Next Article in Journal
Development of Improved Confined Compression Testing Setups for Use in Stress Relaxation Testing of Viscoelastic Biomaterials
Previous Article in Journal
Ceramic Fiber-Reinforced Polyimide Aerogel Composites with Improved Shape Stability against Shrinkage
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Study of the Synthesis of Multi-Cationic Sm-Co-O, Sm-Ni-O, Al-Co-O, Al-Ni-O, and Al-Co-Ni-O Aerogels and Their Catalytic Activity in the Dry Reforming of Methane

1
Central European Institute of Technology, Brno University of Technology, Purkyňova 123, 612 00 Brno, Czech Republic
2
Institute of Materials Science and Engineering, Brno University of Technology, Technicka 2, 616 69 Brno, Czech Republic
3
Institute of Rock Structure and Mechanics of the Czech Academy of Sciences, V Holešovičkách 94/41, 182 09 Praha, Czech Republic
*
Author to whom correspondence should be addressed.
Gels 2024, 10(5), 328; https://doi.org/10.3390/gels10050328
Submission received: 21 April 2024 / Revised: 4 May 2024 / Accepted: 8 May 2024 / Published: 11 May 2024
(This article belongs to the Special Issue Chemical Properties and Application of Gel Materials)

Abstract

:
Dense multi-cationic Sm-Co-O, Sm-Ni-O, Al-Co-O, Al-Ni-O, and Al-Ni-Co-O oxide aerogels were prepared by epoxide-driven sol–gel synthesis. Catalysts for dry reformation of methane, Sm2O3/Co, Sm2O3/Ni, Al2O3/Co, Al2O3/Ni, Al2O3/Co, and Ni were prepared by reduction of aerogels with hydrogen and their catalytic activities and C-deposition during dry reformation of methane were tested. Catalytic tests showed high methane conversion (93–98%) and C-deposition (0.01–4.35 mg C/gcat.h). The highest content of C-deposits after catalytic tests was determined for Al2O3/Co and Al2O3/Ni catalysts, which was related to the formation of Al alloys with Co and Ni. A uniform distribution of Co0 and Ni0 nanoparticles (in the form of a CoNi alloy) was found only for the Al2O3/Co and Ni catalysts, which showed the highest activity as well as low C deposition.

1. Introduction

Climate change is mainly related to burning fossil fuels that produce carbon dioxide. Another greenhouse gas is methane, which comes mainly from fossil sources. One of the essential ways of reducing emissions of these greenhouse gases, CO2 and CH4, is dry reformation of methane (DRM), whose important product is hydrogen or synthesis gas. Due to its ecological and economic benefits, many professional reviews and scientific works are devoted to the dry reformation of methane [1]. These works aim to bring new catalytic systems with high catalytic activity at the lowest possible temperature, which are long-term stable, have low C-deposition, and are affordable [2]. These requirements are currently best met by heterogeneous catalytic systems containing catalytically active particles of transition metals, especially Ni and Co, deposited on the surfaces of inorganic oxides [3,4]. Their catalytic efficiency depends on the interplay of acid–base and redox properties of metal particles and oxide substrates during the transfer of electrons and oxide ions between the reactants participating in DRM (CH4, CO2, H2, and CO). The structural properties of reforming catalysts are no less critical, such as surface size, particle size, the composition of active centers, and their localization on substrate surfaces [5]. The chemical and phase composition, architecture, and surface morphology of reforming catalysts depend on their preparation technology. Among the most valuable are the sol–gel methods, which allow control of the homogeneity of the reactants and the dispersions that arise from them. It is generally considered that the homogeneity of dispersions is one of the prerequisites for forming homogeneously distributed active centers on the surface of heterogeneous catalysts [6]. Aerogels are three-dimensional porous materials with interconnected micropores, often prepared by sol–gel methods and supercritical gel drying, which can preserve the large surface texture of the raw gel [7]. Inorganic oxide aerogels containing several cations can be transformed by hydrogen reduction into heterogeneous catalysts composed of an oxide substrate and an active metal. Achieving the transition of a sol (dispersion) to a polymer gel in multi-cationic systems and creating a sufficiently homogeneous distribution of cations in the gel is not easy and depends primarily on the course of hydrolysis and condensation of reactants during gelation [8].
The work is focused on the study of di-cationic aerogels composed of two oxides, one of which is “irreducible” with hydrogen and serves as a substrate and the other is “reducible”, which is excreted on the surface of the substrate through reduction and forms active centers for dry reformation of methane. The work’s primary goal is to prepare di-cationic aerogels, where both cations are “homogeneously” dispersed in the aerogel structure. During reduction, the reducible cation diffuses through the aerogel structure and is reduced by hydrogen to an active metal at the surface of the substrate. The composition of di-cationic aerogels is designed so that, by reduction, catalytic systems containing oxide substrates (Sm2O3 and Al2O3) with different interactions with the most catalytically active general transition elements (Co and Ni) are created. The three-cationic aerogel Al-Ni-Co-O was included in the work in order to verify the higher activity and lower C-deposition compared to the di-cationic Al-Co-O and Al-Ni-O catalysts. The work contains the study results of epoxide-driven synthesis of dense oxide aerogels Sm-Co-O, Sm-Ni-O, Al-Co-O, Al-Ni-O, and Al-Ni-Co-O. The work also includes the results of the study of the reduction in aerogels with hydrogen to reforming catalysts, Sm2O3/Co, Sm2O3/Ni, Al2O3/Co, Al2O3/Ni, Al2O3/Co, Ni and testing of their catalytic activity and C deposition during dry methane reforming. We see the benefit of the work in a new perspective on C-deposition in aerogel Al2O3/Co and Al2O3/Ni catalysts.

2. Results and Discussion

2.1. Synthesis and Characterization of Aerogels

Sm, Al, Co, and Ni chlorides with six crystalline water molecules were used to synthesize aerogels. Chlorides with the mentioned cations had an octahedral configuration in which water molecules are bound in the inner-sphere of the aqua complex and chloride ions lie in the outer-sphere of the aqua complex (x = 2, 3) [M(H2O)6]x+· xCl [8,9,10,11,12].
All cations used have a high charge/radius ratio (charge density) and their hydrated cations (aqua cationic complex) in aqueous-alcoholic solutions were subject to hydrolysis according to Equation (1) [13], as follows:
[M(H2O)6]r+ + H2O = [M(OH)(H2O)5](r−1)+ + H3O+(5) (M = Co2+, Ni2+, Al3+, Sm3+)
The greater the charge of the cation and the smaller the size of the ion, the more easily the ion polarizes coordinated water molecules. This weakens the OH bond in the water molecule and releases the H+ (or H3O+) ion into the solution. The charge density of cations (C/mm3) increases in the order Sm(86) < Ni(134) < Co(155) < Al(364) [14,15]. The hydrolysis and acidity of chloride solutions of these cations increase in the same order. By pumping out protons (H3O+) from the hydrolyzed aqua cationic complexes, olation and the formation of μ-hydroxo-bridged complexes first occur according to Equation (2), as follows:
2[M(H2O)5(OH)](r−1)+ + H+ = [(H2O)5M-OH-M(H2O)5] (2r−1)+ + H2O
By pumping out protons from μ-hydroxo-bridged metal complexes, oxolation and the formation of μ-oxo-bridged metal complexes occurs according to Equation (3), as follows:
[(H2O)5M-OH-M(H2O)5](2r−1)+ = [(H2O)5M-O-M(H2O)5](2r−2)+ + H+
Slow hydrolysis (deprotonation), olation, and oxolation (condensation) lead to the formation of polymeric species in gels. In the epoxide-driven method, the epoxide (propylene oxide) acts as a base, slowly reacting by the epoxide ring-opening mechanism with protons released by hydrolysis and oxolation, thus promoting the creation of gels according to Equation (4) [16,17], as follows:
Gels 10 00328 i001
In multication systems, namely Sm-Co-O, Sm-Ni-O, Al-Co-O, Al-Ni-O, and Al-Ni-Co-O as well as not only Sm-O-Co, Sm-O-Ni, Al-O-Co, Al-O-Ni-Co bonds but also “homo-metallic” Mi-O-Mi bonds, can be formed by olation and oxolation. Different rates of hydrolysis and condensation of individual aqua-cationic species can lead to inhomogeneities in gels and phase separation in multimetallic aerogels [17].
The precursor chlorides of Sm, Al, Co, and Ni formed in epoxy-driven synthesis of dense and visually homogeneous aerogels (Figure 1).
Hydrolysis and condensation rates of precursors of aerogels were manifested on values gelation times (tgel) listed in Table 1. The longest tgel was observed at gelation di-cationic gels containing Sm3+ cations; the shortest times were found for gels containing Al3+ cations. These findings are consistent with the charge density values of Sm3+ and Al3+ cations (Table 1). If the values of tgel given in Table 1 are compared with the published values for SmCl3 (12 min, [18]) and AlCl3 (3 min, [19]), it can be seen that the gelation times of gels were affected primarily by chlorides containing Sm3+ and Al3+ ions.
The results of simultaneous TGA/MS analysis of supercritically dried Sm-Al-Co-Ni-O aerogels are shown in Figure 2 and Table 2. In the temperature range of 50–220 °C, the aerogels lost 6–15% of their mass due to the evaporation of physically adsorbed water; see endothermic peaks (Q)DSC = −122 to −254 J/g and peaks m/z = 18 (H2O) and (Tmax)H2O = 123 to 144 °C. At a temperature of 220 to 600 °C, the weight loss of the aerogel was 10 to 22%; see exothermic peaks (Q)DSC = 157 to 642 J/g and peaks m/z = 18 (Tmax)H2O = 273 to 317 °C. Peaks of CO2 (m/z = 44) were detected in the temperature range 220 to 600 °C associated with the oxidation of organic molecules formed by the reaction of propylene oxide with chloride precursors (see exothermic peaks (Q)DSC and (Tmax)CO2 = 288 to 359 °C. In the MS spectra of gaseous products of thermal oxidation of aerogels, traces of Cl were also detected, which were released in the form of HCl (m/z = 36) from M-Cl groups or chloropropyl alcohol. It follows from the STA/MS analysis that the multi-cationic dried aerogels contained hydroxides-oxides of Sm, Al, Ni, Co, and the content of M-OH groups, which was in the range of 1.2 to 2.5 mol/100g of aerogel, did not depend on the precursors used.
The phase composition of calcined aerogels (T = 900 °C/10 h) can be seen in Figure 3. Except for the single-phase SmCoO3-δ perovskite, calcination resulted in two-phase or three-phase products (Al-Co-Ni-O). Spinel phases appeared in Al-Co-O and Al-Ni-O aerogels. Calcination was performed on uncrushed samples at a relatively short exposure time to verify the homogeneity in the distribution of cations in aerogels prepared by epoxy-driven synthesis. It was assumed that the formation of single-phase products would manifest the homogeneous composition of the dried aerogels. If we neglect the different activation energies of high-temperature reactions of mixed oxides (Sm-Co-O, Sm-Ni-O, Al-Co-O, Al-Ni-O, and Al-Ni-Co-O), the formation of multiphase products indicates an inhomogeneous distribution of cations in aerogels leading to phase separation.
The results of BET analysis of the surface of aerogels heat-treated at 500 °C/h and calcined at 900 °C/10 h are shown in Figure 4 and Table 3. Gas adsorption experiments revealed that powders after both thermal treatments exhibited similar features of physisorption behavior. All adsorption/desorption isotherms can be classified as Type II isotherms [20]. Such isotherms are associated with an open and stable external surface of non-porous or macroporous powders. Small hysteresis loops (Type H3) observed on some isotherms (Sm-Co-O, Sm-Ni-O) resulted from a substantial agglomeration of primary particles. The particle agglomerates exhibited macropores that were not filled by the condensed adsorbent. The specific surface area of individual samples is shown in Table 2. Heat-treated aerogels at a relatively low temperature (500 °C/1 h) had a very high surface area of up to 470 m2/g. The aerogels did not have a mesoporous structure and were formed by agglomerates of hydroxide-oxide nanoparticles. The surface area decreased substantially after calcination at 900 °C/10 h due to particle coarsening and sintering.

2.2. Reduction in Calcined Aerogels

Since the catalytically active sites must contain metal particles (Co0, Ni0, Ni0Co0), the calcined aerogels were activated by reduction. Reducibility aerogels were studied by TPR-H2, and the results are shown in Figure 5. TPR-H2 aerogel profiles Sm-Co-O contains two reduction peaks, Tmax = 462 and 523 °C. The first peak corresponds to the reduction Co3+ → Co2+ and the second peak corresponds to the reduction Co2+ → Co0 [21,22,23,24]. The TPR-H 2 profile of the Sm-Ni-O aerogel has two peaks, a low-temperature one (Tmax = 402 °C), which is related to the reduction in Ni3+ to Ni2+, and a high-temperature one (Tmax = 537 °C), which corresponds to the reduction Ni2+ to Ni0. The reduction behavior of Sm-Ni-O is similar to that of La-Ni-O [25,26,27]. Aerogels containing Al: Al-Co-O, Al-Ni-O and Al-Ni-Co-O differed from aerogels containing Sm by a significantly higher temperature of the second reduction peak. While the first reduction peak, corresponding to the reduction in Co3+, Ni3+ to Co2+ and Ni2+, was similar to the case of Sm aerogels in the range of 368 to 483 °C, the second reduction peak corresponding to the reduction in Co2+, Ni2+ → Co0, Ni0 reached values of 847 to 903 °C. The high reduction temperatures of Al-containing aerogels showed that Co and Ni ions were tightly bound in Al-containing phases (spinels Al-Ni-O, Al-Co-O, Al-Co-Ni-O). From this point of view, the binding of Co and Ni ions in the perovskite phases was weaker.
Figure 6 shows the field emission scanning electron microscopy (FESEM) and elemental mapping of reduced aerogels. The images show round particles of the substrates (Sm2O3, Al2O3) with a size of several micrometers and micrometer-sized pores formed by sintering during the reduction in aerogels. The distribution of Co0 on the surface of Sm-Co-O and Al-Co-O aerogels is significantly heterogeneous. In addition to small, several nanometer-sized nanoparticles uniformly deposited on the surface, spherical particles or clusters of 500–1000 nm size are present. A similar distribution of Ni0 is seen on the surface of Al-Ni-O and Sm-Ni-O aerogels. A uniform distribution of Co0 and Ni0 nanoparticles (probably in the form of a CoNi alloy) without large spherical particles is evident in the Al-Co-Ni-O aerogel. Large particles could form in two-phase (Sm-Ni-O, Al-Co-O, Al-Ni-O) aerogels so that one phase was reduced more easily and quickly (NiO, Co3O4) than the other (spinel phase). In the case of a single-phase aerogel (Sm-Co-O), the seeds could be formed in the low-temperature phase of the reduction. The practically homogeneous distribution of CoNi nanoparticles on the surface of the substrate is attributed to the interaction of Co0 and Ni0 nuclei during reduction and their growth at a constant rate without forming agglomerates. Elemental compositions of the reduced aerogels were qualitatively investigated using SEM-EDX and the results are listed in Table 4. These results confirm that aerogels comprise their major chemical elements and do not contain any other elements.

2.3. Catalytic Performance of Reduced Aerogels

The study of catalytic activity and C-deposition in the DRM process catalyzed by monometallic and bimetallic aerogels was used at two time and temperature programs. A mixture of gases CO2/CH4 = 3.8/3.2 mL/min (1.2/1.0) flowed through the bed of 100 mg of catalyst “gas space velocity” (GHSV) was 36 l/gcat.h. The composition of reactants (CO2 and CH4) and reaction products (H2, CO, and H2O) was monitored by MS. The catalytic performance of catalysts during dry reformation of CH4 expressed by the conversion of CH4 and CO2 is shown in Figure 7 and Table 5. From Figure 7, it can be seen that the DRM reaction was initiated on catalysts containing Ni (Sm-Ni-O, Al-Ni-O, and Al-Ni-Co-O) at a temperature of around 350 °C and on catalysts containing Co (Al-Co-O and Sm-Co-O) at a temperature of 400 °C to 500 °C. The conversion of X(CH4) increased rapidly when temperature increased from 350 °C to 600 °C and it reached values of up to 100% at 800 °C. The Sm-Co-O catalyst presented lower activity at temperatures below 750 °C due to its shallow surface area (1.51 m2/g). The catalytic parameters of the other catalysts were not significantly unaffected by the surface area. There was no significant difference between the values of X(CH4), X(CO2), Y(H2), and Y(CO) of the Al-Ni-O catalyst with a surface area of 68.7 m2/g and the Al-Co-O catalyst with a surface area of 11.9 m2/g. Instead, the catalytic activity was determined by the density and distribution of the metal particles on the surface, which was similar to these catalysts. The epoxide-driven method, probably at the molecular level (formation of Mi-O-Mj bonds), led to a similar structure of aerogels and reduced catalysts prepared from them. Catalytic tests showed that all aerogel catalysts had excellent activity. The high conversion values of X(CH4) and X(CO2) catalysts with Ni and Co deposited on the surface of the Al2O3 substrate (Al-Co-O, Al-Ni-O, and Al-Ni-Co-O) were comparable to those reported in the literature [3,28,29,30,31,32].
The catalytic activity and deactivation of aerogel catalysts can be seen in Figure 8, which shows the time dependence of conversion (X(CH4) and X(CO2)) and yield (Y(H2) and Y(CO)) at three gradually decreasing temperatures (800 °C/4h + 750 °C/5h + 700 °C/5h). CH4 and CO2 conversions decreased with temperature from 100% to 98% (or 90% in the case of Sm-Co-O) but the time decrease in individual steps was less than 1%. The reason for such minor changes in activity was primarily the short period of DRM testing. The stable behavior of aerogel catalysts under DRM is probably related to the controlled sol–gel synthesis producing Mi-O-Mj mixed bonds.

2.4. Characterization of Spent Catalysts

Temperature-programmed oxidation by oxygen (TPO) profiles of aerogel catalysts after DRM tests (Figure 9) show oxidation peaks at temperatures around 275–335 °C, 474–496 °C and 532–600 °C associated with oxidation of carbonaceous species deposited on the surface catalysts with oxygen. Depending on the temperature, the peaks are assigned to the oxidation of high amorphous, low amorphous, and graphitic carbons [33,34]. The content of the deposit on the surface of the catalyst increased in the order Sm-Co-O < Sm-Ni-O < Al-Co-Ni-O < Al-Ni-O < Al-Co-O (Table 6). C-deposits’ formation depends on the rates of C-deposit formation and its gasification, which is related to the character of the metal and the substrate [35]. The highest C-deposit (2.93 and 4.47 mg C/gcat.h) was found for monometallic metals deposited on Al2O3 substrates (Al-Ni-O and Al-Co-O). Since the reduction temperature of the high-temperature peak of both aerogels was high (around 900 °C), a strong interaction between Co and Ni and the Al2O3 substrate can be assumed [36], which should lead to a decrease in the rate of C-deposition on the catalyst surface. Therefore, the cause of the significant C-deposition in these two catalytic systems probably lies in the composition of the metal particles. XPS spectra showed the presence of NiAl alloy in Al-Ni-O and Al-Co-O samples. Due to the presence of metallic Al, a more robust interaction with CO2 could occur on the surfaces of NiAl and CoAl alloys, leading to a decrease in the rate of CO2 reduction to CO. A higher rate of CH4 cracking than CO2 reduction led to higher C-deposit production. The very low C-deposition in catalysts containing Sm (Sm-Co-O and Sm-Ni-O) was connected, similarly to La2O3, to the basic nature of Sm2O3 [37,38] and the high mobility of oxygen on the active center [37,39]. Electron-deficient Co particles, in combination with a Sm2O3 carrier, supported the reduction in CO2 and the oxidation of C-deposits to CO [27]. The very low C-deposition in the Al-Ni-Co-O catalyst agrees with the published results. The high activity, stability, and low C-deposition of the bimetallic Al-Ni-Co-O catalyst is attributed to the synergistic effect between Co and Ni in the CoNi alloy [40,41]. The synergistic effect resulted in the formation of homogeneously distributed CoNi nanoparticles on the surface of the Al2O3 substrate and a favorable interaction between CoNi and the substrate. The CH4 cracking rate on the CoNi surface was lower than the CO2 reduction rate, retarding the C-deposition [42].
XPS analysis of aerogel catalysts after DRM tests confirmed the presence of NiCo alloy in the Al-Co-Ni-O sample based on the difference in binding energy Ni 2p3/2 and Co 2p3/2 between mono-metallic samples Al-Ni-O, Al-Co-O, and bimetallic Al-Co-Al-O. The difference in Ni 2p3/2 binding energies between Al-Ni-O and Al-Co-Al-O was −0.2 eV (857.6 vs. 857.4 eV). The difference between the Co 2p3/2 binding energy was +0.45 eV (782.35 vs. 782.8 eV). The migration of Ni 2p3/2 peak to higher binding energy and Co 2p3/2 peak to higher binding energy indicates electron migration between Ni and Co and confirms the interaction between Co and Ni and the formation of NiCo alloy [43]. Figure 10 shows the XPS spectra of Al-Co-O and Al-Ni-O catalysts after the DRM test containing peaks of 61.8 eV and 68.3 eV assigned to AlCo and AlNi alloys. Due to the difficult reducibility of Al3+ with hydrogen, the mentioned alloys were probably only formed during DRM tests at high temperatures (800 °C) by the action of reactive carbon species (CxHy, CO, and C). In contrast to the “anti-coking” effect of the Al-Ni-Co-O catalyst containing the NiCo alloy, the Al-Co-O and Al-Ni-O catalysts produced the highest amount of C-deposit. AlCo and AlNi alloys containing reactive Al can form a stronger bond with CO2, retarded CO release, and oxidation of C-deposit on the catalyst surface.
C(1s) XPS spectra of aerogel catalysts are shown in Figure 11. Deconvoluted peaks centered at the binding energy 284.8–285.1 eV were assigned to C=C and CH groups; peaks at binding energy 285.6–286.7 eV were assigned to functional groups C-OH and C-O-C and peaks 288.8–289.6 eV to CO-O and CO2 groups [44,45,46,47]. The relative content of carbon deposit (CC, CH) on Sm-Al-Co-Ni-O catalysts increased in the order Al-Co-O > Al-Ni-Co-O > Al-Ni-O > Sm-Co-O > Sm-Ni-O. The relative content of oxidized C-products (COH, COC, and COO) increased in the opposite order. The highest content (COH, C-O-C, and COO) associated with Sm2O3 substrates shows that similar to La2O3, Sm2O3 participated, due to its basic and redox nature, in the elimination of CxHy species from the surface of the catalysts by interaction with CO2 [37,38,39].

3. Conclusions

Epoxide-driven sol–gel synthesis of multi-cationic aerogels led to dense di- and tri-cationic Sm-Co-O, Sm-Ni-O, Al-Co-O, Al-Ni-O, and Al-Ni-Co-O aerogels. In the course of hydrolysis and condensation (olation and oxolation) of aqua-cationic complexes in the water-ethanolic solution formed not only bonds Sm-O-Co, Sm-O-Ni, Al-O-Co, Al-O-Ni but also Mi-O-Mi bonds which led along the thermal processing to multiphase products. Aerogels heat-treated at a relatively low temperature (500 °C/1 h) had a very high surface area of up to 470 m2/g, which decreased more than ten times after calcination at 900 °C/10 h. The aerogels did not have a mesoporous structure and were formed by agglomerates of hydroxide-oxide nanoparticles of Co2+, Ni2+, Al3+ and Sm3+. Reduction in di-cationic Sm-Co-O, Sm-Ni-O, Al-Co-O, and Al-Ni-O aerogels resulted in catalysts with heterogeneous distribution of Co0 and Ni0 nanoparticles and microparticles on the surface of Sm2O3 and Al2O3. In addition to tiny, several nanometer-sized nanoparticles uniformly deposited on the surface, spherical particles or clusters of 500–1000 nm size were present. Large particles could form in two-phase Sm-Ni-O, Al-Co-O, and Al-Ni-O catalysts, in which one phase was reduced more easily faster (NiO and Co3O4) than the other (spinel phase). A uniform distribution of Co0 and Ni0 nanoparticles (in the form of a CoNi alloy) without large spherical particles was found only in the Al-Co-Ni-O catalyst. Catalytic tests showed that all aerogel catalysts had excellent activity. The epoxide-driven method probably, at the molecular level (formation of Mi-O-Mj bonds), led to a similar structure of aerogels and catalysts prepared from them. The content of C-deposit on the surface of the tested catalysts increased in the order Sm-Co-O < Sm-Ni-O < Al-Co-Ni-O < Al-Ni-O < Al-Co-O. The most C-deposit was found for metals, Co0 and Ni0, lying on the surface of Al2O3 substrates (Al-Ni-O and Al-Co-O). Therefore, the cause of the significant C-deposition in these two catalytic systems probably lies in the composition of the metal particles. XPS spectra showed the presence of NiAl alloy in Al-Ni-O and CoAl alloy in Al-Co-O samples. Due to the presence of metallic Al, strong interaction with CO2 could occur on the surfaces of NiAl and CoAl alloys, leading to a retarding of CO2 reduction. A higher rate of CH4 cracking than CO2 reduction led to higher C-deposit production. The low C-deposition in catalysts containing Sm (Sm-Co-O and Sm-Ni-O) was connected, similarly to La2O3, to the basic nature of Sm2O3 and the high mobility of oxygen on active centers.

4. Materials and Methods

4.1. Synthesis of Sm-Co-O, Sm-Ni-O, Al-Co-O, Al-Ni-O, and Al-Ni-Co-O Aerogels

Aerogels were prepared using the sol–gel method based on a ring-opening polymerization mechanism. Transition metal chlorides (CoCl2.6H2O and NiCl2.6H2O) in combination with samarium (SmCl3.6H2O) or aluminum chloride (AlCl3.6H2O) were used as the starting reactants and were dissolved in ethanol. The molar ratio of the reactants and the concentration of solutions are given in Table 1. Initially, solutions were prepared and stirred for 1 h. Then, propylene oxide was addedand the mixtures were stirred for several minutes and left in the vessels to gel. Gel times are given in Table 1. As a result, jelly samples were aged for 24 h in vessels and then in an acetone environment following supercritical drying, obtaining the aerogels. Aerogels were thermal treated (air, 500 °C/1 h), calcined (air, 900 °C/10 h) and reduced (5 vol.%H2 in Ar, 800 °C/1 h) afterward.

4.2. Characterization of Aerogels

Aerogels were evaluated via nitrogen adsorption, X-ray diffraction (XRD), X-ray photoelectron spectroscopy (XPS), temperature-programmed reduction by hydrogen (H2-TPR), and temperature-programmed oxidation by oxygen (O2-TPO). The phase composition of aerogel particles before and after reduction with H2 was determined by the RTG diffractometer (Philips X’pert) in a central focusing configuration, using CoKa radiation and two types of detectors, the X’celerator and the Microprobe. The corresponding standards were established by comparing the ICSD (Inorganic Crystal Structure Database, FIZ). The crystallographic structures and quantitative phase-analysis were assessed using the Highscore program (PANalytical). The data were entered into the algorithm according to Rietveld [48] and mathematically processed to cover the measured diffractogram completely. The morphology of the aerogel particles (particle shape and size) and their chemical composition were evaluated on a field-emission scanning electron microscope Verios 460L (Thermo Fisher) equipped with an energy dispersion spectrometer (SEM-EDX). The specific surface of perovskites was measured on a CHEMBET 3000 device (Quantachrome) and, using the BET method, evaluated from the adsorption isotherms of nitrogen measured at 77 K on samples degassed at 300 °C. X-ray photoelectron spectroscopy (XPS) of aerogels reduced in the course of DRM was carried out using Kratos Axis Supra (Kratos-XPS), with monochromatic Ka radiation, 300 W emission power, and magnetic lens and the charge compensation turned on. The survey and detailed spectra were measured using pass energies of 160 and 20 eV, respectively. The spectra were evaluated using the Unifit 2013 software.
H2-TPR and O2-TPO profiles and simultaneous thermal analysis of aerogels were measured on STA 409 CD/QMS 403 Skimmer analyzers. In total, 20 mg of samples was placed in an Al2O3 crucible with a reduction or oxidation atmosphere flowing around it. Thermal analysis was measured in a hydrogen or oxygen atmosphere in a temperature range of 50 °C to 950 °C at a heating rate of 10 °C/min. The sample was cooled under an Ar atmosphere at a rate of 10 °C/min. A mixture of 5% H2 in Ar and 20% O2 in Ar was used, flowing through the thermal analyzer at a rate of 100 mL/min. The mass analysis of the gaseous phase was conducted using the Skimmer device in combination with the thermogravimetric analysis.

4.3. Tests of Catalytic Activity of Aerogel Catalysts

Catalytic activity testing of aerogel catalysts was performed on the high-temperature heterogeneous reactor Catlab (Hiden Analytical Ltd., Warrington, UK) from 30 °C to 800 °C at a heating rate of 10 °C/min. The gaseous phase was analyzed using the HPR-20 mass spectrometer (Hiden Analytical Ltd., UK). Temperature-programmed reduction by hydrogen (H2-TPR) and temperature-programmed oxidation by oxygen (O2-TPO) analyses of each sample were repeated three times. The mass spectrometer was calibrated with two calibration mixtures of gases (CO, CO2, H2, CH4 in Ar, and H2O in Ar, Messer, Bridgewater, NJ, USA).
The catalytic bed made up of aerogel particles of 100 mg in weight, deposited on alumina fibers (Saffil ceramic fibers, Unifrax Ltd., Widnes, UK), formed a piston of 40 mm in length. A temperature programmer controlled the furnace temperature in a temperature range of 30 °C to 800 °C. The reaction mixture of model biogas, CH4/CO2/Ar (3.8/3.2/53 mL/min), flowed through the tubular fixed-bed reactor at an overall rate of 60 mL/min (GHSV = 36 L/gcat.h). The catalytic tests were conducted under isothermal conditions in steps of 100 °C in the range from 400 °C to 1000 °C, with isothermal steps taking 30 min under isobaric conditions at a pressure of 1 bar. Long-time-on-the-stream tests of catalytical stability were conducted at 900 °C/15 h under the conditions given above. Prior to catalytic tests, the aerogels were reduced in Ar/5vol.%H2 atmosphere at 800 °C/1 h. Data from the mass spectrometer were used to assess the molar flows of CH4, CO2, H2, and CO. The conversion of methane X(CH4), carbon dioxide X(CO2), the yield of hydrogen Y(H2) and carbon monoxide Y(CO) were calculated according to Equations (5)–(8), where Fi is the molar flow rate of i-component at the entrance to the reactor or output from the reactor.
X C O 2 = F C O 2 ( i n ) F C O 2 ( o u t ) F C O 2 ( i n ) 100 %
X C H 4 = F C H 4 ( i n ) F C H 4 ( o u t ) F C H 4 ( i n ) 100 %
Y H 2 = F H 2 ( o u t ) 2 F C H 4 ( i n ) 100 %
Y C O = F C O ( o u t ) F C H 4 ( i n ) + F C O 2 ( i n ) 100 %

Author Contributions

J.C.—Conceptualization, Investigation, Funding acquisition, Methodology, Project administration, and Writing—original draft; S.T.—Investigation, Visualization, and Writing—original draft; V.B.—Investigation and Writing—review and editing; J.C.J.—Investigation, Writing—review and editing, and Writing—original draft; K.C.—Investigation and Writing—review and editing; M.T.—Investigation and Writing—review and editing; L.C.—Funding acquisition and Writing—review and editing. All authors have read and agreed to the published version of the manuscript.

Funding

The work was carried out in the frame of the COST Action CA18125 “Advanced Engineering and Research of aeroGels for Environment and Life Sciences” (AERoGELS) and funded by the European Commission. The authors acknowledge the support of the Ministry of Education, Youth and Sports of the Czech Republic, under grant no. LTC20019. We also acknowledge CzechNanoLab Research Infrastructure supported by MEYS CR (project LM2018110).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The original contributions presented in the study are included in the article, further inquiries can be directed to the corresponding author.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Singh, R.; Dhir, A.; Mohapatra, S.K.; Mahla, S.K. Dry reforming of methane using various catalysts in the process: Review. Biomass Convers. Bior 2020, 10, 567–587. [Google Scholar] [CrossRef]
  2. Aramouni, N.A.K.; Touma, J.G.; Tarboush, B.A.; Zeaiter, J.; Ahmad, M.N. Catalyst design for dry reforming of methane: Analysis review. Renew. Sustain. Energy Rev. 2018, 82, 2570–2585. [Google Scholar] [CrossRef]
  3. Guo, S.; Sun, Y.; Zhang, Y.; Zhang, C.; Li, Y.; Bai, J. Bimetallic Nickel-Cobalt catalysts and their application in dry reforming reaction of methane. Fuel 2024, 358, 130290. [Google Scholar] [CrossRef]
  4. Bitters, J.S.; He, T.N.; Nestler, E.; Senanayake, S.D.; Chen, J.G.G.; Zhang, C. Utilizing bimetallic catalysts to mitigate coke formation in dry reforming of methane. J. Energy Chem. 2022, 68, 124–142. [Google Scholar] [CrossRef]
  5. Al-Fatesh, A.S.; Patel, N.; Fakeeha, A.H.; Alotibi, M.F.; Alreshaidan, S.B.; Kumar, R. Reforming of methane: Effects of active metals, supports, and promoters. Catal. Rev. 2023, 1–99. [Google Scholar] [CrossRef]
  6. Torimoto, M.; Sekine, Y. Effects of alloying for steam or dry reforming of methane: A review of recent studies. Catal. Sci. Technol. 2022, 12, 3387–3411. [Google Scholar] [CrossRef]
  7. Maleki, H.; Hüsing, N. Current status, opportunities and challenges in catalytic and photocatalytic applications of aerogels: Environmental protection aspects. Appl. Catal. B 2018, 221, 530–555. [Google Scholar] [CrossRef]
  8. Brinker, C.J.; Schere, G.W. Sol-Gel Science: The Physics and Chemistry of Sol-Gel Processing; Academic Press, Inc.: San Diego, CA, USA, 1990. [Google Scholar]
  9. Livage, J. Sol–gel synthesis of heterogeneous catalysts from aqueous solutions. Catal. Today 1998, 41, 3–19. [Google Scholar] [CrossRef]
  10. Sakka, S.E. Handbook of Sol-Gel Science and Technology-Processing, Characterization and Applications; Kluwer Academic Publishers: New York, NY, USA, 2005; p. 1968. [Google Scholar]
  11. Baumann, T.F.; Gash, A.E.; Satcher, J.H., Jr. A Robust Approach to Inorganic Aerogels: The Use of Epoxides in Sol–Gel Synthesis. In Aerogels Handbook; Michel, A., Aegerter, N.L., Koebel, M.M., Eds.; Springer: New York, NY, USA, 2011. [Google Scholar]
  12. Wei, T.Y.; Chen, C.H.; Chien, H.C.; Lu, S.Y.; Hu, C.C. A Cost-Effective Supercapacitor Material of Ultrahigh Specific Capacitances: Spinel Nickel Cobaltite Aerogels from an Epoxide-Driven Sol-Gel Process. Adv. Mater. 2010, 22, 347–351. [Google Scholar] [CrossRef]
  13. House, J.E. Inorganic Chemistry; Elsevier: New York, NY, USA, 2008; p. 864. [Google Scholar]
  14. Shannon, R.D. Revised effective ionic radii and systematic studies of interatomic distances in halides and chalcogenides. Acta Crystallogr. Sect. A 1976, 32, 751–767. [Google Scholar] [CrossRef]
  15. Rayner-Canham, G.; Overton, T. Descriptive Inorganic Chemistry; W. H. Freeman and Company: New York, NY, USA, 2010. [Google Scholar]
  16. Danks, A.E.; Hall, S.R.; Schnepp, Z. The evolution of ‘sol-gel’ chemistry as a technique for materials synthesis. Mater. Horiz. 2016, 3, 91–112. [Google Scholar] [CrossRef]
  17. Zhang, M.; Guo, S.H.; Zheng, L.; Zhang, G.N.; Hao, Z.P.; Kang, L.P.; Liu, Z.H. Preparation of NiMn2O4 with large specific surface area from an epoxide-driven sol-gel process and its capacitance. Electrochim. Acta 2013, 87, 546–553. [Google Scholar] [CrossRef]
  18. Clapsaddle, B.J.N.B.; Wittstock, A.; Sprehn, D.W.; Gash, A.E.; Satcher, J.H.; Simpson, R.L.; Bäumer, M. A sol–gel methodology for the preparation of lanthanide-oxide aerogels: Preparation and characterization. J. Sol-Gel Sci. Technol. 2012, 64, 381–389. [Google Scholar] [CrossRef]
  19. Juhl, S.J.; Dunn, N.J.H.; Carroll, M.K.; Anderson, A.M.; Bruno, B.A.; Madero, J.E.; Bono, M.S. Epoxide-assisted alumina aerogels by rapid supercritical extraction. J. Non-Cryst. Solids 2015, 426, 141–149. [Google Scholar] [CrossRef]
  20. Sing, K.S.W. Reporting physisorption data for gas/solid systems with special reference to the determination of surface area and porosity (Recommendations 1984). Pure Appl. Chem. 1985, 57, 603–619. [Google Scholar] [CrossRef]
  21. Lago, R.; Bini, G.; Pena, M.A.; Fierro, J.L.G. Partial oxidation of methane to synthesis gas using LnCoO3 perovskites as catalyst precursors. J. Catal. 1997, 167, 198–209. [Google Scholar] [CrossRef]
  22. Osazuwa, O.U.; Cheng, C.K. Catalytic conversion of methane and carbon dioxide (greenhouse gases) into syngas over samarium-cobalt-trioxides perovskite catalyst. J. Clean Prod. 2017, 148, 202–211. [Google Scholar] [CrossRef]
  23. Osazuwa, O.U.; Setiabudi, H.D.; Abdullah, S.; Cheng, C.K. Syngas production from methane dry reforming over SmCoO3 perovskite catalyst: Kinetics and mechanistic studies. Int. J. Hydrogen Energy 2017, 42, 9707–9721. [Google Scholar] [CrossRef]
  24. Osazuwa, O.U.; Setiabudi, H.D.; Rasid, R.A.; Cheng, C.K. Syngas production via methane dry reforming: A novel application of SmCoO3 perovskite catalyst. J. Nat. Gas Sci. Eng. 2017, 37, 435–448. [Google Scholar] [CrossRef]
  25. Lima, S.M.; Assaf, J.M.; Pena, 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]
  26. Wu, L.A.; Jiang, X.C.; Wu, S.L.; Yao, R.; Qiao, X.S.; Fan, X.P. Synthesis of monolithic zirconia with macroporous bicontinuous structure via epoxide-driven sol-gel process accompanied by phase separation. J. Sol-Gel Sci. Technol. 2014, 69, 1–8. [Google Scholar] [CrossRef]
  27. Mousavi, M.; Pour, A.N. Performance and structural features of LaNi0.5Co0.5O3 perovskite oxides for the dry reforming of methane: Influence of the preparation method. New J. Chem. 2019, 43, 10763–10773. [Google Scholar] [CrossRef]
  28. Zhao, X.Y.; Cao, Y.; Li, H.R.; Zhang, J.P.; Shi, L.Y.; Zhang, D.S. Sc promoted and aerogel confined Ni catalysts for coking-resistant dry reforming of methane. Rsc Adv. 2017, 7, 4735–4745. [Google Scholar] [CrossRef]
  29. Wang, C.Z.; Sun, N.N.; Zhao, N.; Wei, W.; Zhao, Y.X. Template-free preparation of bimetallic mesoporous Ni-Co-CaO-ZrO2 catalysts and their synergetic effect in dry reforming of methane. Catal. Today 2017, 281, 268–275. [Google Scholar] [CrossRef]
  30. Hao, Z.G.; Zhu, Q.S.; Jiang, Z.; Li, H.Z. Fluidization characteristics of aerogel Co/Al2O3 catalyst in a magnetic fluidized bed and its application to CH4-CO2 reforming. Powder Technol. 2008, 183, 46–52. [Google Scholar] [CrossRef]
  31. Hao, Z.G.; Zhu, Q.S.; Lei, Z.; Li, H.Z. CH4-CO2 reforming over Ni/Al2O3 aerogel catalysts in a fluidized bed reactor. Powder Technol. 2008, 182, 474–479. [Google Scholar] [CrossRef]
  32. Li, P.C.; Li, J.; Zhu, Q.S.; Cui, L.J.; Li, H.Z. Effect of granulation on the activity and stability of a Co-Al2O3 aerogel catalyst in a fluidized-bed reactor for CH4-CO2 reforming. Rsc Adv. 2013, 3, 8939–8946. [Google Scholar] [CrossRef]
  33. Cao, A.N.T.; Pham, C.Q.; Pham, L.H.; Nguyen, D.L.; Phuong, P.T.T.; Tran, T.T.V.; Nguyen, V.; Nguyen, T.B.; Van Le, Q.; Nguyen, N.A.; et al. Boosted methane dry reforming for hydrogen generation on cobalt catalyst with small cerium dosage. Int. J. Hydrogen Energy 2022, 47, 42200–42212. [Google Scholar] [CrossRef]
  34. Cao, A.N.T.; Pham, C.Q.; Nguyen, T.M.; Van Tran, T.; Phuong, P.T.T.; Vo, D.V.N. Dysprosium promotion on Co/Al2O3 catalysts towards enhanced hydrogen generation from methane dry reforming. Fuel 2022, 324, 124818. [Google Scholar] [CrossRef]
  35. Sheng, K.; Luan, D.; Jiang, H.; Zeng, F.; Wei, B.; Pang, F.; Ge, J. NixCoy Nanocatalyst Supported by ZrO2 Hollow Sphere for Dry Reforming of Methane: Synergetic Catalysis by Ni and Co in Alloy. Acs Appl. Mater. Inter. 2019, 11, 24078–24087. [Google Scholar] [CrossRef]
  36. Yang, T.Z.; Chen, W.; Chen, L.; Liu, W.F.; Zhang, D.C. Promotion effect between Ni and Co aerogel catalysts in CH4 reforming with CO2 and O2. J. CO2 Util. 2016, 16, 130–137. [Google Scholar] [CrossRef]
  37. Tran, N.T.; Le, Q.V.; Cuong, N.V.; Nguyen, T.D.; Phuc, N.H.H.; Phuong, P.T.T.; Monir, M.U.; Abd Aziz, A.; Truong, Q.D.; Abidin, S.Z.; et al. La-doped cobalt supported on mesoporous alumina catalysts for improved methane dry reforming and coke mitigation. J. Energy Inst. 2020, 93, 1571–1580. [Google Scholar] [CrossRef]
  38. Bahari, M.B.; Phuc, N.H.H.; Alenazey, F.; Vu, K.B.; Ainirazali, N.; Vo, D.-V.N. Catalytic performance of La-Ni/Al2O3 catalyst for CO2 reforming of ethanol. Catal. Today 2017, 291, 67–75. [Google Scholar] [CrossRef]
  39. Li, K.; Chang, X.; Pei, C.; Li, X.; Chen, S.; Zhang, X.; Assabumrungrat, S.; Zhao, Z.-J.; Zeng, L.; Gong, J. Ordered mesoporous Ni/La2O3 catalysts with interfacial synergism towards CO2 activation in dry reforming of methane. Appl. Catal. B 2019, 259, 118092. [Google Scholar] [CrossRef]
  40. Bian, Z.F.; Kawi, S. Highly carbon-resistant Ni-Co/SiO2 catalysts derived from phyllosilicates for dry reforming of methane. J. CO2 Util. 2017, 18, 345–352. [Google Scholar] [CrossRef]
  41. Gupta, S.; Fernandes, R.; Patel, R.; Spreitzer, M.; Patel, N. A review of cobalt-based catalysts for sustainable energy and environmental applications. Appl. Catal. A 2023, 661, 119254. [Google Scholar] [CrossRef]
  42. San-Jose-Alonso, D.; Juan-Juan, J.; Illan-Gomez, M.J.; Roman-Martinez, M.C. Ni, Co and bimetallic Ni-Co catalysts for the dry reforming of methane. Appl. Catal. A-Gen. 2009, 371, 54–59. [Google Scholar] [CrossRef]
  43. You, X.J.; Wang, X.; Ma, Y.H.; Liu, J.J.; Liu, W.M.; Xu, X.L.; Peng, H.G.; Li, C.Q.; Zhou, W.F.; Yuan, P.; et al. Ni-Co/Al2O3 Bimetallic Catalysts for CH4 Steam Reforming: Elucidating the Role of Co for Improving Coke Resistance. Chemcatchem 2014, 6, 3377–3386. [Google Scholar] [CrossRef]
  44. Akhavan, O.; Ghaderi, E. Photocatalytic Reduction of Graphene Oxide Nanosheets on TiO2 Thin Film for Photoinactivation of Bacteria in Solar Light Irradiation. J. Phys. Chem. C 2009, 113, 20214–20220. [Google Scholar] [CrossRef]
  45. Yumitori, S. Correlation of C1s chemical state intensities with the O1s intensity in the XPS analysis of anodically oxidized glass-like carbon samples. J. Mater. Sci. 2000, 35, 139–146. [Google Scholar] [CrossRef]
  46. Zhang, Z.; Verykios, X.E.; MacDonald, S.M.; Affrossman, S. Comparative Study of Carbon Dioxide Reforming of Methane to Synthesis Gas over Ni/La2O3 and Conventional Nickel-Based Catalysts. J. Phys. Chem. 1996, 100, 744–754. [Google Scholar] [CrossRef]
  47. Yang, D.; Velamakanni, A.; Bozoklu, G.; Park, S.; Stoller, M.; Piner, R.; Stankovich, S.; Jung, I.; Field, D.; Ventrice, J.C.; et al. Chemical Analysis of Graphene Oxide Films after Heat and Chemical Treatments by X-ray Photoelectron and Micro-Raman Spectroscopy. Carbon 2009, 47, 145–152. [Google Scholar] [CrossRef]
  48. Rietveld, H. A profile refinement method for nuclear and magnetic structures. J. Appl. Crystallogr. 1969, 2, 65–71. [Google Scholar] [CrossRef]
Figure 1. Images of supercritically dried aerogels: (a) Sm-Co-O, (b) Sm-Ni-O, (c) Al-Co-O, (d) Al-Ni-O, and (e) Al-Co-Ni-O.
Figure 1. Images of supercritically dried aerogels: (a) Sm-Co-O, (b) Sm-Ni-O, (c) Al-Co-O, (d) Al-Ni-O, and (e) Al-Co-Ni-O.
Gels 10 00328 g001
Figure 2. Simultaneous TGA/MS profiles of supercritically dried aerogels: (a) Sm-Co-O, (b) Sm-Ni-O, (c) Al-Co-O, (d) Al-Ni-O, and (e) Al-Co-Ni-O.
Figure 2. Simultaneous TGA/MS profiles of supercritically dried aerogels: (a) Sm-Co-O, (b) Sm-Ni-O, (c) Al-Co-O, (d) Al-Ni-O, and (e) Al-Co-Ni-O.
Gels 10 00328 g002
Figure 3. XRD patterns of calcined (a) and reduced (b) aerogels.
Figure 3. XRD patterns of calcined (a) and reduced (b) aerogels.
Gels 10 00328 g003
Figure 4. Nitrogen adsorption and desorption isotherms of thermally treated (a) and calcined (b) aerogels.
Figure 4. Nitrogen adsorption and desorption isotherms of thermally treated (a) and calcined (b) aerogels.
Gels 10 00328 g004
Figure 5. TGA and H2-TPR profiles of calcined aerogels: (a) Sm-Co-O, (b) Sm-Ni-O, (c) Al-Co-O, (d) Al-Ni-O, and (e) Al-Co-Ni-O.
Figure 5. TGA and H2-TPR profiles of calcined aerogels: (a) Sm-Co-O, (b) Sm-Ni-O, (c) Al-Co-O, (d) Al-Ni-O, and (e) Al-Co-Ni-O.
Gels 10 00328 g005
Figure 6. SEM image and EDX elemental mapping of the reduced aerogels: (a) Sm-Co-O, (b) Sm-Ni-O, (c) Al-Co-O, (d) Al-Ni-O, and (e) Al-Co-Ni-O.
Figure 6. SEM image and EDX elemental mapping of the reduced aerogels: (a) Sm-Co-O, (b) Sm-Ni-O, (c) Al-Co-O, (d) Al-Ni-O, and (e) Al-Co-Ni-O.
Gels 10 00328 g006
Figure 7. Catalytic performance of reduced aerogel catalysts in DRM test as a function of reaction temperature: (a) Sm-Co-O, (b) Sm-Ni-O, (c) Al-Co-O, (d) Al-Ni-O, and (e) Al-Co-Ni-O.
Figure 7. Catalytic performance of reduced aerogel catalysts in DRM test as a function of reaction temperature: (a) Sm-Co-O, (b) Sm-Ni-O, (c) Al-Co-O, (d) Al-Ni-O, and (e) Al-Co-Ni-O.
Gels 10 00328 g007
Figure 8. CH4 and CO2 conversion and H2 and CO yield with time on stream in the dry reformation of methane over aerogel catalysts at 800, 750, and 700 °C: (a) Sm-Co-O, (b) Sm-Ni-O, (c) Al-Co-O, (d) Al-Ni-O, and (e) Al-Co-Ni-O.
Figure 8. CH4 and CO2 conversion and H2 and CO yield with time on stream in the dry reformation of methane over aerogel catalysts at 800, 750, and 700 °C: (a) Sm-Co-O, (b) Sm-Ni-O, (c) Al-Co-O, (d) Al-Ni-O, and (e) Al-Co-Ni-O.
Gels 10 00328 g008
Figure 9. TPO profiles of spent aerogel catalysts (after 14 h of DRM test at 700 to 800 °C).
Figure 9. TPO profiles of spent aerogel catalysts (after 14 h of DRM test at 700 to 800 °C).
Gels 10 00328 g009
Figure 10. Al 2p XPS spectra of reduced Sm-Al-Co-Ni-O aerogels after DRM tests (a) Al-Co-O and (b) Al-Ni-O.
Figure 10. Al 2p XPS spectra of reduced Sm-Al-Co-Ni-O aerogels after DRM tests (a) Al-Co-O and (b) Al-Ni-O.
Gels 10 00328 g010
Figure 11. C 1s XPS spectra of aerogels after DRM tests: (a) Sm-Co-O, (b) Sm-Ni-O, (c) Al-Co-O, (d) Al-Ni-O, and (e) Al-Co-Ni-O.
Figure 11. C 1s XPS spectra of aerogels after DRM tests: (a) Sm-Co-O, (b) Sm-Ni-O, (c) Al-Co-O, (d) Al-Ni-O, and (e) Al-Co-Ni-O.
Gels 10 00328 g011
Table 1. Summary of synthesis conditions in the preparation of Sm-Co-O, Sm-Ni-O, Al-Co-O, Al-Ni-O, and Al-Co-Ni-O gels.
Table 1. Summary of synthesis conditions in the preparation of Sm-Co-O, Sm-Ni-O, Al-Co-O, Al-Ni-O, and Al-Co-Ni-O gels.
Synthesis PrecursorsConcentration (mol)H2O/Mi (mol/mol)tgel (min)Charge Density of Cations (C/mm3)
CoCl2·6H2O0.0144625155
SmCl3·6H2O0.014486
propylene oxide0.2858
ethanol1.182
NiCl2·6H2O0.014467134
SmCl3·6H2O0.014486
propylene oxide0.2858
ethanol1.182
CoCl2·6H2O0.014463155
AlCl3·6H2O0.0144364
propylene oxide0.2858
ethanol1.182
NiCl2·6H2O0.014463134
AlCl3·6H2O0.0144364
propylene oxide0.2858
ethanol1.182
NiCl2·6H2O0.007264134
CoCl2·6H2O0.0072155
AlCl3·6H2O0.0144364
propylene oxide0.2858
ethanol1.182
Table 2. Simultaneous TGA/MS of supercritically dried aerogels.
Table 2. Simultaneous TGA/MS of supercritically dried aerogels.
Sample/Peak(Dm)TG(Q)DSC(Tmax)DTG(Tmax)DSC(Tmax)H2O(Tmax)CO2(Tmax)HCl
(wt%)(J/g)(°C)
SmCoO/1−6−122134120123
SmCoO/2−17642287294288288254
SmNiO/1−9−241144135138134
SmNiO/2−7588259260259259226
SmNiO/3−16324324317317
AlCoO/1−16−202134134144
AlCoO/2−10363287310286301225
AlNiO/1−12−162134125140122
AlNiO/2−1577291305273293
AlNiO/3−580615376 359
AlNiCoO/1 −254136136144136
AlNiCoO/2 242289337289330
Table 3. The specific surface of thermally treated and calcined aerogels.
Table 3. The specific surface of thermally treated and calcined aerogels.
SampleSpecific Surface (m2/g)
Thermally TreatedCalcined
Sm-Co-O3251.51
Sm-Ni-O47018.9
Al-Co-O45911.9
Al-Ni-O39568.7
Al-Co-Ni-O34634.4
Table 4. Elemental composition of the reduced aerogels obtained from scanning electron microscopy-energy dispersive X-ray spectroscopy (SEM-EDX) analysis.
Table 4. Elemental composition of the reduced aerogels obtained from scanning electron microscopy-energy dispersive X-ray spectroscopy (SEM-EDX) analysis.
SampleAt. %
SmAlCoNiO
Sm-Co-O27.70.025.00.047.4
Sm-Ni-O29.00.024.50.046.5
Al-Co-O0.028.623.40.048.0
Al-Ni-O0.027.90.024.847.4
Al-Co-Ni-O0.024.29.325.341.2
Table 5. Catalytic performance of Sm-Al-Co-N i-O catalysts in dry reformation of methane at 600 to 800 °C.
Table 5. Catalytic performance of Sm-Al-Co-N i-O catalysts in dry reformation of methane at 600 to 800 °C.
SampleConversion (%)Temperature (°C)
Yield (%)600650700750800
Sm-Co-OX(CH4)51.1 ± 0.3 81.2 ± 0.391.8 ± 0.498.0 ± 0.499.6 ± 0.2
X(CO2)60.2 ± 0.280.4 ± 0.290.2 ± 0.392.5 ±0.693.7 ± 0.5
Y(H2)46.6 ± 0.277.3 ± 0.190.6 ± 0.396.4 ± 0.495.8 ± 0.3
Y(CO)36.4 ± 0.169.5 ± 0.387.4 ± 0.493.8 ± 0.395.7 ± 0.6
Sm-Ni-OX(CH4)85.2 ± 0.193.8 ± 0.197.6 ± 0.199.3 ± 0.299.8 ± 0.1
X(CO2)80.6 ± 0.487.6 ± 0.391.4 ± 0.292.5 ± 0.394.1 ± 0.3
Y(H2)88.4 ± 0.288.7 ± 0.293.2 ± 0.293.7 ± 0.294.6 ± 0.4
Y(CO)76.3 ± 0.180.1 ± 0.292.7 ± 0.392.6 ± 0.296.3 ± 0.1
Al-Co-OX(CH4)82.2 ± 0.192.6 ± 0.597.6 ± 0.299.3 ± 0.499.7 ± 0.2
X(CO2)76.4 ± 0.284.0 ± 0.389.2 ± 0.191.2 ± 0.292.3 ± 0.5
Y(H2)76.6 ± 0.288.6 ± 0.194.3 ± 0.294.7 ± 0.295.4 ± 0.1
Y(CO)70.4 ± 0.185.2 ± 0.692.4 ± 0.194.6 ± 0.295.2 ± 0.2
Al-Ni-OX(CH4)85.1 ± 0.393.7 ± 0.298.4 ± 0.199.2 ± 0.299.9 ± 0.1
X(CO2)80.6 ± 0.188.4 ± 0.291.3 ± 0.493.1 ± 0.294.6 ± 0.2
Y(H2)80.5 ± 0.491.0 ± 0.394.7 ± 0.195.6 ± 0.596.4 ± 0.2
Y(CO)74.6 ± 0.287.4 ± 0.293.3 ± 0.294.5 ± 0.395.7 ± 0.2
Al-Co-Ni-OX(CH4)84.3 ± 0.284.3 ± 0.197.5 ± 0.299.4 ± 0.299.7 ± 0.3
X(CO2)79.4 ± 0.286.5 ± 0.289.6 ± 0.291.3 ± 0.492.3 ± 0.2
Y(H2)81.1 ± 0.192.3 ± 0.495.3 ± 0.396.4 ± 0.395.7 ± 0.2
Y(CO)73.1 ± 0.386.2 ± 0.292.1 ± 0.294.0 ± 0.695.3 ± 0.1
Table 6. Carbon deposition after 14 h of methane dry reformation catalyzed by aerogel catalysts.
Table 6. Carbon deposition after 14 h of methane dry reformation catalyzed by aerogel catalysts.
Sample/PeakTCO2 (°C)ICCO2 (A·s·109)C-Deposit
(mg C/gcat.h)
SC-Deposit
MeanSdevMeanSdevMeanSdevMeanSdev
Sm-Co-O/127533.80.60.0280.0110.0420.021
Sm-Co-O/253261.90.30.0140.01
Sm-Ni-O/13432235.35.20.260.0050.290.008
Sm-Ni-O/249654.30.60.030.004
Al-Co-O/148416607.789.44.470.434.470.43
Al-Ni-O/15383282.741.62.080.0232.930.077
Al-Ni-O/26003115.317.00.850.074
Al-Co-Ni-O/1335514.42.10.110.0050.350.011
Al-Co-Ni-O/2483433.14.90.240.006
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

Cihlar, J.; Tkachenko, S.; Bednarikova, V.; Cihlar, J., Jr.; Castkova, K.; Trunec, M.; Celko, L. Study of the Synthesis of Multi-Cationic Sm-Co-O, Sm-Ni-O, Al-Co-O, Al-Ni-O, and Al-Co-Ni-O Aerogels and Their Catalytic Activity in the Dry Reforming of Methane. Gels 2024, 10, 328. https://doi.org/10.3390/gels10050328

AMA Style

Cihlar J, Tkachenko S, Bednarikova V, Cihlar J Jr., Castkova K, Trunec M, Celko L. Study of the Synthesis of Multi-Cationic Sm-Co-O, Sm-Ni-O, Al-Co-O, Al-Ni-O, and Al-Co-Ni-O Aerogels and Their Catalytic Activity in the Dry Reforming of Methane. Gels. 2024; 10(5):328. https://doi.org/10.3390/gels10050328

Chicago/Turabian Style

Cihlar, Jaroslav, Serhii Tkachenko, Vendula Bednarikova, Jaroslav Cihlar, Jr., Klara Castkova, Martin Trunec, and Ladislav Celko. 2024. "Study of the Synthesis of Multi-Cationic Sm-Co-O, Sm-Ni-O, Al-Co-O, Al-Ni-O, and Al-Co-Ni-O Aerogels and Their Catalytic Activity in the Dry Reforming of Methane" Gels 10, no. 5: 328. https://doi.org/10.3390/gels10050328

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