Next Article in Journal
Adsorption Sites on Pd Nanoparticles Unraveled by Machine-Learning Potential with Adaptive Sampling
Next Article in Special Issue
Effect of Cactus (Opuntia ficus-indica) and Acacia (Acacia seyal) Gums on the Pasting, Thermal, Textural, and Rheological Properties of Corn, Sweet Potato, and Turkish Bean Starches
Previous Article in Journal
Effect of Spirulina (Arthrospira platensis) Supplementation on Physical and Chemical Properties of Semolina (Triticum durum) Based Fresh Pasta
Previous Article in Special Issue
Relationship between the Chemical Composition and the Biological Functions of Coffee
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

The Role of Strontium in CeNiO3 Nano-Crystalline Perovskites for Greenhouse Gas Mitigation to Produce Syngas

1
Department of Chemistry, College of Science, King Saud University, P.O. Box 2454, Riyadh 11451, Saudi Arabia
2
Department of Zoology, College of Science, King Saud University, P.O. Box 2454, Riyadh 11451, Saudi Arabia
3
Central Laboratory, Department of Botany and Microbiology, College of Science, King Saud University, P.O. Box 2454, Riyadh 11451, Saudi Arabia
4
Central Laboratory, College of Science, King Saud University, P.O. Box 2454, Riyadh 11451, Saudi Arabia
*
Author to whom correspondence should be addressed.
Molecules 2022, 27(2), 356; https://doi.org/10.3390/molecules27020356
Submission received: 29 November 2021 / Revised: 27 December 2021 / Accepted: 4 January 2022 / Published: 6 January 2022

Abstract

:
The transition metal-based catalysts for the elimination of greenhouse gases via methane reforming using carbon dioxide are directly or indirectly associated with their distinguishing characteristics such as well-dispersed metal nanoparticles, a higher number of reducible species, suitable metal–support interaction, and high specific surface area. This work presents the insight into catalytic performance as well as catalyst stability of CexSr1−xNiO3 (x = 0.6–1) nanocrystalline perovskites for the production of hydrogen via methane reforming using carbon dioxide. Strontium incorporation enhances specific surface area, the number of reducible species, and nickel dispersion. The catalytic performance results show that CeNiO3 demonstrated higher initial CH4 (54.3%) and CO2 (64.8%) conversions, which dropped down to 13.1 and 19.2% (CH4 conversions) and 26.3 and 32.5% (CO2 conversions) for Ce0.8Sr0.2NiO3 and Ce0.6Sr0.4NiO3, respectively. This drop in catalytic conversions post strontium addition is concomitant with strontium carbonate covering nickel active sites. Moreover, from the durability results, it is obvious that CeNiO3 exhibited deactivation, whereas no deactivation was observed for Ce0.8Sr0.2NiO3 and Ce0.6Sr0.4NiO3. Carbon deposition during the reaction is mainly responsible for catalyst deactivation, and this is further established by characterizing spent catalysts.

1. Introduction

Among the various well-established CO2 conversion processes, such as electrochemical catalysis, photocatalysis, and thermal catalysis, methane reforming using carbon dioxide, commonly known as dry reforming of methane (DRM), has recently attracted scientists primarily because DRM converts major greenhouse gases, i.e., carbon dioxide and methane, to synthesize hydrogen and carbon monoxide, also called synthesis gas, which is further utilized to produce liquid hydrocarbons [1,2,3,4,5,6,7,8]. Hence, the DRM process not only plays a role in greenhouse gas mitigation and thus serves as a cause of climate change, but also generates synthesis gas, a mixture of equimolar hydrogen and carbon monoxide suitable for the production of hydrocarbon via the well-known Fischer–Tropsch synthesis process [9,10,11,12,13]. It is well established that the proper choice and suitable design of a catalyst plays significant role in catalytic activity and stability during DRM [14,15,16]. Both noble metal-based and transition metal-based catalysts have been reported for DRM, but noble metal-based catalysts are expensive and less abundant despite their excellent catalytic activities [8]. Therefore, transition metal-based catalysts serve as an efficient alternative for DRM. Among transition metal-based catalysts, nickel (Ni)- and cobalt (Co)-based catalysts are most investigated for DRM due to the fact that these catalysts are cost-effective, more abundant, and offer quick turnover rates [17,18,19,20]. The major challenge related to Ni-based catalysts is their deactivation during DRM, associated with sintering (which leads to loss of active metal surface area) and carbon deposition [21,22,23].
In general, the metallic nanoparticles supported on basic oxide supports and/or promoters, such as Sr2+, La2O3, and CeO2, etc., demonstrate better catalytic activity and benefit CO2 chemisorption compared to acidic supports. Ceria and modified-ceria supports have been found to play a promising role in the endothermic DRM process due to the decisive factors associated with these supports, including their high oxygen storage capacity and/or oxygen vacancies, which not only activate CO2 but also help to gasify carbon deposited during DRM and their basicity, which further promotes CO2 adsorption [15,23,24,25]. These characteristic features, when combined with tiny (nano)-metallic particles, exhibit better catalytic efficiency, lower the reaction temperature, and promote the water–gas-shift reaction. For instance, Makri et al. [26] synthesized 5 wt% Ni/Ce1−xMxO2−δ (M = Zr4+, Pr3+, x = 0.2 and 0.5) catalysts by citrate route sol-gel and studied the influence of chemical composition on nickel particle size, which in turn influenced the different kind of carbon deposition under dry reforming reaction conditions.
In addition to the metallic nanoparticle-supported oxides, nickel-based double-layered hydroxides (LDHs) have shown excellent catalytic performances towards DRM reaction [27,28,29,30]. Modification of double-layered hydroxides with Y, Zr, and Ce as dopants or promotors significantly improves CH4 and CO2 conversions, lowers carbon deposition, and limits side reactions [28,29,30].
Although supported catalysts and LDHs have excellent catalytic performance in the DRM process, they eventually deactivate due to rapid growth of randomly sized impregnated nanoparticles, which assist the carbon deposition and blockage of active sites at higher temperature. To sort out the coalescence and coking deposition, well-defined crystal structures such as perovskite (ABO3) and pyrochlore (A2B2O7) are more suitable candidates for the DRM process because they produce uniformly distributed metallic nanoparticles under a reducing atmosphere (B/AOx).
Ni-based perovskites are thermally stable and offer high metal dispersion, which leads to their excellent performance in DRM [31,32]. The partial substitution at the A or B sites in ABO3 has a distinguished influence for the development of the perovskite phase, especially when substitutions are performed with the metal ions that have a different valence state—for example, substitution of the A trivalence site by bivalence or tetravalence ions (Sr2+, Ca2+, or Ce4+) would relatively improve the size and dispersion of exsolved active metal nanoparticles and increase the synergetic interaction between the active metals and the derived support [33,34,35,36,37,38]. These derived factors are useful for better catalytic activity, more carbon, and metal sintering resistance.
The investigation of the role of supported LaxNiOy/MgAl2O4 and bulk LaxNiOy catalysts for dry methane reforming showed that the supported LaxNiOy/MgAl2O4 catalysts exhibited improved catalytic activity and catalyst durability for over 65 h time-on-stream with less carbon deposition, which was attributed to their larger specific surface areas and higher nickel dispersions. It was also found by XRD analysis of spent LaxNiOy/MgAl2O4 catalysts that metallic nickel active sites were intact even after 65 h of reaction [39]. Ruocco et al. [40] studied the influence of ternary perovskites AZrRuO3 (A = Ca, Ba, Sr), their synthesis technique, and operating conditions such as reaction temperature and gas hourly space velocity for the dry methane-reforming reaction. They discovered that ternary perovskite with Sr (SrZrRuO3) demonstrated higher activity and durability than the rest of the perovskites [40]. Utilizing perovskite as a support and subsequent deposition of nickel and cobalt to obtain bimetallic catalysts (Ni-Co/La2O3-LaFeO3) to analyze their activity and stability during dry methane reforming was reported by Wang et al. [41]. The crystalline structure of the perovskite, after a suitable amount of Co was added to the bimetallic catalyst, resulted in enhanced catalytic activity and stability with less carbon formation during the reforming reaction.
This research article demonstrates the impact of strontium incorporation into cerium- and nickel-based nanocrystalline perovskites (CeNiO3) in dry and/or CO2 reforming of methane. The catalytic activity performances are specifically elucidated in relation to their characterizations before and after reaction. Overall, the study is focused on exploring the insights of cerium replacement with strontium and the influence of this replacement on catalytic conversions and durability under reforming conditions.

2. Results

2.1. Characterization of Fresh Catalysts

2.1.1. Thermal Decomposition of Precursors

Thermal decomposition curves, including thermogravimetric (TG) and their differentials (DTG) of the precursors as a measure of their calcination temperatures, are shown in Figure 1. It can be seen from Figure 1 that each precursor, a gel-like mixture of metal nitrates and glycine, decomposed differently and both CeNiO3 and Ce0.8Sr0.2NiO3 demonstrated two-step decomposition, whereas three-step decomposition was observed for Ce0.6Sr0.4NiO3. Overall, the decomposition temperatures of CeNiO3 were relatively higher than strontium-incorporated Ce0.8Sr0.2NiO3 and Ce0.6Sr0.4NiO3. The onset temperature for initial weight loss (first-step decomposition) of ~8%, in the case of CeNiO3, was found to be ~75 °C; weight loss then reached its maximum at ~150 °C, and finally finished at ~300 °C. This first-step decomposition was assigned to the removal of the adsorbed water and/or release of adsorbed gases. The second-step decomposition exhibited a weight loss of ~21% observed in the temperature range of 300–550 °C (with peak maximum at 420 °C), which could be ascribed to the decomposition and/or combustion of organic matter. Despite showing an almost similar decomposition trend, the DTG peak temperatures for first- and second-step decomposition for Ce0.8Sr0.2NiO3 dropped slightly down to 145 and 410 °C, respectively.
In contrast, Ce0.6Sr0.4NiO3 demonstrated three-step decomposition, and the onset temperature for initial weight loss was observed to be 130 °C, which was attributed to the release of physically adsorbed water and/or gases. Second- and third-step decompositions, occurring at 450 and 870 °C, elucidated the combustion of organic matter present in the precursor.

2.1.2. X-ray Diffraction (XRD)

Figure 2 presents the XRD patterns of the as-prepared nanocrystalline perovskites with and without the incorporation. MDI Jade® software was used to analyze XRD data. The data analysis shows that a significant number of peaks corresponding to CeNiO3 were recorded [42,43,44,45], whereas diffraction peaks detected at 2θ values of 37.3°, 43.3°, and 62.9° were related to cubic NiO. Moreover, remaining peaks were assigned to oxides and carbonates of strontium, as labeled in Figure 2 [46].

2.1.3. Textural Properties and Elemental Analysis

Figure 3 shows the N2 adsorption–desorption isotherms and pore-size distributions of air-calcined precursors (700 °C, 5 h), and their BET surface areas and pore parameters are summarized in Table 1. Figure 3 shows that all of the samples displayed a type II isotherm with a H3 hysteresis loop in the relative pressure (P/P0) range of 0.8–1.0, indicating that these samples possessed macro-pores. The specific surface areas observed (17.3–25.6 m2/g) are in accordance with the literature, except that these values were slightly higher than the reported specific surface area, i.e., below 10 m2/g for similar types of materials [47]. The surface area of Ce0.8Sr0.2NiO3 catalyst was higher (25.6 m2/g) than that of the other fabricated catalysts. However, no clear trend as a function of Sr substitution extent was observed. Although Sr doping did not much affect the surface area, it greatly influenced the pore size and pore volume; the highest values were found for x = 0.2 and the lowest for x = 0.0. The increase in pore volume post Sr addition indicates that Sr could play a role in influencing the catalytic activity, which is discussed in Section 2.2 and Section 3. Furthermore, the elemental analysis of as-synthesized catalysts revealed that CeNiO3 catalyst exhibited an Ni/Ce ratio of 1, and hence, both Ni and Ce had similar compositions as measured from ICP-OES (Table 1). The chemical compositions of Sr-incorporated catalysts show that Ce was replaced with Sr.

2.1.4. Morphological Study (TEM) of Fresh and Reduced Perovskites

Transmission electron microscopy (TEM) was utilized to analyze the morphology of the fresh, reduced, and used nanocrystalline perovskite catalysts, and microscopic images of all the catalysts before and after reaction are shown in Figure 4. ImageJ® software was used for image processing and data analysis. The particles were observed to be spherical in shape, and the average particle sizes were between 5–34, 12–27, and 14–25 nm for CeNiO3, Ce0.8Sr0.2NiO3, and Ce0.6Sr0.4NiO3 fresh catalysts, respectively. It is interesting to note that insignificant sintering was observed for reduced catalysts, and particle sizes were 8–45, 15–42, and 18–40 nm for CeNiO3-, Ce0.8Sr0.2NiO3-, and Ce0.6Sr0.4NiO3- reduced catalysts, respectively.

2.1.5. Temperature-Programmed Reduction (TPR)

TPR is mainly utilized to examine the reducibility, the metal–support interaction, and to find the activation and/or reduction temperature that results in metallic particles required for the catalytic reforming reaction. Figure 5 presents the reduction profiles, indicating the variations in metal–support interaction and/or the reducibility of CeNiO3 post Sr addition. The reduction profiles can be divided into two regions (region I < 525 and region II > 525 °C). The small reduction peak in region I below 220 °C in each catalyst is related to the reduction of weakly interacting nickel oxide species, whereas the reduction peaks appearing between 220 and 525 °C correspond to two-step reduction of Ni3+ to Ni2+ and Ni2+ to Ni0 [44,48]. The reduction profiles in region II are ascribed to either the reduction of ceria and/or the reduction of perovskite to metallic nickel at higher temperature [49]. CeNiO3- and Sr-incorporated catalysts reduce according to the reactions below:
CeNiO3 + H2 = Ni0 + CeO2 + H2O
CexSr1−xNiO3 + H2 = Ni0 + (1 − x)SrO1−x + xCeO2 + H2O
There were no significant change in the reduction peak temperatures (~290 °C) of CeNiO3 or Ce0.8Sr0.2NiO3; however, the peak intensity of Ce0.8Sr0.2NiO3 reduced to one third of CeNiO3, indicating the suppression of a number of reducible species post Sr addition. This trend completely changed when Sr was further added (Ce0.6Sr0.4NiO3), and more species reducible at higher temperatures were observed. It is reported that the addition of Sr to a supported Ni catalyst enhanced metal–support interaction and new peaks were found at higher temperature, which is in agreement with the TPR results of Ce0.6Sr0.4NiO3 but is contradictory to the Ce0.8Sr0.2NiO3 reduction profile, where a small shoulder was observed in region II. The enhanced reducibility and/or easier reduction of Ce0.8Sr0.2NiO3 can be attributed to a possible lowering of the valence state of Ni cations incorporated into the oxide. The role of these findings in influencing the catalytic activity is discussed in Section 2.2.

2.2. Catalytic Performances

As-synthesized nanocrystalline CexSr1−xNiO3 (x = 0.6–1) perovskites were studied for their performance in dry methane-reforming reaction at 700 °C. Prior to the dry-reforming reaction, all the catalysts were subjected to reduction/activation using hydrogen as a reducing gas to generate the required metallic nickel crystallites as active sites. Figure 6a,b shows methane (CH4) and carbon dioxide (CO2) conversions versus time-on-stream. It can be seen from Figure 5a that CeNiO3 demonstrated a higher CH4 conversion than Sr-incorporated catalysts. CeNiO3 exhibited an initial CH4 conversion of 54.3% and deactivated down to reach 50.1% after 7 h time-on-stream, leading to a deactivation factor of 7.7% (Table 1). The deactivation factor based on first-order kinetics also showed the same deactivation trends, as shown in Table 1 [50,51]. The impact of incorporating Sr into CeNiO3 over initial CH4 conversion was obvious, and decreased from 54.3% (CeNiO3) to 13.1% (Ce0.8Sr0.2NiO3) and 19.4% (Ce0.6Sr0.4NiO3) when x changed from 1 to 0.6. The loss in catalytic activity, in the case of Sr-incorporated catalysts, can be attributed to either nickel species agglomeration facilitated by Sr addition after the reduction or the covering of nickel active sites with Sr. Based on the TEM images of reduced catalysts, it is obvious that no significant sintering was observed after reduction; hence, the main reason behind the lower activity of Sr-incorporated catalysts might be nickel active sites covered with Sr. Interestingly, no deactivation was observed for strontium-incorporated perovskites (Ce0.8Sr0.2NiO3 and Ce0.6Sr0.4NiO3) and hence, deactivation factors were negative in these catalysts (Table 1). Moreover, conversions increased for the initial 90 min before becoming constant for strontium-incorporated perovskites, which is associated with the reduction of nickel species remaining oxidized even after activation under hydrogen atmosphere prior to reaction. The catalyst’s deactivation and associated factors behind deactivation are discussed in Section 3. Like Figure 6a,b shows a similar trend for CO2 conversions as a function of time. CO2 conversions of 64.8, 26.3, and 32.5% were exhibited by CeNiO3, Ce0.8Sr0.2NiO3, and Ce0.6Sr0.4NiO3, respectively, which eventually reached final CO2 conversions of 58, 36, and 33.6%, respectively. It is noteworthy from Figure 6 that CO2 conversions were higher than CH4 conversions. This finding leads to the fact that some side reactions were simultaneously taking place, such as the reverse water–gas-shift reaction (CO2 + H2 → CO + H2O) producing H2/CO molar ratios lower than unity, i.e., H2/CO = 1.0, since hydrogen was consuming CO2 (Figure 6c), as well as the Boudouard reaction and CO disproportionation reaction (2CO → CO2 + C). A dedicated section discusses the catalytic performance results in relation to their characterizations in Section 3.

2.3. Characterization of Spent Perovskites

The post dry reforming perovskites were further analyzed using different characterizations to explain the catalytic activity and/or stability results. Among the tested catalysts, only CeNiO3 deactivated over time, whereas Sr-incorporated catalysts showed increasing CO2 and CH4 conversions. Moreover, CO2 conversions higher than CH4 conversions (Figure 6) paved the way for side reactions, including reverse CO disproportionation and reverse water–gas-shift reaction. These post dry-reforming catalysts were characterized by temperature-programmed oxidation (TPO) and transition electron microscope (TEM) to explore the modifications experienced by perovskites during the dry-reforming reaction.

2.3.1. Temperature-Programmed Oxidation (TPO)

The possibility of carbon formation over the catalyst’s surface post dry-reforming reaction was demonstrated by TPO analysis, as shown in Figure 7. It is obvious that one broad peak was displayed in the temperature range of 110 to 500 °C for each catalyst. The TPO peak maxima were 285, 300, and 335 °C for Ce0.6Sr0.4NiO3, Ce0.8Sr0.2NiO3, and CeNiO3, respectively. These peak maxima suggest the formation of less-reactive surface carbides and/or polymeric carbon species during dry methane-reforming reaction, as reported earlier in the literature [52,53]. The hydrogenation degree of surface carbon species is related to TPO peak temperature, and these surface carbon species transform into graphitic carbon as TPO peak temperature moves towards higher temperatures. Furthermore, the peak temperature implies that the interaction between polymeric carbon species and catalyst surface varies and becomes weaker after Sr incorporation, as exhibited by the decrease in TPO peak maxima from 335 to 285 °C for CeNiO3 and Ce0.6Sr0.4NiO3. It is noteworthy that the carbon deposition over perovskite samples in this work was significantly lower than scandium- and magnesium-modified Ni- and Co-based bimetallic catalysts supported over an expensive SBA-15 support [3] and strontium-promoted Ni and Co-Al2O3 catalysts [4]. It can be concluded that cerium oxide and strontium incorporation influence the formation of carbon, carbon suppression, and carbon interaction with the catalyst surface.

2.3.2. Transition Electron Microscopy (TEM) and Elemental Analysis

In order to gain insight into the morphological changes and/or formation of deactivating coke over the surface of spent catalysts, TEM microscopic images were recorded post dry-reforming reaction. From the TEM images of spent catalysts in Figure 4g–i, particle sizes were found to be 10–70, 18–50, and 20–67 nm for CeNiO3, Ce0.8Sr0.2NiO3, and Ce0.6Sr0.4NiO3, respectively. The deposition of carbon over the catalysts’ surface and obvious catalyst particle agglomeration were also evident, which suggests the contribution of sintering towards the deactivation of CeNiO3. No deactivation, in the case of Sr-incorporated perovskites, was observed even after carbon deposition and particle agglomeration. These findings are further elaborated below in Section 3. Moreover, the elemental analysis of spent catalysts demonstrated that no variations in the chemical compositions of the catalyst were observed during the reaction.

3. Discussion

Prior to the dry methane-reforming reaction study, as-prepared nanocrystalline perovskites were characterized to envisage their activity performance during reforming reaction. Thermogravimetric analysis results, i.e., TG-DTG analysis data, illustrate the conversion of precursors into perovskites post calcination treatment at temperatures demonstrated by DTG (Figure 1 and Table S1). XRD diffraction profiles of as-synthesized catalysts (Figure 2) revealed the presence of nickel oxide, perovskite structure, and cerium dioxide along with carbonates and oxides of strontium. The formation of macroporous higher-specific-surface-area perovskite, e.g., 25.6 m2/g for Ce0.8Sr0.2NiO3, was obtained from nitrogen adsorption–desorption isotherms (Figure 3 and Table 1). TEM demonstrated (Figure 4) spherical-shaped particles with various sizes with and without strontium incorporation. The reduction profiles using TPR (Figure 5) exhibited easier reduction of oxides in CeNiO3 and Ce0.8Sr0.2NiO3, whereas stronger metal–support interaction and difficult-to-reduce oxides were formed in Ce0.6Sr0.4NiO3. Pre-reforming reaction characterization results indicate a higher number of reducible species, larger specific surface area, and a wide range of spherically shaped nanoparticles post strontium incorporation. Hence, it was expected that Sr-incorporated perovskite, such as Ce0.8Sr0.2NiO3, would potentially outperform the rest of the catalysts, but Figure 6 shows lower activity of this catalyst. The mechanism associated with dry methane reforming requires reactant adsorption over the active sites of the catalyst as an initial step. The reactants then dissociate and finally react with each other to form products that eventually leave the catalyst surface following desorption [4]. Nickel, in its metallic form, serves as the active site for the adsorption of CH4 [3]. Catalytic activity results reveal the loss in activity in terms of both CH4 and CO2 conversions post Sr incorporation into CeNiO3, which could be ascribed to the loss of active sites due to metal particle agglomeration, the presence of unreduced nickel oxides post activation, and/or active site coverage with carbonates or oxides of strontium [4,7]. TEM data show the extent of metal agglomeration or sintering prior to, during, and post reforming reaction, and it is evident that the degree of sintering remained the same and can be ruled out as a deactivation factor contributing to the lower activity of Sr-incorporated catalysts. Furthermore, the presence of strontium oxide and strontium carbonates, as evidenced by XRD, substantiate the hypothesis that coverage of nickel active sites by strontium species causes loss of catalytic activity in Sr-incorporated catalysts.
The activity results versus time-on-stream as a function of catalyst stability are shown in Figure 6. Despite higher activity, CeNiO3 perovskite deactivated over time, which is attributed to carbon formation as well as active metal sintering. These observations are substantiated by TEM and TPO characterization results of used catalysts. It is interesting to note that strontium-incorporated perovskites did not show any deactivation despite carbon deposition and sintering. The degree of agglomeration or sintering remained the same for all perovskites during the dry methane-reforming reaction. This suggests that methane decomposition [3], a prevalent DRM side reaction at higher reaction temperature, is the main source of carbon deposition causing catalyst deactivation, in accordance with TPO peak maxima—i.e., a higher temperature (335 °C) is required to gasify carbon formed over the surface of CeNiO3 than that of strontium-incorporated perovskites. Moreover, CO2 conversions in strontium-incorporated perovskites increased over time, which was not the case with CeNiO3. These findings infer that carbon gasification is promoted in an oxidative environment in strontium-incorporated perovskites, leading to no deactivation in these catalysts. The reaction mechanism and associated impact of Sr incorporation is further shown in Figure 8. The investigation of reduced La2-xSrxNiO4 perovskite oxides by Rynkowski et al. [7] inferred that Sr incorporation in smaller amounts exhibited less-active but stable catalytic performance in comparison with strontium-free catalysts. The study of the role of LaNiO3 perovskites after partially substituting Ni and La for dry methane-reforming reaction revealed that La replacement with Sr leads to a loss of catalytic activity [54]. Furthermore, the comparison of findings in this work with already published literature (Table 2) revealed that the current conversions outperformed similar perovskites such as CeNiO3 [55] or similar combinations with varying A and/or B substitutions in ABO3 [56,57,58,59]. Since the main focus of this work was to demonstrate the deactivation factors as a function of catalytic stability of perovskites, it is obvious that the perovskites in this work showed lower deactivation factors than [55,59] despite using higher space velocity.

4. Materials and Methods

Nanocrystalline perovskites, i.e., CexSr1−xNiO3 (x = 0.6–1), were prepared via the self-combustion method, and glycine and metallic nitrates were used as precursors. Initially, in 100 mL of deionized water, appropriate amount corresponding to x of hydrated cerium nitrate (Ce(NO3)3·nH2O, purchased from Sigma-Aldrich, St. Louis, MO, USA, with a purity of 99.9%), 1 mmole of nickel nitrate hexahydrate (Ni(NO3)2·6H2O purchased from Sigma-Aldrich with a purity of 99.9%), and appropriate amount corresponding to 1 − x of Sr(NO3)2 were dissolved. Then, glycine (purity of 99.5%) was added to metal nitrate solution as an ignition promoter by keeping the glyceine-to-metal-ions ratio close to unity. The mixtures were systematically stirred to obtain a homogenous gel-like solution by water elimination at 60–70 °C. Since the onset temperature for ignition reaction is ~250 °C, the gel was subjected to this temperature to produce powdered precursors that were eventually calcined under oxidative atmosphere at 700 °C for 6 h, leading to the formation of the perovskite structure. The nomenclature CexSr1−xNiO3, where x varies between 0.6 and 1, was used in all cases just as standardization, as it is usual in the literature.
The thermogravimetric analysis (TGA), including differential curves (DTG) of the perovskite precursors, was recorded using a TGA/SDTA851e thermal analyzer (purchased from Mettler-Toledo, Switzerland). The samples were subjected to heating from room temperature to 1000 °C using a heating rate of 20 °C/min under inert atmosphere (by flowing nitrogen at 100 mL/min). The following characterizations were employed over calcined catalysts. X-ray diffraction profiles were recorded using a scan rate of 0.2°/min between 10 and 80° on an XRD-6000 diffractometer (purchased from Shimadzu, Columbia, MD, USA) equipped with monochromatic radiation of CuK (λ = 1.5406 Å). Braunner, Emmet, and Teller (BET) specific surface areas of the samples were measured on a NOVA2000 BET system (purchased from Quantachrome, Boynton Beach, FL, USA). Temperature-programmed oxidation (TPO) and reduction (TPR) profiles were conducted on an Autochem 2920 apparatus (purchased from Micromeritics, Norcross, GA, USA). A total of 30 mg of each sample was placed in a U-shape quartz tube. The sample was first pretreated under synthetic air flowing at 50 mL/min at 300 °C for 1 h followed by cooling it down to room temperature. In order to record TPR patterns, samples were subjected to reducing atmosphere by flowing 10% H2/Ar at 25 mL/min while the sample temperature was raised from room temperature to 900 °C at 10 °C/min. The TPO samples were pretreated in a way similar to TPR. TPO profiles were recorded for the spent catalysts post dry-reforming stability tests to identify carbon formed over the catalyst surface. The samples were subjected to oxidative environment by flowing 10% O2/He at 25 mL/min and raising the sample temperature from room temperature to 800 °C at 10 °C/min. The morphology of fresh and used perovskite catalysts was analyzed by transmission electron microscopy (TEM) using a JEM-210 microscope (purchased from JEOL, Tokyo, Japan) equipped with an accelerating voltage of 80 kV. The chemical compositions in the as-synthesized catalysts were measured by using inductively coupled plasma optical emission spectroscopy (ICP-OES ELAN 6100 purchased from Perkin Elmer, Waltham, MA, USA).
The dry methane-reforming reaction was conducted in a tubular fixed-bed reactor (i.d. 1 cm, catalytic bed length 3 mm) at 700 °C and 1 atm pressure. The feed gas, comprising methane and carbon dioxide (1:1) balanced with nitrogen, was flown at 70 mL/min with a space velocity of 84,000 mL gcat.−1h−1. The catalysts were subjected to activation/reduction, prior to reaction, by flowing 10% H2 in nitrogen mixture (40 mL/min) at 700 °C for 2 h. After activation, the hydrogen mixture was replaced with helium to remove any leftover hydrogen and subsequently the reaction was carried out at 700 °C for over 7 h time-on-stream. The analysis of the reforming products and unconverted reactants was carried out using on-line gas chromatograph GC 3800 (purchased from VARIAN, Palo Alto, CA, USA) equipped with two columns (packed with Porapak N and 13X molecular sieves) and two thermal conductivity detectors. The results were reproducible with an error of ±5%. The conversions and molar ratios were calculated using following equations [60,61]:
CH 4   Converson   % = 100 × F CH 4 , in F CH 4 ,   out F CH 4 , in
CO 2   Converson   % = 100 × F CO 2 , in F CO 2 ,   out F CO 2 , in
nH 2 nCO = F H 2 , out F CO , out

5. Conclusions

This study demonstrated the catalytic performance results of strontium-incorporated CexSr1−xNiO3 (x = 0.6–1) nanocrystalline perovskites for dry methane reforming. The extent of finely dispersed nanocrystalline particles over the surface of the catalysts, their reducibility, and their textural properties, including specific surface area, were determined by characterizing the catalysts prior to the reforming reaction. It was revealed that these factors significantly influenced the catalytic activity as well as durability. Strontium-free CeNiO3 perovskite demonstrated higher activity in terms of CH4 and CO2 conversions but showed deactivation over time-on-stream. Despite relatively lower conversions, strontium incorporated perovskites (Ce0.8Sr0.2NiO3 and Ce0.6Sr0.4NiO3) remained stable over the 7 h time-on-stream. The lower activity in strontium-incorporated perovskites was attributed to the coverage of active (nickel) sites with carbonates of strontium, which is also reported in the literature. The post-reforming characterizations of perovskites helped to investigate the deactivation cause. Despite finding active metal agglomeration or sintering over all catalysts post-reforming, the insignificant sintering was ruled out as major cause of deactivation substantiated with no activity loss for strontium-incorporated catalysts. This suggests that carbon formation during the reforming reaction was the main cause of deactivation. This was further confirmed by analyzing TEM and TPO results of spent perovskites. Stable and/or increasing CO2 conversions higher than CH4 conversions in strontium-incorporated perovskites facilitated carbon gasification, which prevented deactivation. These results presented the key role of the strontium incorporation in CexSr1−xNiO3 (x = 0.6–1) perovskites to ensure stable catalytic performance for longer periods of time to avoid deactivation, a common challenge to cope with for nickel-based catalysts.

Supplementary Materials

The following supporting information can be downloaded, Table S1: Textural properties and deactivation factors of CexSr1−xNiO3 (x = 0.6–1) perovskites.

Author Contributions

N.A., F.A.A. and R.W. synthesized the catalysts, carried out all the experiments and characterization tests, and wrote the manuscript. B.F.A., S.M. and M.A. prepared the catalysts and contributed to proofreading the manuscript. N.A. and F.A.A. contributed to the analysis of the data and writing—review of the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by King Saud University under NPST project 14-PET851-02.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The authors would like to extend their sincere appreciation to King Saud University for its funding of this NPST project (14-PET851-02).

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Samples of the compounds CexSr1−xNiO3 (x = 0.6–1) perovskites are available from the authors.

References

  1. Wang, Y.; Yao, L.; Wang, S.; Mao, D.; Hu, C. Low-temperature catalytic CO2 dry reforming of methane on Ni-based catalysts: A review. Fuel Process. Technol. 2018, 169, 199–206. [Google Scholar] [CrossRef]
  2. Li, H.; He, Y.; Shen, D.; Cheng, S.; Wang, J.; Liu, H.; Xing, C.; Shan, S.; Lu, C.; Yang, R. Design an in-situ reduction of Ni/C-SiO2 catalyst and new insight into pretreatment effect for CH4-CO2 reforming reaction. Int. J. Hydrogen Energy 2017, 42, 10844–10853. [Google Scholar] [CrossRef]
  3. Al-Fatesh, A.S.; Arafat, Y.; Atia, H.; Ibrahim, A.A.; Ha, Q.L.M.; Schneider, M.; M-Pohl, M.; Fakeeha, A.H. CO2-reforming of methane to produce syngas over Co-Ni/SBA-15 catalyst: Effect of support modifiers (Mg, La and Sc) on catalytic stability. J. CO2 Util. 2017, 21, 395–404. [Google Scholar] [CrossRef]
  4. Ibrahim, A.A.; Fakeeha, A.H.; Al-Fatesh, A.S. Enhancing hydrogen production by dry reforming process with strontium promoter. Int. J. Hydrogen Energy 2014, 39, 1680–1687. [Google Scholar] [CrossRef]
  5. Li, D.; Lu, M.; Xu, S.; Chen, C.; Zhan, Y.; Jiang, L. Preparation of supported Co catalysts from Co-Mg-Al layered double hydroxides for carbon dioxide reforming of methane. Int. J. Hydrogen Energy 2017, 42, 5063–5071. [Google Scholar] [CrossRef]
  6. Al-Fatesh, A.S.; Arafat, Y.; Ibrahim, A.A.; Atia, H.; Fakeeha, A.H.; Armbruster, U.; Abasaeed, A.E.; Frusteri, F. Evaluation of Co-Ni/Sc-SBA–15 as a novel coke resistant catalyst for syngas production via CO2 reforming of methane. Appl. Catal. A Gen. 2018, 567, 102–111. [Google Scholar] [CrossRef]
  7. Rynkowski, J.; Samulkiewicz, P.; Ladavos, A.K.; Pomonis, P.J. Catalytic performance of reduced La2−xSrxNiO4 perovskite-like oxides for CO2 reforming of CH4. Appl. Catal. A Gen. 2004, 263, 1–9. [Google Scholar] [CrossRef]
  8. El Hassan, N.; Kaydouh, M.N.; Geagea, H.; El Zein, H.; Jabbour, K.; Casale, S.; El Zakhem, H.; Massiani, P. Low temperature dry reforming of methane on rhodium and cobalt based catalysts: Active phase stabilization by confinement in mesoporous SBA-15. Appl. Catal. A Gen. 2016, 520, 114–121. [Google Scholar] [CrossRef] [Green Version]
  9. Khan, W.U.; Li, X.; Baharudin, L.; Yip, A.C.K. Copper-promoted cobalt/titaniananorod catalyst for CO hydrogenation to hydrocarbons. Catal. Lett. 2021, 151, 2492–2501. [Google Scholar] [CrossRef]
  10. Khan, W.U.; Baharudin, L.; Choi, J.; Yip, A.C.K. Recent progress in CO hydrogenation over bimetallic catalysts for higher alcohol synthesis. ChemCatChem 2021, 13, 111–120. [Google Scholar] [CrossRef]
  11. Khan, W.U.; Chen, S.S.; Tsang, D.C.W.; Hu, X.; Lam, F.L.Y.; Yip, A.C.K. Catalytically active interfaces in titaniananorod-supported copper catalysts for CO oxidation. Nano Res. 2020, 13, 533–542. [Google Scholar] [CrossRef]
  12. Horn, R.; Schlögl, R. Methane Activation by Heterogeneous Catalysis. Catal. Lett. 2015, 45, 23–39. [Google Scholar] [CrossRef] [Green Version]
  13. Rostrup-Nielsen, J.R.; Sehested, J.; Nørskov, J.K. Hydrogen and synthesis gas by steam- and CO2 reforming. Adv. Catal. 2002, 47, 65–139. [Google Scholar] [CrossRef]
  14. Kang, D.; Yu, J.; Ma, W.; Zheng, M.; He, Y.; Li, P. Synthesis of Cu/Ni-La0.7Sr0.3Cr0.5Mn0.5O3–δ and its catalytic performance on dry methane reforming. J. Rare Earth 2019, 37, 585–593. [Google Scholar] [CrossRef]
  15. Wang, F.; Han, K.; Yu, W.; Zhao, L.; Wang, Y.; Wang, X.; Yu, H.; Shi, W. Low Temperature CO2 Reforming with Methane Reaction over CeO2-Modified Ni@SiO2 Catalysts. ACS Appl. Mater. Interfaces 2020, 12, 35022–35034. [Google Scholar] [CrossRef] [PubMed]
  16. Han, B.; Zhao, L.; Wang, F.; Xu, L.; Yu, H.; Cui, Y.; Zhang, J.; Shi, W. Effect of calcination temperature on performance of Ni@SiO2 catalyst in methane dry reforming. Ind. Eng. Chem. Res. 2020, 59, 13370–13379. [Google Scholar] [CrossRef]
  17. Chein, R.-Y.; Fung, W.-Y. Syngas production via dry reforming of methane over CeO2 modified Ni/Al2O3 catalysts. Int. J. Hydrogen Energy 2019, 44, 14303–14315. [Google Scholar] [CrossRef]
  18. Ma, Q.; Guo, L.; Fang, Y.; Li, H.; Zhang, J.; Zhao, T.-S.; Yang, G.; Yoneyam, Y.; Tsubaki, N. Combined methane dry reforming and methane partial oxidization for syngas production over high dispersion Ni based mesoporous catalyst. Fuel Process. Technol. 2019, 188, 98–104. [Google Scholar] [CrossRef]
  19. Abdullah, B.; Ghani, N.A.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] [Green Version]
  20. Ali, S.; Khader, M.M.; Almarri, M.J.; Abdelmoneim, A.G. Ni-based nano-catalysts for the dry reforming of methane. Catal. Today 2020, 343, 26–37. [Google Scholar] [CrossRef]
  21. Shen, J.; Reule, A.A.C.; Semagina, N. Ni/MgAl2O4 catalyst for low-temperature oxidative dry methane reforming with CO2. Int. J. Hydrogen Energy 2019, 44, 4616–4629. [Google Scholar] [CrossRef]
  22. Wanga, F.; Hana, B.; Zhanga, L.; Xub, L.; Yuc, H.; Shi, W. CO2 reforming with methane over small-sized Ni@SiO2 catalysts with unique features of sintering-free and low carbon. Appl. Catal. B Environ. 2018, 235, 26–35. [Google Scholar] [CrossRef]
  23. Han, K.; Yu, W.; Xu, L.; Deng, Z.; Yu, H.; Wang, F. Reducing carbon deposition and enhancing reaction stability by ceria for methane dry reforming over Ni@SiO2@CeO2 catalyst. Fuel 2021, 291, 120182. [Google Scholar] [CrossRef]
  24. Charisiou, N.D.; Siakavelas, G.; Papageridis, K.N.; Baklavaridis, A.; Tzounis, L.; Avraam, D.G.; Goula, M.A. Syngas production via the biogas dry reforming reaction over nickel supported on modified with CeO2 and/or La2O3 alumina catalysts. J. Nat. Gas Sci. Eng. 2016, 31, 164–183. [Google Scholar] [CrossRef]
  25. Goula, M.A.; Charisiou, N.D.; Siakavelas, G.; Tzounis, L.; Tsiaoussis, I.; Panagiotopoulou, P.; Goula, G.; Yentekakis, I.V. Syngas production via the biogas dry reforming reaction over Ni supported on zirconia modified with CeO2 or La2O3 catalysts. Int. J. Hydrogen Energy 2017, 42, 13724–13740. [Google Scholar] [CrossRef]
  26. Makri, M.M.; Vasiliades, M.A.; Petallidou, K.C.; Efstathiou, A.M. Effect of support composition on the origin and reactivity of carbon formed during dry reforming of methane over 5 wt% Ni/Ce1−xMxO2−δ (M= Zr4+, Pr3+) catalysts. Catal. Today 2016, 259, 150–164. [Google Scholar] [CrossRef]
  27. Summa, P.; Swirk, K.; Wierzbicki, D.; Motak, M.; Alxneit, I.; Rønning, M.; Da Costa, P. Co-Precipitated Ni-Mg-Al Hydrotalcite-Derived Catalyst Promoted with Vanadium for CO2 Methanation. Molecules 2021, 26, 6506. [Google Scholar] [CrossRef]
  28. Świrka, K.; Gálvez, M.E.; Motak, M.; Grzybek, T.; Rønning, M.; Da Costa, P. Dry reforming of methane over Zr- and Y-modified Ni/Mg/Al double-layered Hydroxides. Catal. Commun. 2018, 117, 26–32. [Google Scholar] [CrossRef]
  29. Świrk, K.; Gálvez, M.E.; Motak, M.; Grzybek, T.; Rønning, M.; Da Costa, P. Yttrium promoted Ni-based double-layered hydroxides for dry methane Reforming. J. CO2 Util. 2018, 27, 247–258. [Google Scholar] [CrossRef]
  30. Swirk, K.; Rønning, M.; Motak, M.; Beaunier, P.; Da Costa, P.; Grzybek, T. Ce- and Y-Modified Double-Layered Hydroxides as Catalysts for Dry Reforming of Methane: On the Effect of Yttrium Promotion. Catalysts 2019, 9, 56. [Google Scholar] [CrossRef] [Green Version]
  31. Jacobson, A.J. Materials for Solid Oxide Fuel Cells. Chem. Mater. 2010, 22, 660–674. [Google Scholar] [CrossRef]
  32. Kim, W.Y.; Jang, J.S.; Ra, E.C.; Kim, K.Y.; Kim, E.H.; Lee, J.S. Reduced perovskite LaNiO3 catalysts modified with Co and Mn for low coke formation in dry reforming of methane. Appl. Catal. A Gen. 2019, 575, 198–203. [Google Scholar] [CrossRef]
  33. Wei, T.; Jia, L.; Zheng, H.; Chi, B.; Pu, J.; Li, J. LaMnO3-based perovskite with in-situ exsolved Ni nanoparticles: A highly active, performance stable and coking resistant catalyst for CO2 dry reforming of CH4. Appl. Catal. A Gen. 2018, 564, 199–207. [Google Scholar] [CrossRef]
  34. Yang, Q.; Liu, G.; Yuan Liu, Y. Perovskite-type oxides as the catalyst precursors for preparing supported metallic nanocatalysts: A review. Ind. Eng. Chem. Res. 2018, 57, 1–17. [Google Scholar] [CrossRef]
  35. Bhattar, S.; Abedin, M.A.; Kanitkar, S.; Spivey, J.J. A review on dry reforming of methane over perovskite derived catalysts. Catal. Today 2021, 365, 2–23. [Google Scholar] [CrossRef]
  36. Khalesi, A.; Arandiyan, H.R.; Parvari, M. Production of Syngas by CO2 Reforming on MxLa1−xNi0.3Al0.7O3−d (M = Li, Na, K). Ind. Eng. Chem. Res. 2008, 47, 5892–5898. [Google Scholar] [CrossRef]
  37. Pérez-Camacho, M.N.; Abu-Dahrieh, J.; Goguet, A.; Sun, K.; Rooney, D. Self-cleaning perovskite-type catalysts for the dry reforming of methane. Chin. J. Catal. 2014, 35, 1337–1346. [Google Scholar] [CrossRef]
  38. Ruocco, C.; Caprariis, B.D.; Palma, V.; Petrullo, A.; Ricca, A.; Scarsella, M.; Filippis, P.D. Methane dry reforming on Ru perovskites, AZrRuO3: Influence of preparation method and substitution of A cation with alkaline earth metals. J. CO2 Util. 2019, 30, 222–231. [Google Scholar] [CrossRef]
  39. Wang, H.; Dong, X.; Zhao, T.; Yu, H.; Li, M. Dry reforming of methane over bimetallic Ni-Co catalyst prepared from La(CoxNi1-x)0.5Fe0.5O3 perovskite precursor: Catalytic activity and coking resistance. Appl. Catal. B Environ. 2009, 245, 302–313. [Google Scholar] [CrossRef]
  40. Wang, F.; Xu, L.; Shi, W. Syngas production from CO2 reforming with methane over core-shell Ni@SiO2 catalysts. J. CO2 Util. 2016, 16, 318–327. [Google Scholar] [CrossRef]
  41. Han, B.; Wang, F.; Zhang, L.; Wang, Y.; Fan, W.; Xu, L.; Yu, H.; Li, Z. Syngas production from methane steam reforming and dry reforming reactions over sintering-resistant Ni@SiO2 catalyst. Res. Chem. Intermed. 2020, 46, 1735–1748. [Google Scholar] [CrossRef]
  42. Dehghani, F.; Ayatollahi, S.; Bahadorikhalili, S.; Esmaeilpour, M. Synthesis and Characterization of Mixed–Metal Oxide Nanoparticles (CeNiO3, CeZrO4, CeCaO3) and Application in Adsorption and Catalytic Oxidation–Decomposition of Asphaltenes with Different Chemical Structures. Pet. Chem. 2020, 60, 731–743. [Google Scholar] [CrossRef]
  43. Harikrishnan, M.P.; Mary, A.J.C.; Bose, A.C. Electrochemical performance of ANiO3 (A= La, Ce) perovskite oxide material and its device performance for supercapattery application. Electrochim. Acta 2020, 362, 137095. [Google Scholar] [CrossRef]
  44. Chen, J.; He, Z.; Li, G.; An, T.; Shi, H.; Li, Y. Visible-light-enhanced photothermocatalytic activity of ABO3-type perovskites for the decontamination of gaseous styrene. Appl. Catal. B Environ. 2017, 209, 146–154. [Google Scholar] [CrossRef] [Green Version]
  45. Hu, Q.; Yue, B.; Shao, H.; Yang, F.; Wang, J.; Wang, Y.; Liu, J. Facile syntheses of cerium-based CeMO3 (M = Co, Ni, Cu) perovskite nanomaterials for high-performance supercapacitor electrodes. J. Mater. Sci. 2020, 55, 8421–8434. [Google Scholar] [CrossRef]
  46. García de la Cruz, R.M.; Falcón, H.; Peña, M.A.; Fierro, J.L.G. Role of bulk and surface structures of La1−xSrxNiO3 perovskite-type oxides in methane combustion. Appl. Catal. B Environ. 2001, 33, 45–55. [Google Scholar] [CrossRef]
  47. Wang, Y.; Cui, X.; Li, Y.; Shu, Z.; Chen, H.; Shi, J. A simple co-nanocasting method to synthesize high surface area mesoporous LaCoO3 oxides for CO and NO oxidations. Microporous Mesoporous Mater. 2013, 176, 8–15. [Google Scholar] [CrossRef]
  48. Wei, Y.; Zhao, Z.; Jiao, J.; Liu, J.; Duan, A.; Jiang, G. Facile synthesis of three-dimensionally ordered macroporous LaFeO3-supported gold nanoparticle catalysts with high catalytic activity and stability for soot combustion. Catal. Today 2015, 245, 37–45. [Google Scholar] [CrossRef]
  49. Sudhakaran, M.S.P.; Hossain, M.M.; Gnanasekaran, G.; Mok, Y.S. Dry Reforming of Propane over γ-Al2O3 and Nickel Foam Supported Novel SrNiO3 Perovskite Catalyst. Catalysts 2019, 9, 68. [Google Scholar]
  50. Pino, L.; Italiano, C.; Lagana, M.; Vita, A.; Recupero, V. Kinetic study of the methane dry (CO2) reforming reaction over the Ce0.70La0.20Ni0.10O2−δ catalyst. Catal. Sci. Technol. 2020, 10, 2652–2662. [Google Scholar] [CrossRef]
  51. Ostrovskii, N.M. Problems in the Study of Catalyst Deactivation Kinetics. Kinet. Catal. 2005, 46, 693–704. [Google Scholar] [CrossRef]
  52. Moral, A.; Reyero, I.; Alfaro, C.; Bimbela, F.; Gandía, L.M. Syngas production by means of biogas catalytic partial oxidation and dry reforming using Rh-based catalysts. Catal. Today 2018, 299, 280–288. [Google Scholar] [CrossRef]
  53. Verykios, X. Mechanistic aspects of the reaction of CO2 reforming of methane over Rh/Al2O3 catalyst. Appl. Catal. A Gen. 2003, 255, 101–111. [Google Scholar] [CrossRef]
  54. Choudhary, V.R.; Uphade, B.S.; Belhekar, A.A. Oxidative Conversion of Methane to Syngas over LaNiO3 Perovskite with or without Simultaneous Steam and CO2 Reforming Reactions: Influence of Partial Substitution of La and Ni. J. Catal. 1996, 163, 312–318. [Google Scholar] [CrossRef]
  55. Lima, S.M.; Assaf, J.M.; Peña, M.A.; Fierro, J.L.G. Structural features of La1−xCexNiO3 mixed oxides and performance for the dry reforming of methane. Appl. Catal. A Gen. 2006, 311, 94–104. [Google Scholar] [CrossRef]
  56. Valderrama, G.; Goldwasser, M.R.; de Navarro, C.U.; Tatibouët, J.M.; Barrault, J.; Batiot-Dupeyrat, C.; Martínez, F. Dry reforming of methane over Ni perovskite type oxides. Catal. Today 2005, 107–108, 785–791. [Google Scholar] [CrossRef]
  57. Das, S.; Bhattar, S.; Liu, L.; Wang, Z.; Xi, S.; Spivey, J.J.; Kawi, S. Effect of Partial Fe Substitution in La0.9Sr0.1NiO3 Perovskite-derived Catalysts on Reaction Mechanism of Methane Dry Reforming. ACS Catal. 2020, 10, 12466–12486. [Google Scholar] [CrossRef]
  58. Komarala, E.P.; Komissarov, I.; Rosen, B.A. Effect of Fe and Mn Substitution in LaNiO3 on Exsolution, Activity, and Stability for Methane Dry Reforming. Catalysts 2020, 10, 27. [Google Scholar] [CrossRef] [Green Version]
  59. Mousavi, M.; Pour, A.N.; Gholizadeh, M.; Mohammadi, A.; Shahri, S.M.K. Dry Reforming of Methane by La0.5Sr0.5NiO3 Perovskite Oxides: Influence of Preparation Method on Performance and Structural Features of the Catalysts. J. Chem. Technol. Biotechnol. 2020, 95, 2911–2920. [Google Scholar] [CrossRef]
  60. Voorhoeve, R.J.H. 5—Perovskite-Related Oxides as Oxidation—Reduction Catalysts. In Advanced Materials in Catalysis; Burton, J.J., Garten, R.L., Eds.; Academic Press: New York, NY, USA, 1977; pp. 129–180. [Google Scholar]
  61. Messaoudi, H.; Thomas, S.; Djaidja, A.; Slyemi, S.; Barama, A. Study of LaxNiOy and LaxNiOy/MgAl2O4 catalysts in dry reforming of methane. J. CO2 Util. 2018, 24, 40–49. [Google Scholar] [CrossRef]
Figure 1. TG−DTG curves versus temperature of CexSr1−xNiO3 (x = 0.6–1) perovskites.
Figure 1. TG−DTG curves versus temperature of CexSr1−xNiO3 (x = 0.6–1) perovskites.
Molecules 27 00356 g001
Figure 2. XRD patterns of CexSr1−xNiO3 (x = 0.6–1) perovskites.
Figure 2. XRD patterns of CexSr1−xNiO3 (x = 0.6–1) perovskites.
Molecules 27 00356 g002
Figure 3. N2 adsorption–desorption isotherms of CexSr1−xNiO3 (x = 0.6–1) perovskites.
Figure 3. N2 adsorption–desorption isotherms of CexSr1−xNiO3 (x = 0.6–1) perovskites.
Molecules 27 00356 g003
Figure 4. TEM images of CexSr1−xNiO3 (x = 0.6–1) perovskites; fresh catalysts: (a) CeNiO3, (b) Ce0.8Sr0.2NiO3, (c) Ce0.6Sr0.4NiO3; reduced catalysts: (d) CeNiO3, (e) Ce0.8Sr0.2NiO3, (f) Ce0.6Sr0.4NiO3; spent catalysts: (g) CeNiO3, (h) Ce0.8Sr0.2NiO3, (i) Ce0.6Sr0.4NiO3.
Figure 4. TEM images of CexSr1−xNiO3 (x = 0.6–1) perovskites; fresh catalysts: (a) CeNiO3, (b) Ce0.8Sr0.2NiO3, (c) Ce0.6Sr0.4NiO3; reduced catalysts: (d) CeNiO3, (e) Ce0.8Sr0.2NiO3, (f) Ce0.6Sr0.4NiO3; spent catalysts: (g) CeNiO3, (h) Ce0.8Sr0.2NiO3, (i) Ce0.6Sr0.4NiO3.
Molecules 27 00356 g004
Figure 5. H2-TPR profiles of (a) CeNiO3, (b) Ce0.8Sr0.2NiO3, and (c) Ce0.6Sr0.4NiO3 perovskites.
Figure 5. H2-TPR profiles of (a) CeNiO3, (b) Ce0.8Sr0.2NiO3, and (c) Ce0.6Sr0.4NiO3 perovskites.
Molecules 27 00356 g005
Figure 6. (a) CH4 conversion, (b) CO2 conversion, and (c) H2/CO ratios versus time-on-stream (TOS) of CexSr1−xNiO3 (x = 0.6–1) perovskites.
Figure 6. (a) CH4 conversion, (b) CO2 conversion, and (c) H2/CO ratios versus time-on-stream (TOS) of CexSr1−xNiO3 (x = 0.6–1) perovskites.
Molecules 27 00356 g006
Figure 7. TPO profiles of (a) CeNiO3, (b) Ce0.8Sr0.2NiO3, and (c) Ce0.6Sr0.4NiO3 perovskites.
Figure 7. TPO profiles of (a) CeNiO3, (b) Ce0.8Sr0.2NiO3, and (c) Ce0.6Sr0.4NiO3 perovskites.
Molecules 27 00356 g007
Figure 8. Schematic diagram of the reaction mechanism over Sr-free and Sr-incorporated perovskites.
Figure 8. Schematic diagram of the reaction mechanism over Sr-free and Sr-incorporated perovskites.
Molecules 27 00356 g008
Table 1. Textural properties and deactivation factors of CexSr1−xNiO3(x = 0.6–1) perovskites.
Table 1. Textural properties and deactivation factors of CexSr1−xNiO3(x = 0.6–1) perovskites.
PerovskitesNi/Ce/Sr Content (%) aSBET
(m2/g)
Pore Volume
(cm3/g)
Pore size
(nm)
Deactivation Factor (%)Coke (wt%) d
FreshUsed
CeNiO349.8/50.1/-50.2/49.6/-20.70.16230.17.7 b (0.88) c9.1
Ce0.8Sr0.2NiO351.1/44.1/4.750.9/43.8/5.325.60.26140.8−62.6 b (1.71) c4.7
Ce0.6Sr0.4NiO350.7/32.1/17.251/31.1/17.817.30.16438.1−13.4 b (1.15) c2.1
a Determined from ICP-OES. b Deactivation factor (D.F., %) = 100 × (CH4conversioninitial − CH4conversionfinal)/(CH4conversioninitial). c D.F based on first-order deactivation, % = ln(1 − CH4conversionfinal)/ln(1 − CH4conversioninitial). d Calculated from TPO data.
Table 2. Comparison of current work with previously reported work. DF refers to deactivation factor. TOS indicates time-on-stream.
Table 2. Comparison of current work with previously reported work. DF refers to deactivation factor. TOS indicates time-on-stream.
CatalystReaction Temp. (°C)/GHSV (L/h/gcat)Highest CH4
Conversion (%)
%DF a TOS (h)Ref.
CeNiO3750/723212.9~26[55]
La0.6Sr0.4NiO3700/-85-20[56]
La0.9Sr0.1NiO3700/18703.68[57]
LaNi0.8Fe0.2O3800/13.7657.735[58]
La0.5Sr0.5NiO3750/186947.824[59]
CeNiO3700/84557.78This work
Ce0.8Sr0.2NiO322.5−62.6
Ce0.6Sr0.4NiO322.5−13.4
a Deactivation Factor (D.F., %) = 100 × (CH4conversioninitial − CH4conversionfinal)/(CH4conversioninitial).
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Ahmad, N.; Wahab, R.; Manoharadas, S.; Alrayes, B.F.; Alam, M.; Alharthi, F.A. The Role of Strontium in CeNiO3 Nano-Crystalline Perovskites for Greenhouse Gas Mitigation to Produce Syngas. Molecules 2022, 27, 356. https://doi.org/10.3390/molecules27020356

AMA Style

Ahmad N, Wahab R, Manoharadas S, Alrayes BF, Alam M, Alharthi FA. The Role of Strontium in CeNiO3 Nano-Crystalline Perovskites for Greenhouse Gas Mitigation to Produce Syngas. Molecules. 2022; 27(2):356. https://doi.org/10.3390/molecules27020356

Chicago/Turabian Style

Ahmad, Naushad, Rizwan Wahab, Salim Manoharadas, Basel F. Alrayes, Manawwer Alam, and Fahad A. Alharthi. 2022. "The Role of Strontium in CeNiO3 Nano-Crystalline Perovskites for Greenhouse Gas Mitigation to Produce Syngas" Molecules 27, no. 2: 356. https://doi.org/10.3390/molecules27020356

Article Metrics

Back to TopTop