiO and the supports is responsible for the absorption peaks observed between 600 and 800 °C. The formation of Mg2NiO3, MgNiO2, and MgNi2O3 as a result of a solid solution between NiO and MgO could be the cause of the high-temperature peaks for all catalysts [25]. In Table 1, the catalyst with 5Ni/MgO + ZrO2 displayed a minimum hydrogen consumption of 2.22 cm3/g, whereas the catalyst with 5Ni/MgO + TiO2 displayed a maximum hydrogen consumption of 13.32 cm3/g. Additional reduction features over 5Ni/MgO + TiO2 may be due to the reduction of TiO2 and rhombohedral NiTiO3 phases [26]. Among 5Ni/MgO and 5Ni/MgO + Al2O3, 5Ni/MgO + Al2O3 has a higher amount of reducible NiO which is under strong metal support interaction. The basicity profile of catalysts is studied by CO2 temperature-programmed desorption (CO2-TPD) (Figure 2A). The CO2-TPD profile of catalysts shows desorption peaks at about 260 °C and 330 °C, which are attributed to weak basic sites and medium strength basic sites, respectively [25]. Interestingly, 5Ni/MgO has the highest quantity of desorbed CO2, indicating the optimum number of basic sites over the catalyst surface. Over the MgO + ZrO2-supported Ni catalyst, the intensity of medium strength basic sites is depleted faster than weak basic sites, whereas over the MgO + TiO2-supported Ni catalyst, the population of both types of basic sites decreased proportionally. The basic profile of 5Ni/MgO + Al2O3 is the worst and it had minimum CO2 desorption. Overall, the catalyst surfaces of 5Ni/MgO, “5Ni/MgO + ZrO2 and 5Ni/MgO + TiO2”, and 5Ni/MgO + Al2O3 have high, moderate, and low concentrations of basic sites. The acidic profile of the catalysts was studied by NH3 temperature-programmed desorption (NH3-TPD) (Figure 2B). It is found that 5Ni/MgO and 5Ni/MgO + ZrO2 have a low concentration of acid sites, 5Ni/MgO + TiO2 retains a moderate concentration of acid sites and 5Ni/MgO + Al2O3 attains a high concentration of acid sites. The minimum basicity of the MgO + Al2O3-supported Ni catalyst can be justified by neutralization of basic sites over MgO by acidic alumina oxide. Again, the highest concentration of acid sites over 5Ni/MgO + Al2O3 is solely due to the presence of acidic aluminum oxide in the support.

2.4. Infrared (IR) Spectroscopy of Fresh and CO2-Treated Catalyst

The infrared spectra of fresh and the CO2-treated catalysts are shown in Figure 3. Infrared spectra of the CO2-treated catalysts (for 2 h) show CO2-surface-interacting species grown over the catalyst surface. Over 5Ni/MgO, the bidentate carbonate species (about 1087 cm−1), carboxylate species (about 1417 cm−1), ionic carbonate species (1460 cm−1), and bidentate formate species (about 2850–2925 cm−1) are cultivated upon treatment with CO2 [27]. Over 5Ni/MgO + TiO2, IR vibration peaks for bidentate carbonate are not observed, whereas over 5Ni/MgO + ZrO2 and 5Ni/MgO/Al2O3, only carboxylate and ionic carbonates are grown.

2.5. Raman Analysis of Spent Catalysts

The Raman spectra of the spent catalysts are shown in Figure 4. The method displays the type and degree of graphitization on the catalysts that are being employed. The temperature range between 1250 and 1800 cm−1 displays D and G band characteristic peaks. The D band located at 1341 cm−1 is represents imperfect, haphazard carbon deposits (amorphous), while the G band located at 1574 cm−1 represents graphite. The degree of carbon crystallinity that was produced during the reaction can be accurately determined thanks to the relative intensity sizes of the D and G bands (ID/IG). Smaller ratios suggest that the graphitized carbon is primarily responsible for the crystallinity. The ID/IG values for our spent catalysts were 0.72, 1.13, 1.70, and 1.72 for the 5Ni/MgO + ZrO2, 5Ni/MgO, 5Ni/MgO + TiO2, and 5Ni/MgO + Al2O3. 5Ni/MgO + Al2O3 had the highest (ID/IG) ratios, indicating low crystallinity due to a deficiency in graphitized carbon. Consequently, the degree of carbon graphitization over the 5Ni/MgO + Al2O3 catalyst was lower than that of the other catalysts. The catalyst did not deactivate early due to amorphous or poorly crystalline carbon deposits; however, the catalytic activity declined by the crystalline graphitic carbon, as seen in the 5Ni/MgO + ZrO2 sample.

2.6. TGA Analysis of Spent Catalysts

In Table 1, the weight loss of spent catalysts is shown. 5Ni/MgO, 5Ni/MgO + Al2O3, and 5Ni/MgO + ZrO2 catalysts had weight losses ranging between 11 and 13%, while the 5Ni/MgO + TiO2 sample had a high weight loss of 64%. These TGA profiles might be connected to the catalysts’ reactivity. Indeed, the 5Ni/MgO + TiO2 catalyst exhibited lower performance than the 5Ni/MgO and 5Ni/MgO + Al2O3 catalysts, thus generating higher carbon deposition.

2.7. Transmission Electron Microscopy

Figure S3 displays the 5Ni/MgO + Al2O3 catalyst’s fresh and spent TEM images and the fresh and used TEM images of 5Ni/MgO. The fresh catalysts exhibit small Ni particles of semispherical and cylindrical shapes, with varying sizes. The particles are uniformly distributed on the support, and their boundaries are extremely clear. However, the used catalysts show agglomerated particles that lead to the formation of interparticle voids. The formation of nanotubes of different diameters and lengths is found, particularly for the non-modified supported MgO.

2.8. Catalytic Activity Results

Before the developed catalysts were examined, the catalytic performance was investigated in a blank experiment carried out in a stainless-steel reactor with the same input ratio and reaction temperatures. The blank’s CH4 and CO2 conversions were 1.0% and 0.5%, respectively. The profiles of the CO2 and CH4 conversions as well as the H2/CO ratio are displayed in Figure 5. The performance of the catalyst was assessed at 700 °C. Compared to the equivalent CH4 conversions, the CO2 conversions are greater. The existence of the reverse water gas shift reaction in Equation (2) may be the cause of this phenomenon. The catalyst 5Ni/MgO + Al2O3 had the best dry reforming performance, with an H2/CO ratio of 0.94 and CH4 and CO2 conversions of 73.8% and 79.1%, respectively.
The reduction of the active metal is proportional to hydrogen consumption. The turnover frequency (TOF), which indicates how many times the catalytic cycle occurs on a single site per unit time, is computed as shown in Table 1. The 5Ni/MgO + Al2O3 catalyst has the highest value of 48.4 h−1; the 5Ni/MgO + ZrO2 catalyst has the lowest value of 34.88 h−1. The low TOF over 5Ni/MgO + ZrO2 may be related to its low hydrogen intake, which results in the least amount of reduced active Ni. Table S1 presents a comparative analysis of the current catalyst and those found in previous studies that were used for methane dry reforming. The outcome demonstrates the catalyst’s applicability and the need for more development.

2.9. Catalytic Activity Results at Different Reaction Temperatures

The effect of reaction temperature was further examined using the pristine supported catalyst 5Ni/MgO and the best-obtained modified catalyst 5Ni/MgO + Al2O3. Both 600 °C and 800 °C reaction temperatures were tested. Figure 6 exhibits the conversions of both methane and carbon dioxide versus time on the stream. The performance increased 2.5 times when the reaction temperature was increased from 600 °C to 800 °C. The increase could be attributed to the endothermic nature of the process. The best performance was attained using an 800 °C reaction temperature. The methane and carbon dioxide conversions acquired were 83% and 87%, respectively.

2.10. Temperature-Programmed Oxidation (TPO) Study over the Spent Catalyst

Temperature-programmed oxidation (TPO) analysis was utilized to examine the type of carbon collected on the surface of the DRM catalysts being used. In the DRM process, the TPO experiments were performed for seven hours of reaction time using spent catalysts, as shown in Figure 7 and Figure S4. Structurally organized carbon with a reduced oxidation reactivity is responsible for the peaks corresponding to the evolution of CO2 at high temperatures. According to the TPO characterization analysis, inert carbon was not present when the catalysts were operated at 800 °C and the low-temperature peak was attributed to the existence of amorphous/active carbon. However, at 600 °C or 700 °C, the catalysts’ peaks appeared due to structurally ordered carbon with lower oxidation reactivity. The extra 2% rise in carbon deposition seen on the modified supported catalyst could be attributed to its higher reactivity.

2.11. Impact of Reaction Temperature on the TGA Analysis

The impact of reaction temperature on the TGA analysis of 5Ni/MgO and 5Ni/MgO + Al2O3 catalysts is shown in Figure 8. The methane dry reforming reaction appears to reach its maximum carbon deposition at 600 °C, moderate carbon deposition at 700 °C, and least carbon deposition at 800 °C. This behavior was noted in an earlier study [28]. The decrease of carbon deposition upon increasing reaction temperature can be justified thermodynamically. At low reaction temperatures, CH4 decomposition reactions, hydrogen-consuming reactions, and other coke deposition reaction pathways (like CO + H2 → H2O + C and CO2 + 2H2 → 2H2O + C) are thermodynamically feasible [29]. At about 700 °C, even coke removal reactions like C + H2O → CO + H2 become feasible, and the above coke deposition pathways have lost their thermodynamic feasibility. At 800 °C, spontaneous thermal cracking of CH4 and their consequent oxidation results in the least coke deposit.

2.12. Raman Analysis of Spent Catalysts

Figure 9 shows the Raman spectra of 5Ni/MgO and 5Ni/MgO + Al2O3 catalysts at different reaction temperatures. The relative intensity indices (ID/IG) increase with the decrease in reaction temperature, favoring the formation of low-crystallinity carbon.

3. Discussion

Based on the characterization results, it can be said that combining oxides with support modifies the catalyst texture, crystallinity, reducibility, and acid–basic property of 5Ni/MgO + MOx (M = Zr, Ti, and Al) catalysts. Further, these surface properties govern catalytic activity and coke deposition accordingly. Under the CO2 stream, 5Ni/MgO + ZrO2 was found to be populated with carboxylate and ionic carbonates (as per the IR results). The feeble active sites (2.22 cm3/g) and unstable monoclinic ZrO2 phase over 5Ni/MgO + ZrO2 make the catalyst inferior (50% CH4 conversion) in the dry reforming of methane reactions. 5Ni/MgO + TiO2 has a comparable surface area (to 5Ni/MgO) and the highest number of reducible species over the catalyst surface. However, all reducible species are not active sites. Some reduced species may be formed by the reduction of TiO2 phases, which are not active sites for DRM. Under the CO2 stream, the 5Ni/MgO + TiO2 catalyst has one additional surface-interacted CO2 species (bidentate formate species). Overall, surface area increases, increasing the number of catalytic active sites and additional CO2-surface interacting species (than 5Ni/MgO + ZrO2), which make the catalyst more active (64% CH4 conversion) for DRM. However, the excessive coke deposition, which is over six times more than other catalysts, and the presence of a higher amount of graphitic carbon, limit the catalyst’s performance further. 5Ni/MgO attains the highest amount of basicity by which various types of CO2-interacting species, such as bidentate carbonate, carboxylate species, ionic carbonate, and bidentate formate species, are populated over the catalyst surface (Figure 10A). It has the highest crystallinity where Ni sites are derived from cubic MgNiO2 phases. The 5Ni/MgO achieves as high as 70% CH4 conversion during 430 min on stream.
It is interesting to note that the 5Ni/MgO + Al2O3 catalyst has relatively fewer catalytic active species and fewer CO2-interacting surface species than 5Ni/MgO, but it attains the highest catalyst activity (Figure 10B). In DRM, CH4 decomposition (into CH4−x) is the rate-determining step. After that, the oxidation of dissociated-CH4 (or CH4−x) by “surface interacting CO2 species” or “gases CO2” does not affect the reaction rate. Therefore, over the 5Ni/MgO + Al2O3 catalyst, oxidation of dissociated CH4 species by gaseous CO2 under the Eley–Rideal model [30,31] can be expected (Figure 10B, route (b)). In terms of population of catalytic active sites over 5Ni/MgO + Al2O3, one possible answer can be drawn that the 5Ni/MgO + Al2O3 catalyst has a higher number of Ni sites under strong interaction (than 5Ni/MgO), which are derived at 700–1000 °C under H2-stream (Figure S3). This means that the 5Ni/MgO + Al2O3 catalyst has more stable active sites in the DRM reaction temperature than 5Ni/MgO. Again, in the acid–base profile, the 5Ni/MgO catalyst has the highest number of basic sites, whereas the 5Ni/MgO + Al2O3 has the highest number of acid sites. Acid sites are responsible for coke formation, which may affect the catalytic activity if coke covers the catalytic active sites. The coke deposition over the 5Ni/MgO + Al2O3 catalyst is slightly higher than the 5Ni/MgO. Again, the O2-TPO study shows the presence of more inert carbon over 5Ni/MgO + Al2O3 than 5Ni/MgO at about 700 °C. However, Raman analysis shows the presence of the least degree of carbon graphitization over the 5Ni/MgO + Al2O3 catalyst. The inert and less graphitic carbon over 5Ni/MgO + Al2O3 is not found to affect the activity, even though the activity of the 5Ni/MgO + Al2O3 catalyst is higher than 5Ni/MgO. It is noticeable that the 5Ni/MgO + Al2O3 catalyst has the highest surface area (50% more than 5Ni/MgO) and less crystallinity, whereas the 5Ni/MgO catalyst has the highest crystalline surface. That means that the 5Ni/MgO + Al2O3 catalyst has more space and fewer crystalline constraints for diffusion of carbon deposit than 5Ni/MgO. Therefore, over 5Ni/MgO + Al2O3, the rate of CH4 deposition is counter-supported by the rate of carbon diffusion (away from the active sites) over a “less crystalline and expanded” surface [32] (Figure 10B, route c). However, over a highly crystalline 5Ni/MgO catalyst, the rate of carbon formation may not properly match the rate of carbon diffusion (Figure 10A, route c). So, even after high coke deposition over 5Ni/MgO + Al2O3, carbon deposit is transferred away from the active sites properly and leaves the active sites exposed to the reaction. This means it can be said that stable catalytic active sites and better diffusion of carbon deposit away from the active site over a more expanded surface make 5Ni/MgO + Al2O3 a better catalyst than 5Ni/MgO. It conveys about 73% CH4 conversion at 700 °C and 83% CH4 conversion at 800 °C in about 430 min on stream.

4. Experiments

4.1. Materials

Nickel nitrate hexahydrate (Ni (NO3)2.6H2O, 98%, Alfa Aesar, Haverhill, MA, USA), titania (TiO2-P25, 99.9%, from Degussa p25, Nanoshel, Cheshire, UK), alumina (Al2O3, 97.7% Norton Chemical Process Products Corp, Northboro, MA, USA), magnesia (MgO, 99.5%. BDH, Dubai, United Arab Emirates), and zirconia (ZrO2, 99.8%, Anhui-Elite, Hefei, China) were obtained commercially and were used as received.

4.2. Catalyst Preparation

The impregnation procedure was adopted to prepare the catalysts. A pristine support of MgO and 80% MgO combined with 20%m (m = Al2O3, ZrO2, and TiO2) mixed oxide supports were prepared by mixing 30 mL of distilled water and the appropriate amount of Ni (NO3)2·6H2O to get a 5.0 wt.% loading of nickel oxide, and the needed amount of support was mixed and stirred at room temperature in a crucible. For 30 min, the solution was dried while being stirred at 80 °C. The catalysts were calcined at 600 °C for three hours. The catalyst characterization is described in Supplementary Information S1.

4.3. Catalytic Activity Test

The catalytic activity was performed in a fixed-bed, continuous-flow reactor (PID Eng. & Tech, Madrid, Spain). At atmospheric pressure, the dry reforming reaction was carried out over a sample of 0.1 g of catalyst in a 0.94 cm internal diameter and 30 cm long stainless-steel tube. The reaction temperature was monitored using a thermocouple placed in the center of the catalyst bed. Before the reaction, hydrogen (H2) was introduced at a flow rate of 30 mL/min, at 800 °C, to convert nickel oxide to nickel metal, the active form of the catalyst. Afterward, the reactor was purged under nitrogen (N2) flow (20 mL/min) to remove any H2 left in the reactor. Then, the reactor temperature was adjusted under the flow of N2 until the reaction temperatures of 600, 700 °C and 800 °C were reached. A 70 mL/min feed flow rate comprised 30, 30, and 10 mL/min of CH4, CO2, and N2, respectively. The space velocity was set at 42,000 mL (h·gcat)−1. The reactor outlet gases, containing the product of syngas and unconverted feed gases, were analyzed using online gas chromatography (Shimadzu 2014, Kyoto, Japan), equipped with a TCD. The entire separation of reaction products was achieved by using two columns in series/bypass connections: Porapak Q and Molecular Sieve 5A. The following equations were used in this work to compute the H2/CO mole ratio and methane and carbon dioxide conversions:
C H 4   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 O 2   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
H 2 C O = m o l e s   o f   H 2   p r o d u c e d m o l e s   o f   C O   p r o d u c e d

5. Conclusions

The dry reforming of methane (DRM) generated syngas utilizing Ni catalysts supported on MgO and stabilizing MgO with TiO2, ZrO2, and Al2O3. Various tests were conducted on Ni catalysts supported on different oxides, exploring the impact of reaction temperature. The 5Ni/MgO + ZrO2 catalyst has a shortage of sufficient active sites, resulting in its inferior catalytic performance (~50% CH4 conversion at 430 min TOS). The 5Ni/MgO + TiO2 catalyst is enriched with active sites and CO2-interacting surface species, but excessive coke deposition (six times more than other catalysts) limits the CH4 conversion below 65% at 430 min TOS. The 5Ni/MgO catalyst has the richest basic profile, and this catalyst is populated by reducible NiO-interacted species (having modest and strong interaction with support) as well as a wide population of CO2-interacting surface species. The catalyst attains 70% CH4 conversion. The 5Ni/MgO + Al2O3 catalyst has the richest acidic profile and worst basic profile, but it exhibits the highest catalytic activity for DRM, achieving CH4 and CO2 conversions of 73% and 79% over seven hours at 700 °C. The superiority of 5Ni/MgO + Al2O3 over 5Ni/MgO is due to stronger metal–support interaction, oxidation of carbon deposit by mostly gaseous CO2, and timed diffusion of carbon over a “less crystalline and expanded” surface. Regarding reaction temperatures, the least quantity of carbon was deposited at 800 °C, with CH4 and CO2 conversion rates reaching 83% and 87% over 5Ni/MgO + Al2O3, respectively.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/catal14010033/s1, Paragraph S1: Catalyst Characterization; Figure S1: N2 physisorption for 5 Ni/MgO and 5 Ni\MgO+m catalysts (where m= Al2O3); Figure S2: H2-TPR curves of the fresh samples; Figure S3: TEM images of 5Ni\MgO+Al2O3. (A) Fresh; (B) spent, 5Ni\MgO; (C) fresh; (D) spent; Figure S4: TPO images of spent-5Ni\MgO (A) and spent-5Ni\MgO+Al2O3 (B) catalysts showing the effect of reaction temperatures; Figure S5: DTG plots of 5Ni\MgO (A) and 5Ni\MgO+Al2O3 (B) catalysts; Table S1: Comparison of this catalyst’s efficiency with previous ones; References S1: [33,34,35,36,37,38,39,40,41].

Author Contributions

A.M.A.: methodology, writing, formal analysis, data curation; A.A.I., A.S.A.-F.: methodology, conceptualization, investigation, data curation, writing—original draft. F.A.A.A.: review, data curation, validation, software, formal analysis. S.A.A.-Z.: data curation, visualization, software, writing—review, funding acquisition, data curation, conceptualization, investigation, methodology. N.A.B., S.B.A.: funding acquisition, project administration, writing—review and editing, investigation, methodology. A.H.F., F.F.: data curation, investigation, writing—review and editing. A.I.O.: data curation, investigation, writing—review and editing. All authors have read and agreed to the published version of the manuscript.

Funding

The authors extend their appreciation to the Deputyship for Research & Innovation, Ministry of Education in Saudi Arabia for funding this research through the project number IFP-IMSIU-2023098. The authors also appreciate the Deanship of Scientific Research at Imam Mohammad Ibn Saud Islamic University (IMSIU) for supporting and supervising this project.

Data Availability Statement

No new data were created or analyzed in this study. Data sharing is not applicable to this article.

Conflicts of Interest

The authors state that none of the work disclosed in this study could have been influenced by any known conflicting financial interests or personal relationships.

References

  1. Chatla, A.; Almanassra, I.W.; Kallem, P.; Atieh, M.A.; Alawadhi, H.; Akula, V.; Banat, F. Dry (CO2) Reforming of Methane over Zirconium Promoted Ni-MgO Mixed Oxide Catalyst: Effect of Zr Addition. J. CO2 Util. 2022, 62, 102082. [Google Scholar] [CrossRef]
  2. Akri, M.; Zhao, S.; Li, X.; Zang, K.; Lee, A.F.; Isaacs, M.A.; Xi, W.; Gangarajula, Y.; Luo, J.; Ren, Y.; et al. Atomically Dispersed Nickel as Coke-Resistant Active Sites for Methane Dry Reforming. Nat. Commun. 2019, 10, 5181. [Google Scholar] [CrossRef] [PubMed]
  3. Bachiller-Baeza, B.; Mateos-Pedrero, C.; Soria, M.A.; Guerrero-Ruiz, A.; Rodemerck, U.; Rodríguez-Ramos, I. Transient Studies of Low-Temperature Dry Reforming of Methane over Ni-CaO/ZrO2-La2O3. Appl. Catal. B Environ. 2013, 129, 450–459. [Google Scholar] [CrossRef]
  4. Jamsaz, A.; Pham-Ngoc, N.; Wang, M.; Jeong, D.H.; Oh, E.S.; Shin, E.W. Synergistic Effect of Macroporosity and Crystallinity on Catalyst Deactivation Behavior over Macroporous Ni/CexZr1-XO2–Al2O3 for Dry Reforming of Methane. Chem. Eng. J. 2023, 476, 146821. [Google Scholar] [CrossRef]
  5. de la Cruz-Flores, V.G.; Martinez-Hernandez, A.; Gracia-Pinilla, M.A. Deactivation of Ni-SiO2 Catalysts That Are Synthetized via a Modified Direct Synthesis Method during the Dry Reforming of Methane. Appl. Catal. A Gen. 2020, 594, 117455. [Google Scholar] [CrossRef]
  6. 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]
  7. Zhang, M.; Zhang, J.; Wu, Y.; Pan, J.; Zhang, Q.; Tan, Y.; Han, Y. Insight into the Effects of the Oxygen Species over Ni/ZrO2 Catalyst Surface on Methane Reforming with Carbon Dioxide. Appl. Catal. B Environ. 2019, 244, 427–437. [Google Scholar] [CrossRef]
  8. Liu, D.; Quek, X.Y.; Cheo, W.N.E.; Lau, R.; Borgna, A.; Yang, Y. MCM-41 Supported Nickel-Based Bimetallic Catalysts with Superior Stability during Carbon Dioxide Reforming of Methane: Effect of Strong Metal-Support Interaction. J. Catal. 2009, 266, 380–390. [Google Scholar] [CrossRef]
  9. Yuan, J.; Lu, C.; Gu, Z.; Cai, J.; Zhao, H.; Li, D.; Jiang, L.; Xu, H.; Li, Z.q.; Li, K. Ni-Co Catalyst-Assisted Carbon Cycling for CH4-CO2 Reforming. Appl. Catal. B Environ. 2024, 341, 123318. [Google Scholar] [CrossRef]
  10. Liang, D.; Wang, Y.; Chen, M.; Xie, X.; Li, C.; Wang, J.; Yuan, L. Dry Reforming of Methane for Syngas Production over Attapulgite-Derived MFI Zeolite Encapsulated Bimetallic Ni-Co Catalysts. Appl. Catal. B Environ. 2023, 322, 122088. [Google Scholar] [CrossRef]
  11. Gangadharan, P.; Kanchi, K.C.; Lou, H.H. Evaluation of the Economic and Environmental Impact of Combining Dry Reforming with Steam Reforming of Methane. Chem. Eng. Res. Des. 2012, 90, 1956–1968. [Google Scholar] [CrossRef]
  12. Zhao, B.; Yan, B.; Yao, S.; Xie, Z.; Wu, Q.; Ran, R.; Weng, D.; Zhang, C.; Chen, J.G. LaFe0.9Ni0.1O3 Perovskite Catalyst with Enhanced Activity and Coke-Resistance for Dry Reforming of Ethane. J. Catal. 2018, 358, 168–178. [Google Scholar] [CrossRef]
  13. Lenggoro, I.W.; Itoh, Y.; Iida, N.; Okuyama, K. Control of Size and Morphology in NiO Particles Prepared by a Low-Pressure Spray Pyrolysis. Mater. Res. Bull. 2003, 38, 1819–1827. [Google Scholar] [CrossRef]
  14. Jafarbegloo, M.; Tarlani, A.; Mesbah, A.W.; Muzart, J.; Sahebdelfar, S. NiO-MgO Solid Solution Prepared by Sol-Gel Method as Precursor for Ni/MgO Methane Dry Reforming Catalyst: Effect of Calcination Temperature on Catalytic Performance. Catal. Lett. 2016, 146, 238–248. [Google Scholar] [CrossRef]
  15. Gao, X.; Ge, Z.; Zhu, G.; Wang, Z.; Ashok, J.; Kawi, S. Reforming of Methane: Recent Progress and Prospects. Catalysts 2021, 11, 1003. [Google Scholar] [CrossRef]
  16. Zhang, S.; Ying, M.; Yu, J.; Zhan, W.; Wang, L.; Guo, Y.; Guo, Y. NixAl1O2-δ Mesoporous Catalysts for Dry Reforming of Methane: The Special Role of NiAl2O4 Spinel Phase and Its Reaction Mechanism. Appl. Catal. B Environ. 2021, 291, 120074. [Google Scholar] [CrossRef]
  17. Rogers, J.L.; Mangarella, M.C.; D’Amico, A.D.; Gallagher, J.R.; Dutzer, M.R.; Stavitski, E.; Miller, J.T.; Sievers, C. Differences in the Nature of Active Sites for Methane Dry Reforming and Methane Steam Reforming over Nickel Aluminate Catalysts. ACS Catal. 2016, 6, 5873–5886. [Google Scholar] [CrossRef]
  18. Oni, B.A.; Tomomewo, O.S.; Sanni, S.E.; Ojo, V.O. Dry Reforming of Methane with CO2 over Co-La1−xCaxNiO3 Perovskite-Type Oxides Supported on ZrO2. Mater. Today Commun. 2023, 36, 106802. [Google Scholar] [CrossRef]
  19. Zhang, M.; Zhou, X.; Yang, J.; Yang, T.; Liu, Z.; Han, Y. Deciphering the ZrO2 Phase Engineering Effects on Dry Reforming of Methane over the Ni/ZrO2 Catalysts. Fuel 2023, 349, 128705. [Google Scholar] [CrossRef]
  20. Androulakis, A.; Yentekakis, I.V.; Panagiotopoulou, P. Dry Reforming of Methane over Supported Rh and Ru Catalysts: Effect of the Support (Al2O3, TiO2, ZrO2, YSZ) on the Activity and Reaction Pathway. Int. J. Hydrogen Energy 2023, 48, 33886–33902. [Google Scholar] [CrossRef]
  21. Mateos-Pedrero, C.; González-Carrazán, S.R.; Soria, M.A.; Ruíz, P. Effect of the Nature of TiO2 Support over the Performances of Rh/TiO2 Catalysts in the Partial Oxidation of Methane. Catal. Today 2013, 203, 158–162. [Google Scholar] [CrossRef]
  22. Yokota, S.; Okumura, K.; Niwa, M. Support Effect of Metal Oxide on Rh Catalysts in the CH4-CO2 Reforming Reaction. Catal. Lett. 2002, 84, 131–134. [Google Scholar] [CrossRef]
  23. Mao, Y.; Zhang, L.; Zheng, X.; Liu, W.; Cao, Z.; Peng, H. Coke-Resistance over Rh–Ni Bimetallic Catalyst for Low Temperature Dry Reforming of Methane. Int. J. Hydrogen Energy 2023, 48, 13890–13901. [Google Scholar] [CrossRef]
  24. Shen, H.; Li, H.; Yang, Z.; Li, C. Magic of Hydrogen Spillover: Understanding and Application. Green Energy Environ. 2022, 7, 1161–1198. [Google Scholar] [CrossRef]
  25. Al-Fatesh, A.S.; Kumar, R.; Kasim, S.O.; Ibrahim, A.A.; Fakeeha, A.H.; Abasaeed, A.E.; Alrasheed, R.; Bagabas, A.; Chaudhary, M.L.; Frusteri, F.; et al. The Effect of Modifier Identity on the Performance of Ni-Based Catalyst Supported on γ-Al2O3 in Dry Reforming of Methane. Catal. Today 2020, 348, 236–242. [Google Scholar] [CrossRef]
  26. de Bokx, P.K.; Bonne, R.L.C.; Geus, J.W. Strong Metal-Support Interaction in Ni/TiO2 Catalysts: The Origin of TiOx Moieties on the Surface of Nickel Particles. Appl. Catal. 1987, 30, 33–46. [Google Scholar] [CrossRef]
  27. Khatri, J.; Fakeeha, A.H.; Kasim, S.O.; Lanre, M.S.; Abasaeed, A.E.; Ibrahim, A.A.; Kumar, R.; Al-Fatesh, A.S. Ceria Promoted Phosphate-Zirconia Supported Ni Catalyst for Hydrogen Rich Syngas Production through Dry Reforming of Methane. Int. J. Energy Res. 2021, 45, 19289–19302. [Google Scholar] [CrossRef]
  28. Al-Fatish, A.S.A.; Ibrahim, A.A.; Fakeeha, A.H.; Soliman, M.A.; Siddiqui, M.R.H.; Abasaeed, A.E. Coke Formation during CO2 Reforming of CH4 over Alumina-Supported Nickel Catalysts. Appl. Catal. A Gen. 2009, 364, 150–155. [Google Scholar] [CrossRef]
  29. Al-Fatesh, A.S.; Arafat, Y.; Kasim, S.O.; Ibrahim, A.A.; Abasaeed, A.E.; Fakeeha, A.H. In Situ Auto-Gasification of Coke Deposits over a Novel Ni-Ce/W-Zr Catalyst by Sequential Generation of Oxygen Vacancies for Remarkably Stable Syngas Production via CO2-Reforming of Methane. Appl. Catal. B Environ. 2021, 280, 119445. [Google Scholar] [CrossRef]
  30. Mark, M.F.; Mark, F.; Maier, W.F. Reaction Kinetics of the CO2 Reforming of Methane. Chem. Eng. Technol. 1997, 20, 361–370. [Google Scholar] [CrossRef]
  31. Kathiraser, Y.; Oemar, U.; Saw, E.T.; Li, Z.; Kawi, S. Kinetic and Mechanistic Aspects for CO2 Reforming of Methane over Ni Based Catalysts. Chem. Eng. J. 2015, 278, 62–78. [Google Scholar] [CrossRef]
  32. Bayahia, H.; Fakeeha, A.H.; Al-Zahrani, S.A.; Alreshaidan, S.B.; Al-Awadi, A.S.; Alotibi, M.F.; Kumar, R.; Al-Fatesh, A.S. COx -Free H2 Production via Catalytic Decomposition of CH4 over Fe Supported on Tungsten Oxide-Activated Carbon Catalyst: Effect of Tungsten Loading. Arab. J. Chem. 2023, 16, 104781. [Google Scholar] [CrossRef]
  33. Shah, M.; Bordoloi, A.; Nayak, A.K.; Mondal, P. Effect of Ti/Al ratio on the performance of Ni/TiO2-Al2O3 catalyst for methane reforming with CO2. Fuel Process. Technol. 2019, 192, 21–35. [Google Scholar] [CrossRef]
  34. Bian, Z.; Zhong, W.; Yu, Y.; Wang, Z.; Jiang, B.; Kawi, S. Dry reforming of methane on Ni/mesoporous-Al2O3 catalysts: Effect of calcination temperature. Int. J. Hydrogen Energy 2021, 46, 31041–31053. [Google Scholar] [CrossRef]
  35. 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]
  36. Jeon, K.-W.; Kim, H.-M.; Kim, B.-J.; Lee, Y.-L.; Na, H.-S.; Shim, J.-O.; Jang, W.-J.; Roh, H.-S. Synthesis gas production from carbon dioxide reforming of methane over Ni-MgO catalyst: Combined effects of titration rate during co-precipitation and CeO2. Fuel Process. Technol. 2021, 219, 106877. [Google Scholar] [CrossRef]
  37. Khan, M.M.; Jin, L.; Khan, M.M.; Li, Y.; Saulat, H.; Zhang, Y.; Sarfraz, M.; Zhu, J.; Hu, H. CO2 reforming of methane over activated carbon-Ni/MgO-Al2O3 composite catalysts for syngas production. Fuel Process. Technol. 2021, 211, 106595. [Google Scholar] [CrossRef]
  38. Alabi, W.O.; Adesanmi, B.M.; Wang, H.; Patzig, C. Correlation of MgO loading to spinel inversion, octahedral site occupancy, site generation and performance of bimetal CoeNi catalyst for dry reforming of CH4. Int. J. Hydrogen Energy 2024, 51, 1087–1098. [Google Scholar] [CrossRef]
  39. Gutiérrez, M.C.; Hernández, P.S.; Anzures, F.M.; Martínez, A.G.; Galicia, G.M.; García, M.E.F.; Escobar-Alarcón, L.; Morales, F.J.T.; Pérez-Hernández, R. MgO impregnation to Al2O3 supported Ni catalyst for SYNGAS production using greenhouse gases: Some aspects of chemical state of Ni species. Int. J. Hydrogen Energy 2024, 52, 1131–1140. [Google Scholar] [CrossRef]
  40. binti Rosdin, R.D.; Yusuf, M.; Abdullah, B. Dry reforming of methane over Ni-based catalysts: Effect of ZrO2 and MgO addition as support. Mater. Lett. 2021, 12, 100095. [Google Scholar] [CrossRef]
  41. Kim, B.-J.; Park, H.-R.; Lee, Y.-L.; Ahn, S.-Y.; Kim, K.-J.; Hong, G.-R.; Roh, H.-S. Customized Ni-MgO-ZrO2 catalysts for the dry reforming of methane using coke oven gas: Optimizing the MgO content. J. CO2 Util. 2023, 68, 102379. [Google Scholar] [CrossRef]
Figure 1. The XRD patterns of catalysts.
Figure 1. The XRD patterns of catalysts.
Catalysts 14 00033 g001
Figure 2. Temperature-programmed desorption study of catalysts. (A) CO2-TPD profile; (B) NH3-TPD.
Figure 2. Temperature-programmed desorption study of catalysts. (A) CO2-TPD profile; (B) NH3-TPD.
Catalysts 14 00033 g002
Figure 3. FTIR spectra for different supports (A) MgO, (B) MgO + TiO2, (C) MgO + ZrO2, and (D) MgO + Al2O3 of fresh and (CO2 Treated) catalysts.
Figure 3. FTIR spectra for different supports (A) MgO, (B) MgO + TiO2, (C) MgO + ZrO2, and (D) MgO + Al2O3 of fresh and (CO2 Treated) catalysts.
Catalysts 14 00033 g003
Figure 4. Raman spectra of the spent samples operated at 700 °C.
Figure 4. Raman spectra of the spent samples operated at 700 °C.
Catalysts 14 00033 g004
Figure 5. Conversions of CO2, CH4, CO2, and H2/CO mole ratio as a function of TOS (reaction conditions: CH4/CO2/N2 = 3/3/1 (v/v/v); GHSV = 42,000 mL/gcat/h; Mcat = 0.1 g; t = 700 °C.
Figure 5. Conversions of CO2, CH4, CO2, and H2/CO mole ratio as a function of TOS (reaction conditions: CH4/CO2/N2 = 3/3/1 (v/v/v); GHSV = 42,000 mL/gcat/h; Mcat = 0.1 g; t = 700 °C.
Catalysts 14 00033 g005
Figure 6. Conversions at various reaction temperatures of (A) CH4 and (B) CO2 as a function of TOS (reaction conditions: CH4/CO2/N2 = 3/3/1 (v/v/v); GHSV = 42,000 mL/gcat/h; Mcat = 0.1 g; t = 600–800 °C).
Figure 6. Conversions at various reaction temperatures of (A) CH4 and (B) CO2 as a function of TOS (reaction conditions: CH4/CO2/N2 = 3/3/1 (v/v/v); GHSV = 42,000 mL/gcat/h; Mcat = 0.1 g; t = 600–800 °C).
Catalysts 14 00033 g006
Figure 7. TPO analysis for 5Ni/MgO and 5Ni/MgO + Al2O3 catalysts at different temperatures (A) 600 °C; (B)700 °C; (C) 800 °C.
Figure 7. TPO analysis for 5Ni/MgO and 5Ni/MgO + Al2O3 catalysts at different temperatures (A) 600 °C; (B)700 °C; (C) 800 °C.
Catalysts 14 00033 g007
Figure 8. TGA analysis for (A) 5Ni/MgO and (B) 5Ni/MgO + Al2O3 catalysts at different reaction temperatures.
Figure 8. TGA analysis for (A) 5Ni/MgO and (B) 5Ni/MgO + Al2O3 catalysts at different reaction temperatures.
Catalysts 14 00033 g008
Figure 9. Raman analysis for (a) 5Ni/MgO and (b) 5Ni/MgO + Al2O3 catalysts at different Reaction temperature.
Figure 9. Raman analysis for (a) 5Ni/MgO and (b) 5Ni/MgO + Al2O3 catalysts at different Reaction temperature.
Catalysts 14 00033 g009
Figure 10. The proposed reaction scheme over (A) 5Ni/MgO and (B) 5Ni/MgO + Al2O3. (a) Decomposition of CH4 over active sites. (b) Oxidation of CH4−x species by surface CO2-interacting species or CO2 (gas). (c) Polymerization of CH4−x species.
Figure 10. The proposed reaction scheme over (A) 5Ni/MgO and (B) 5Ni/MgO + Al2O3. (a) Decomposition of CH4 over active sites. (b) Oxidation of CH4−x species by surface CO2-interacting species or CO2 (gas). (c) Polymerization of CH4−x species.
Catalysts 14 00033 g010
Table 1. Summarizes the results of adsorption–desorption characterization, CH4 conversion ( X CH 4 ), deactivation factor (DF), turnover frequency (TOF), frequency, and weight loss (WL).
Table 1. Summarizes the results of adsorption–desorption characterization, CH4 conversion ( X CH 4 ), deactivation factor (DF), turnover frequency (TOF), frequency, and weight loss (WL).
SamplesSA
m2/g
P v
cm3/g
P D
Nm
Total
H2 Quantity
cm3/g
Total
CO2 Quantity
cm3/g
Total
NH3 Quantity
cm3/g
X C H 4
%
D.F aTOF b
h−1
WL %
IF
5Ni/MgO550.5239.811.114.498.4071.370.01.8246.3011.0
5Ni/MgO + Al2O3850.5323.69.722.9012.4073.873.10.9548.4313.0
5Ni/MgO + ZrO2530.3933.02.222.958.6057.550.811.6534.8812.5
5Ni/MgO + TiO2580.4836.313.323.838.1267.963.95.8944.5863.7
SA = Surface area, P v = Pore volume, P D = Pore diameter, I = initial, F = Final, D.F a = (Initial CH4 conversion-Final CH4 conversion) × 100/(Initial CH4 conversion), WL = weight loss.; TOF b = turnover frequency.
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

Alghamdi, A.M.; Ibrahim, A.A.; Ali, F.A.A.; Bamatraf, N.A.; Fakeeha, A.H.; Osman, A.I.; Alreshaidan, S.B.; Fadhillah, F.; Al-Zahrani, S.A.; Al-Fatesh, A.S. Tailored Ni-MgO Catalysts: Unveiling Temperature-Driven Synergy in CH4-CO2 Reforming. Catalysts 2024, 14, 33. https://doi.org/10.3390/catal14010033

AMA Style

Alghamdi AM, Ibrahim AA, Ali FAA, Bamatraf NA, Fakeeha AH, Osman AI, Alreshaidan SB, Fadhillah F, Al-Zahrani SA, Al-Fatesh AS. Tailored Ni-MgO Catalysts: Unveiling Temperature-Driven Synergy in CH4-CO2 Reforming. Catalysts. 2024; 14(1):33. https://doi.org/10.3390/catal14010033

Chicago/Turabian Style

Alghamdi, Ahmad M., Ahmed A. Ibrahim, Fekri Abdulraqeb Ahmed Ali, Nouf A. Bamatraf, Anis H. Fakeeha, Ahmed I. Osman, Salwa B. Alreshaidan, Farid Fadhillah, Salma A. Al-Zahrani, and Ahmed S. Al-Fatesh. 2024. "Tailored Ni-MgO Catalysts: Unveiling Temperature-Driven Synergy in CH4-CO2 Reforming" Catalysts 14, no. 1: 33. https://doi.org/10.3390/catal14010033

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