Next Article in Journal
The Behavior of Divalent Metals in Double-Layered Hydroxides as a Fenton Bimetallic Catalyst for Dye Decoloration: Kinetics and Experimental Design
Previous Article in Journal
FeCo-LDH/CF Cathode-Based Electrocatalysts Applied to a Flow-Through Electro-Fenton System: Iron Cycling and Radical Transformation
Previous Article in Special Issue
Supported TiO2 Photocatalysis of Spiked Contaminants in Water and Municipal Wastewater
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

The Development of Ni-Al Aerogel-Based Catalysts via Supercritical CO2 Drying for Photocatalytic CO2 Methanation

by
Daniel Estevez
,
Haritz Etxeberria
and
Victoria Laura Barrio
*
School of Engineering of Bilbao, University of the Basque Country (UPV/EHU), Plaza Ingeniero Torres Quevedo 1, 48013 Bilbao, Spain
*
Author to whom correspondence should be addressed.
Catalysts 2025, 15(7), 686; https://doi.org/10.3390/catal15070686
Submission received: 30 May 2025 / Revised: 14 July 2025 / Accepted: 15 July 2025 / Published: 16 July 2025
(This article belongs to the Special Issue Advancements in Photocatalysis for Environmental Applications)

Abstract

The conversion of CO2 into CH4 through the Sabatier reaction is one of the key processes that can reduce CO2 emissions into the atmosphere. This work aims to develop Ni-Al aerogel-based thermo-photocatalysts with large specific surface areas prepared using a sol–gel method and subsequent supercritical drying in CO2. Different Al/Ni molar ratios were selected for the development of the catalysts, characterized using ICP-OES, N2 adsorption–desorption isotherms, XRD, H2-TPR, TEM, UV-Vis DRS, and XPS techniques. Thermo-photocatalytic activity tests were performed in a photoreactor with two different light sources (λ = 365 nm, λ = 470 nm) at a temperature range from 300 °C to 450 °C and a pressure of 10 bar. The catalyst with the highest Ni loading (AG 1/3) produced the best catalytic results, reaching CO2 conversion and CH4 selectivity levels of 82% and 100%, respectively, under visible light at 450 °C. In contrast, the catalysts with the lowest nickel loading produced the lowest results, most likely due to their low amounts of active Ni. These results suggest that supercritical drying is an efficient method for developing active thermo-photocatalysts with high Ni dispersion, suitable for Sabatier reactions under mild reaction conditions.

1. Introduction

Global CO2 emissions reached a new high in 2023, increasing by 0.1% compared to the previous year [1]. Different carbon capture and utilization (CCU) technologies have been developed with the objective of preventing further growth in emissions and the greenhouse effect. Power-to-X technologies are useful strategies for converting excess electric energy into chemical molecules, such as methane [2], in power-to-gas processes. Moreover, if the hydrogen utilized in the reaction has a renewable origin, the process will be sustainable. Methane is a widely used gas; it is a precursor for a variety of reactions, such as hydrocarbon synthesis [3] (through methane reforming [4]) and halide synthesis [5] and it is used as a rocket propellant [6].
One of the systems included in power-to-gas technologies is the CO2 methanation reaction. This reaction was first studied by Paul Sabatier in 1902, where methane was generated due to contact between CO2 and H2 at high temperatures and in the presence of a catalyst [7]. The Sabatier reaction is an exothermic reaction (1), in which CH4 and H2O are the obtained products alongside CO, which can appear as a byproduct.
C O 2 + 4 H 2 C H 4 + 2 H 2 O H = 165.0 K J · m o l 1
To achieve a sustainable process, temperatures should be minimized to the lowest value at which methane can still be obtained. To this end, different catalysts have been used. For example, noble metal-based catalysts have been developed, utilizing metals such as Ru [8,9,10], Rh [11,12,13,14], and Pd [15,16,17]. Ru shows the highest activity and stability [18]; it achieves methane conversion rates close to 90% at 375 °C for 100 h [19], followed by Rh and Pd. Considering the high cost and scarcity of these metals, cheaper alternative solutions have been found. One solution is to use non-noble metals, such as Ni [20,21,22,23,24,25], Co [26,27,28,29], and Fe [30,31,32], which have been used as catalysts. Nickel is the most widely used due to its low price and high availability. The activity of Ni-based catalysts can vary depending on several conditions, e.g., catalytic support, active metal dispersion, particle size, and the addition of other metals acting as dopants. In some studies, CO2 conversions of 92% were obtained at 350 °C [33]. Nevertheless, even if these materials present high methanation activity, they deactivate easily via coke deposition, which blocks the active sites of the catalyst. In addition, the Ni particles can suffer from sintering at high temperatures, decreasing the catalyst’s activity. Solving this issue requires searching for adequate support, adding other metals to form a co-catalyst, and identifying the proper reaction conditions.
Moreover, the support of each catalyst determines different textural properties, including surface area and pore volume, the acidity or basicity of the catalyst, and the presence of higher or lower active metal dispersion, which yields better conversion and selectivity results. Traditionally, different metal oxides have been used as catalytic supports, and depending on which one is applied, the methanation reaction will follow one reaction path or another. For example, if Al2O3 or MgO is used as the support, the reaction will follow a bicarbonate formation route; however, if TiO2 or ZrO2 is used as the support, the reaction will occur through the formation of bidentate carbonate species [34].
As stated previously, one of the most important properties when synthesizing a Ni-based catalyst is the surface area; a larger surface area increases the active metal dispersion, and thus, sintering is avoided. In response to the need for higher surface areas, new supports were developed, such as hydrotalcites, zeolites, and aerogels. Hydrotalcites are double-layered hydroxides, the properties of which vary depending on the hydroxides of the compound and the ratios between them. In some cases, hydrotalcite-based catalysts with surface areas as high as 240 m2/g were synthesized [35] in comparison to the 125 m2/g of traditional Ni catalysts [36]. Zeolites are crystalline microporous materials; their activity varies depending on the properties of the materials, which are determined via the synthesis method. CO2 conversions as high as 70 to 80% were obtained at 400 °C with different zeolite-supported Ni-based catalysts [37,38,39]. In the case of aerogels, these materials are gelified and supercritically dried so that the original gel structure is preserved; thus, high porosities and surface areas are obtained. While their properties have been widely studied, their potential as catalysts for the CO2 methanation reaction has not been investigated. Even so, aerogels with surface areas above 300 m2/g have been developed [40], making them adequate candidates for the dispersal of active metals such as nickel. This was tested in a case where a Ni-based graphene aerogel was synthesized, rendering 80% CO2 conversion at 350 °C in comparison to its graphene oxide counterpart, which reached 68% conversion at the same temperature [41].
Given the high CO2 emissions and their consequences, global warming, there is a need to develop CO2 methanation technology for industrial scale-up, reducing the rate of human-generated CO2 emissions. For this purpose, catalysts with appropriate physical, chemical, and structural properties (such as high specific surface area or high metal dispersion) offering high methanation activity must be studied. Economic and abundant metals like nickel, cobalt, or iron, which are highly active as catalysts for the Sabatier reaction, need to be investigated for this process to become economically affordable. It is also necessary to develop and improve existing catalytic systems with previously validated high CO2 conversion and selectivity towards CH4, so that the whole technique can be optimized, either via the search for new catalytic formulations with different metals and supports or using innovative synthesis methods.
In this project, a series of Ni and Al aerogel-based catalysts were prepared. The Al/Ni molar ratio was varied via a sol–gel reaction, followed by subsequent supercritical CO2 drying. The structural, compositional and physicochemical properties were studied using N2 adsorption–desorption isotherms, inductively coupled plasma optical emission spectroscopy (ICP-OES), hydrogen temperature-programmed reduction (H2-TPR), powder X-ray diffraction (XRD), X-ray photoelectron spectroscopy (XPS), ultraviolet–visible diffuse reflectance spectroscopy (UV-Vis DRS), and transmission electron microscopy (TEM). The thermo-photocatalytic activity of CO2 methanation was determined in a bench-scale plant with an 8 mm open, removable quartz window, and an LED light source was attached. The reactions were performed at a temperature range of 300 to 450 °C, at a constant pressure of 10 bar, with no light, a 470 nm light source, or a 365 nm light source, and a catalytic mass of 5 mg.

2. Results and Discussion

2.1. Textural Properties

The Al/Ni molar ratio obtained using ICP-OES is displayed in Table 1. The molar fractions of each catalyst were similar to the theoretical values. Regarding the N2 adsorption and desorption results, the calcined catalysts exhibited large specific surface areas. These values increased with the aluminum content, as expected, owing to alumina’s large surface area. The total pore volume and pore diameter increased following the same trend with higher aluminum loading. Notably, the pore diameter values were larger than those typically observed in other large surface area materials, such as hydrotalcites [42].
The formation of large pores in the Al/Ni aerogels can be attributed to the synthesis method used, which maintains the gel’s structure throughout the supercritical drying process. In all cases, the shape of the isotherms corresponded to Type IV isotherms, which is characteristic of mesoporous materials. These are shown in Figure S1 of the Supplementary Material.
Table 1 shows the complete results of the prepared catalysts, where SBET, Vt, pore, and DBJH refer to the surface area calculated using the Brunauer–Emmett–Teller (BET) method, the total pore volume, and the pore diameter obtained using the Barrett–Joyner–Halenda (BJH) model, respectively.

2.2. X-Ray Diffraction

The X-ray diffraction results shown in Figure 1 provide insight into the crystalline structure of the catalysts. As highlighted, the different peaks present in all samples correspond to NiO (PDF 00-001-1239), along with a minor presence of NiAl2O4 spinel (PDF 00-001-1299). No peaks were detected for Al2O3; thus, the support was amorphous. Moreover, the sample with the highest Ni content (AG 1/3) exhibited the highest degree of crystallinity.

2.3. H2 Temperature-Programmed Reduction (H2-TPR)

The H2-TPR profiles of the calcined AGs are shown in Figure 2. In every catalyst, similar behavior was observed—one peak with three different observable regions. The first section, up to 530 °C, was attributed to the interaction between the NiO and the alumina [43,44]. The intermediate peak (from 530 to approximately 640 °C) corresponded to the same species with stronger interactions. Finally, the H2 consumption from 640 to 680 °C was ascribed to the reduction in the NiAl2O4 spinel [45,46], which became apparent in the calcination step of the synthesis. The presence of these species was confirmed by the XRD results. Figure 2 shows how AG 2/1, AG 3/1, and AG 4/1 presented a lower amount of NiO than the rest of the catalysts. In the case of AG 1/2 and AG1/3 with higher nickel contents, Ni was more extensively reduced, and the formation of aluminates was observed.

2.4. Transmission Electron Microscopy (TEM)

Transmission electron microscopy (TEM) measurements revealed that the nanoparticles in all the samples presented an oval morphology, except for AG 1/3, where hexagonal-shaped nanoparticles were observed. This was supported by the XRD results, where the crystallinity of the AG 1/3 sample was superior. Particle size distributions were relatively uniform for each sample, with diameters ranging between 2.0 and 5.0 nm in most cases. However, a shift to larger particle sizes was noted in samples with a higher nickel content from 3.0 to 6.0 nm for AG 1/2 and 3.0 to 8.0 nm for AG 1/3. To evaluate the effect of thermal treatment on the morphology of the materials, the non-calcined sample AG 1/1 (named AG 1/1 V) was also examined. Notably, the virgin aerogel presented larger Ni particles than its calcined counterpart, indicating that the calcination process enhanced the dispersion of the active metal phase. Moreover, EDX mapping confirmed the homogeneous distribution of both Ni and Al samples, with no significant evidence of particle agglomeration. A comparison of the morphological differences for the prepared catalysts is shown in Figure 3, while the TEM micrographs of the remaining catalysts are provided in Figure S2 of the Supplementary Material.

2.5. UV-Visible Diffuse Reflectance Spectroscopy (UV-Vis DRS)

The bands displayed in the UV-visible diffuse reflectance spectroscopy (UV-vis DRS) spectra (Figure 4) correspond to the different Ni-containing species present in the catalytic samples. The absorption bands in the 250–300 nm region are attributed to NiO species. These bands increase in intensity with metal loading and shift to longer wavelengths (red shift) based on the strong interaction between the Ni2+ and the metal support. The bands appearing at these wavelengths have been previously related to the transition of the O2− (2p)→Ni2+ (3d) charge transfer, which is characteristic of NiO species [47]. The second band, located at the 600–800 nm region, can be attributed to the contributions of the NiAl2O4 spinel and NiO species. Catalysts with lower Ni contents (AG 3/1 and AG 4/1) showed a higher absorbance and a flatter shape, which explains the small contribution of the d→d transitions from the tetrahedral Ni2+ of the spinel species [48] (as seen in the XRD results). On the other hand, catalysts with higher metal contents (such as AG 1/3 or AG 1/2) showed a profile similar in shape to the typical d→d transitions of Ni2+ in the NiO [49].
Additionally, the direct band gap value was calculated by using the Tauc equation, and by plotting (αhν)2 vs. (hν). The Tauc plot of every catalyst can be found in Figure S3 of the Supplementary Material, while the resulting band gap energies are shown in Table 2. As can be observed, the band gap energy lowers as the Ni content increases, which could be attributed to the interaction between the Ni 3d and O 2p states in the valence and conduction bands [50].
To summarize, UV-vis DRS was used to confirm the presence of two different species in our samples: NiO, as the principal species in samples with higher metal contents (AG 1/3 and AG 1/2), and NiAl2O4, which appeared in samples with lower loadings (AG 3/1 and AG 4/1) along with the NiO signal.

2.6. X-Ray Photoelectron Microscopy (XPS)

The elemental composition of the surface and oxidation state of the metals were studied via X-ray photoelectron spectroscopy (XPS) and are shown in Table 3. In all cases, the Al/Ni atomic ratios obtained using XPS were higher than the ratios obtained using ICP-OES, indicating lower nickel loading on the surface than expected, probably due to the defective distribution of the metals during the synthesis or the difference in the general content of each metal, as shown by ICP-OES. In the case of AG-Pt, 0.01% of Pt was detected with the Pt 4d5 signal, which translates into an Al/Pt atomic ratio of 2068.
The Ni 2p XPS spectra of every aerogel catalyst are presented in Figure 5. In all cases, three different peaks were observed for three different species: the first one at 852.8 eV corresponds to metallic Ni on the surface of the catalysts; the second peak at 858.2 eV appears due to the Ni2+ coming from the NiO; and the third peak corresponds to its satellite at 864.2 eV [42]. This behavior is apparent in both the Ni 2p 3/2 and Ni 2p 1/2 zones. In this regard, no significant differences were spotted when changing the Al/Ni ratio.
In the case of the O 1s spectra, only one peak was found at 528.2 eV, corresponding to the O2− present in the NiO of the catalysts. Moreover, the Al 2p peak of the AG 1/1 sample appeared at 73.8 eV; thus, it was assigned to the Al2O3 species in the sample [51]. The position of this peak appeared at a lower binding energy when compared to the rest of the catalysts (e.g., 71.5 eV for AG 4/1 and 70.6 eV for AG 1/3). No peaks were observed for Pt in the AG-Pt sample, probably due to the low loading present in the sample.

2.7. CO2 Methanation Activity Performance

The performance results of the different materials, in terms of CO2 conversion, are shown in Figure 6.
For the case under dark conditions, CO2 conversions of 65–75% were achieved at high temperatures (>400 °C), with the catalytic systems containing higher Ni loadings (AG 1/1, AG 1/2, and AG 1/3). In contrast, the aerogels with a higher aluminum content (AG 2/1 and AG 3/1) only reached 30% conversion at the same temperature. Finally, the catalyst with the highest amount of alumina (AG 4/1) did not produce any methane, probably due to its low content of active nickel, as measured using the H2-TPR results. Regarding methane selectivity, all catalysts except AG 4/1 exhibited values close to 100%, while AG 4/1 showed a reduced selectivity of 75% at the highest temperature.
When visible light was used as an activity enhancer, CO2 conversions were generally higher in comparison to those achieved under dark conditions. Additionally, almost all the catalysts presented activity starting at around 400 °C, in contrast to the results achieved under dark conditions. In particular, AG 1/1 and AG-Pt showed catalytic activity starting at 300 and 350 °C, respectively, highlighting their potential to reduce the energy demand of methanation, despite AG-Pt displaying slightly lower overall conversion levels (besides AG 1/3, which showed catalytic activity as low as 250 °C). Furthermore, under visible light conditions, the selectivity towards CH4 improved with respect to dark conditions, reaching values close to 100% at the highest temperatures in all cases.
Using ultraviolet light, the results were intermediate between those obtained under dark and visible light conditions in terms of both CO2 conversion and CH4 selectivity. Regarding CO2 conversion, a decrease was observed when increasing the aluminum content, and the AG 1/3 catalytic system reached almost 80% conversion at 450 °C. In this case, the catalytic activity became significant at higher temperatures compared to visible light and at lower temperatures compared to dark conditions. Regarding methane selectivity, all the catalysts reached values close to 100%, except AG 4/1, which reached a lower maximum selectivity of 83% at the highest temperature.
Under all three experimental conditions, catalysts with higher nickel contents exhibited higher CO2 conversion rates, with AG 1/3 obtaining the highest conversion rate under visible light as the catalyst with the largest amount of Ni. This tendency can be attributed to the high rate of nickel dispersion and low particle size in all the samples, as noted by the TEM results. Additionally, the behavior of the catalysts under both light sources can be explained with the absorbance results, in which absorbance in the UV and the visible regions was noticed for all the catalysts. As can be noticed, the activity results with visible light are enhanced with the drop in the band gap energy, which enables the higher use of the visible-light spectrum [52] and triggers the formation of plasmonic hot electrons that are used by the CO2 conversion intermediates [53].
It is observed that high conversions are achieved at medium–high temperatures, with visible light enhancing catalytic activity the most in all cases. The use of a visible light source enables higher conversions at 300 °C in comparison to the other conditions. Regarding the AG-Pt catalyst, an improvement in the conversion rate was expected over the original catalyst; however, it did not surpass the conversions obtained by AG 1/1. Therefore, the synthesis and drying method used to obtain the aerogel-based catalysts was successful in synthesizing catalysts with highly dispersed nickel, including those with higher loadings, which were active as thermo-photocatalysts, and those that avoided noble metals.
The use of nickel aerogels as catalysts has yet to be developed in the CO2 methanation reaction, despite conventional Ni catalysts being widely studied. Lin et al. synthesized Ni-based mesoporous alumina catalysts and tested them in a temperature range between 200 and 500 °C [54]. Their reported conversions are similar to those obtained with our aerogels under visible light (77% with their catalysts at 360 °C, 80% with AG 1/3 at 350 °C); however, considering that only 1/100 of the catalyst mass (500 vs. 5 mg) was used in our case, the productivity of our catalysts is much higher.
As previously stated, one of the most used metals as a catalyst for CO2 methanation is the noble metal ruthenium. Garbarino et al. studied the catalytic activity of a pre-reduced commercial 3 wt% Ru/Al2O3 catalyst in the temperature range of 250 to 500 °C [19]. They obtained CO2 conversion values similar to our maximum at 375 °C and reached their highest value (around 80%) at 400 °C. Even if the metal content of their catalyst was significantly lower than our AG 1/3 catalyst, both achieved similar CO2 conversions. From an economic point of view, nickel is much cheaper than ruthenium, so even if our catalysts have a higher amount of metal, they remain less expensive overall.
In another study, Italiano et al. investigated the CO2 methanation reaction using Ni-based catalysts, synthesized via solution combustion, and tested in a temperature range from 250 to 500 °C [55]. Their catalyst started to exhibit significant activity when 400 °C was reached (57% of CO2 conversion); when compared to the results obtained in the present work, this indicates notably lower efficiency. Furthermore, they employed 20 times higher mass for the catalyst compared to ours, emphasizing the high productivity of aerogel-based materials.

3. Materials and Methods

3.1. Catalyst Preparation

The Ni/Al aerogels were prepared following the sol–gel method reported by Gash et al. [56], in which propylene oxide is the gelation agent. Afterward, the supercritical drying of the gels was performed using CO2 as the solvent. The reagents used in this synthesis were nickel (II) chloride hexahydrate (NiCl2·6H2O; 99.9% purity; Sigma-Aldrich, St. Louis, MO, USA) and aluminum nitrate (III) nonahydrate (Al(NO3)3·9H2O; >98% purity; Honeywell, Seelze, Germany) as metal precursors; liquid propylene oxide (PO; >99.5% purity; Sigma-Aldrich, Steinheim, Germany) as the gelation agent; and ethanol (EtOH; pure; PanReac, Darmstadt, Germany) as the solvent. The metal precursors were dissolved and stirred in ethanol with different Al:Ni molar ratios (1:3, 1:2, 1:1, 2:1, 3:1, and 4:1) until the complete dissolution of the salts (total 3.12 mmol in the case of 1:1) was achieved, followed by the slow addition of as received PO, the volume of which varied with the molar proportion of each material (PO mol/precursor mol = 11).
After aging for 24 h and several cleanings with EtOH, supercritical drying was carried out. Small pieces of the as-synthesized gel were placed inside the dryer (10 °C, 1 bar), and it was filled with liquid CO2 (lCO2), reaching a pressure of 50 bar. Then, several lCO2 exchanges were made to remove the remaining EtOH, as well as other impurities. Finally, the temperature of the dryer was raised to 40 °C (70–80 bar), enabling the lCO2 to reach the supercritical state (scCO2). These conditions were maintained for 1 h before depressurizing at a slow rate. After performing the supercritical drying, the aerogels were calcined at 700 °C in air [57,58] at a heating rate of 2 °C/min for 4 h. The prepared samples were named AG 1/3, AG 1/2, AG 1/1, AG 2/1, AG 3/1, and AG 4/1.
Considering the previous good performance results of Pt as a catalyst for CO2 methanation, AG 1/1 with a 0.25 wt% loading of Pt was synthesized. To perform the synthesis, the Pt precursor ([Pt(NH3)4][NO3]2; 99.9% purity; Thermo Scientific, Kandel, Germany) was added in the dissolution step. The rest of the synthesis and drying methods were followed as described above, and this sample was named AG-Pt.

3.2. Catalyst Characterization

Inductively coupled plasma–optical emission spectroscopy (ICP-OES) experiments were conducted (Optima 2000 DV; Perkin Elmer, Tres Cantos, Spain) to study the metal loading of the samples. They were digested in an acid solution before analysis at a plasma flux of 15 L/min.
The BET surface area (SBET) and the total pore volume (Vp) were calculated with the N2 adsorption and desorption isotherms (Autosorb iQ3; Quantachrome (Anton Paar), Madrid, Spain) after prior degasification at 250 °C.
X-ray diffraction (XRD) analysis was carried out (X’pert PRO automatic diffractometer; Panalytical, San Sebastián de los Reyes, Spain) at 40 kV to elucidate the crystalline structure of the samples, applying a Cu-Kα radiation of λ = 1.5418 Å, with a scanning 2θ range of 5 to 80°.
The reducibility of the materials was studied through H2 temperature-programmed reduction (H2-TPR) experiments (AutoChem II; Micromeritics, L’Hospitalet de Llobregat, Spain) after the calcination of the catalysts. They were heated up to 850 °C at a rate of 10 °C/min, under a constant 5% H2/Ar flow, after a previous drying step with He at 200 °C for 30 min.
Transmission electron microscopy (TEM) was performed (Talos F200X, 200 kV; Thermo Fisher, Alcobendas, Spain) to observe the morphology of the obtained nanoparticles and measure the particle size distribution. Elemental mapping was performed using the HAADF-EDX method.
The binding energies, and thus, the oxidation states of the metals on the samples were measured via X-ray photoelectron spectroscopy (XPS) analyses (5000 VersaProbe II; PHI (Irida), Madrid, Spain), using a monochromatic X-ray source of Al (Kα, 1486.6 eV) and a beam diameter of 200.0 µm. The signal corresponding to the C 1s at 284.8 eV was used as an internal standard to correct the binding energy shift.
UV-visible diffuse reflectance spectroscopy (UV-vis DRS) was used to understand the photochemical behavior of the catalysts (V-770; Jasco, Madrid, Spain), measuring their absorbance at different wavelengths. In addition, this technique enabled us to calculate the band gap of each material. The diffuse reflectance of the samples was measured in a range of 2200 to 200 nm, and prior to that, a blank background correction was made.

3.3. Activity Tests

The thermo-photocatalytic activity tests were performed in a photoreactor with an 8 mm open quartz window. Two different light sources can be attached to this reactor for ultraviolet light (365 nm, 3.4 eV) and visible light (470 nm, 2.7 eV). The intensity of the light was analyzed using a GL Optic Spectis 1.0 spectrometer, which was coupled to an Opti Sphere 48 (GL Optic, Puszczykowo, Poland), and a light intensity of 2.4 W/cm2 was fixed for both light sources. The bench-scale plant (PID Eng&Tech, Alcobendas, Spain) reached temperatures of up to 450 °C and pressures of 10 bar. In all cases, the total gas flow was 50 mLN/min with a stoichiometric gas composition (CO2:H2 = 1:4). The composition of the products was analyzed using an online CompactGC 4.0 gas chromatograph (GAS, Breda, The Netherlands), which had both TCD and FID detectors.
Before each activity test, the catalysts were sieved to a particle size of between 0.42 and 0.5 mm and subsequently reduced in a flow of 20% H2/N2. The reduction temperature for each material was determined by considering the H2-TPR results. For all catalysts, a common temperature of 600 °C was used to perform the reduction, at a heating rate of 5 °C /min and for 2 h. During the reaction, the temperature varied from 300 to 450 °C, and the pressure was fixed at 10 bar. In each test, a catalyst mass of 5 mg and a ramp of 5 °C /min were used to heat the system to each temperature. The system was maintained for 1.5–2 h while the effluent gas composition was constantly analyzed.
The activity results, in terms of CO2 conversion and CH4 selectivity, were calculated using the following equations:
C O 2   C o n v e r s i o n   % = C O 2 i n C O 2 o u t C O 2 i n × 100
C H 4   S e l e c t i v i t y   ( % ) = C H 4 o u t C H 4 o u t + C O o u t × 100

4. Conclusions

The objective of this study was to develop aluminum and nickel aerogel-based catalysts for the methanation of CO2 to decrease emissions and reduce the greenhouse gas effect. Different Al/Ni molar ratios were established, and a sol–gel process followed by supercritical drying with CO2 was chosen to synthesize the aerogels. Various characterization techniques confirmed the large specific surface areas of these materials, reaching values of almost 500 m2/g in some cases, alongside the presence of reducible NiO. When performing the thermo-photocatalytic activity tests, a trend in the results was observed depending on the light conditions used. In dark conditions, significant CO2 conversion was reached at temperatures above 400 °C, such as 65.9%, when using the AG 1/2 catalyst. The visible light source helped improve methanation activity in terms of CO2 conversion. Values as high as 80% were obtained in the case of AG 1/3. UV light provided activity results between dark and visible light conditions, achieving 75% CO2 conversion with the same catalyst at 450 °C. Regarding metallic loadings, catalysts with higher Ni loadings produced the best results due to the high levels of active metal in the samples. Due to their lightweight nature, large surface area, and high catalytic activity, these materials could provide an alternative for using the Sabatier reaction.
Even if these catalysts provide good thermo-photocatalytic results, some future work should be performed in order to address some of the issues recognized in the present work. For instance, activity tests should be performed at lower temperatures to try to reduce energy consumption and highlight the photocatalytic contribution of the materials. Furthermore, stability tests will be carried out in the future, with the objective of testing out the long-term stability of these materials and evaluating their industrial viability.

Supplementary Materials

The following supporting information can be downloaded at https://www.mdpi.com/article/10.3390/catal15070686/s1, Figure S1: N2 adsorption–desorption isotherms of the different catalysts; Figure S2: TEM microgrpahs and EDX mapping of AG 2/1, AG 3/1, AG 4/1, AG-Pt, and AG 1/2; Figure S3: Tauc plots for all catalysts, used to calculate the direct band gap energy.

Author Contributions

Conceptualization, V.L.B. and H.E.; methodology, H.E. and D.E.; investigation, D.E.; writing—original draft preparation, review and editing, D.E. and V.L.B. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by the University of the Basque Country (UPV/EHU), the Basque Government (Project: IT993-16), the Spanish Ministry of Economy, Industry and Competitiveness (PDI2020-112889RB-10), and SINTETIK (KK-2024).

Data Availability Statement

The data that support the findings of this study are available from the corresponding author upon reasonable request.

Acknowledgments

The authors thank the technical and human support provided by SGIker (UPV/EHU/ ERDF, EU), especially Aitor Larrañaga and Juan Carlos Raposo.

Conflicts of Interest

The authors declare no conflicts of interest.

Abbreviations

The following abbreviations are used in this manuscript:
AGAerogel
lCO2Liquid CO2
ICP-OESInductively Coupled Plasma–Optical Emission Spectroscopy
BETBrunauer, Emmet, and Teller Theory
BJHBarrett–Joyner–Halenda
XRDX-Ray Diffraction
H2-TPRH2 Temperature-Programmed Reduction
TEMTransmission Electron Microscopy
HAADFHigh-Angle Annular Dark Field
EDXEnergy-Dispersive X-Ray Spectroscopy
UV-Vis DRSUV-Visible Diffuse Reflectance Spectroscopy
XPSX-Ray Photoelectron Spectroscopy
TCDThermal Conductivity Detector
FIDFlame Ionization Detector

References

  1. Liu, Z.; Deng, Z.; Davis, S.J.; Ciais, P. Global Carbon Emissions in 2023. Nat. Rev. Earth Environ. 2024, 5, 253–254. [Google Scholar] [CrossRef]
  2. Lewandowska-Bernat, A.; Desideri, U. Opportunities of power-to-gas technology in different energy systems architectures. Appl. Energy 2018, 228, 57–67. [Google Scholar] [CrossRef]
  3. Karakaya, C.; Kee, R.J. Progress in the direct catalytic conversion of methane to fuels and chemicals. Prog. Energy Combust. Sci. 2016, 55, 60–97. [Google Scholar] [CrossRef]
  4. Sun, L.; Wang, Y.; Guan, N.; Li, L. Methane Activation and Utilization: Current Status and Future Challenges. Energy Technol. 2019, 8, 1900826. [Google Scholar] [CrossRef]
  5. Ma, J.; Zhu, C.; Mao, K.; Jiang, W.; Low, J.; Duan, D.; Ju, H.; Liu, D.; Wang, K.; Zang, Y.; et al. Sustainable methane utilization technology via photocatalytic halogenation with alkali halides. Nat. Commun. 2023, 14, 1–8. [Google Scholar] [CrossRef] [PubMed]
  6. van Schyndel, J.; Goos, E.; Naumann, C.; Hardi, J.S.; Oschwald, M. Effects of Compounds in Liquefied Methane on Rocket Engine Operation. Aerospace 2022, 9, 698. [Google Scholar] [CrossRef]
  7. Vogt, C.; Monai, M.; Kramer, G.J.; Weckhuysen, B.M. The renaissance of the Sabatier reaction and its applications on Earth and in space. Nat. Catal. 2019, 2, 188–197. [Google Scholar] [CrossRef]
  8. Liu, Y.; Zhang, Z.; Tian, J.; Liao, Y.; Zhai, J.; Fang, X.; Xu, X.; Wang, X. Tuning CO adsorption states via enhanced metal-support interaction for boosting CO2 methanation activity of Ru/rutile-TiO2. Fuel 2023, 360, 130600. [Google Scholar] [CrossRef]
  9. Rueda-Navarro, C.M.; González-Fernández, M.; Cabrero-Antonino, M.; Dhakshinamoorthy, A.; Ferrer, B.; Baldoví, H.G.; Navalón, S. Solar-Assisted CO2 Methanation via Photocatalytic Sabatier Reaction by Calcined Titanium-based Organic Framework Supported RuOx Nanoparticles. ChemCatChem 2024, 16, e202400991. [Google Scholar] [CrossRef]
  10. Yang, C.; Wang, W.; Zhuo, H.; Shen, Z.; Zhang, T.; Yang, X.; Huang, Y. Phase engineering of Ru-based nanocatalysts for enhanced activity toward CO2 methanation. Chin. J. Catal. 2024, 61, 226–236. [Google Scholar] [CrossRef]
  11. Sun, S.; Higham, M.D.; Zhang, X.; Catlow, C.R.A. Multiscale Investigation of the Mechanism and Selectivity of CO2 Hydrogenation over Rh(111). ACS Catal. 2024, 14, 5503–5519. [Google Scholar] [CrossRef] [PubMed]
  12. Gómez-Flores, K.; Solís-García, A.; Lam, S.J.; Cervantes-Gaxiola, M.; Castillo-López, R.; Leyva, J.R.; Gómez, S.; Flores-Aquino, E.; Zepeda, T. Modulating selectivity in CO2 Methanation through Rhodium catalysts supported on Zirconia-Chemically grafted SBA-15. Mol. Catal. 2024, 558, 114035. [Google Scholar] [CrossRef]
  13. Dai, X.; Sun, Y. Mechanism of photocatalytic CO2 methanation on ultrafine Rh nanoparticles. Nanoscale Horizons 2024, 9, 627–636. [Google Scholar] [CrossRef] [PubMed]
  14. Meloni, E.; Cafiero, L.; Renda, S.; Martino, M.; Pierro, M.; Palma, V. Ru- and Rh-Based Catalysts for CO2 Methanation Assisted by Non-Thermal Plasma. Catalysts 2023, 13, 488. [Google Scholar] [CrossRef]
  15. Wu, Y.; Bhalothia, D.; Chou, J.-P.; Lee, G.-W.; Hu, A.; Chen, T.-Y. Electronic metal–support interaction in Pd/CoNi-hydroxides with enhanced CO adsorption for boosting CO2 methanation. Sustain. Energy Fuels 2024, 9, 847–854. [Google Scholar] [CrossRef]
  16. Yang, T.; Beniwal, A.; Bhalothia, D.; Yan, C.; Wang, C.-H.; Chen, T.-Y. Oxygen vacancies coupled with surface silicide facilitate CO2 activation at near-room temperature for efficient methane productivity on Ni-oxide supported Pd nanoparticles. Sustain. Energy Fuels 2024, 8, 3399–3411. [Google Scholar] [CrossRef]
  17. Danielis, M.; Jiménez, J.D.; Rui, N.; Moncada, J.; Betancourt, L.E.; Trovarelli, A.; Rodriguez, J.A.; Senanayake, S.D.; Colussi, S. Tuning the hydrogenation of CO2 to CH4 over mechano-chemically prepared palladium supported on ceria. Appl. Catal. A Gen. 2023, 660, 119185. [Google Scholar] [CrossRef]
  18. Memon, M.A.; Jiang, Y.; Hassan, M.A.; Ajmal, M.; Wang, H.; Liu, Y. Heterogeneous Catalysts for Carbon Dioxide Methanation: A View on Catalytic Performance. Catalysts 2023, 13, 1514. [Google Scholar] [CrossRef]
  19. Garbarino, G.; Bellotti, D.; Riani, P.; Magistri, L.; Busca, G. Methanation of carbon dioxide on Ru/Al2O3 and Ni/Al2O3 catalysts at atmospheric pressure: Catalysts activation, behaviour and stability. Int. J. Hydrogen Energy 2015, 40, 9171–9182. [Google Scholar] [CrossRef]
  20. Spataru, D.; Carvalho, G.; Quindimil, A.; Lopes, J.M.; Henriques, C.; Bacariza, C. CO2 methanation under more realistic conditions: Influence of O2 and H2O on Ni-based catalysts’ performance. Chem. Eng. J. 2024, 496, 153709. [Google Scholar] [CrossRef]
  21. Ma, L.; Ye, R.; Huang, Y.; Reina, T.R.; Wang, X.; Li, C.; Zhang, X.L.; Fan, M.; Zhang, R.; Liu, J. Enhanced low-temperature CO2 methanation performance of Ni/ZrO2 catalysts via a phase engineering strategy. Chem. Eng. J. 2022, 446, 137031. [Google Scholar] [CrossRef]
  22. Zhang, Z.; Yu, Z.; Feng, K.; Yan, B. Eu3+ doping-promoted Ni-CeO2 interaction for efficient low-temperature CO2 methanation. Appl. Catal. B Environ. 2022, 317, 121800. [Google Scholar] [CrossRef]
  23. Cañón-Alvarado, M.; Blanco, C.; Daza, C. Exploring dolomite as a promising support for Ni catalysts in CO2 methanation. J. Environ. Chem. Eng. 2024, 12, 112224. [Google Scholar] [CrossRef]
  24. Liu, F.; Park, Y.S.; Diercks, D.; Kazempoor, P.; Duan, C. Enhanced CO2 Methanation Activity of Sm0.25Ce0.75O2-δ–Ni by Modulating the Chelating Agents-to-Metal Cation Ratio and Tuning Metal–Support Interactions. ACS Appl. Mater. Interfaces 2022, 14, 13295–13304. [Google Scholar] [CrossRef] [PubMed]
  25. Kristiani, A.; Takeishi, K. CO2 methanation over nickel-based catalyst supported on yttria-stabilized zirconia. Catal. Commun. 2022, 165, 106435. [Google Scholar] [CrossRef]
  26. Song, W.; Wang, K.; Wang, X.; Ma, Q.; Zhao, T.-S.; Bae, J.W.; Gao, X.; Zhang, J. Boosting low temperature CO2 methanation by tailoring Co species of CoAlO catalysts. Chem. Eng. Sci. 2024, 298, 120405. [Google Scholar] [CrossRef]
  27. Fu, W.; Liu, S.; He, Y.; Chen, J.; Ren, J.; Chen, H.; Sun, R.; Tang, Z.; Mebrahtu, C.; Zeng, F. CO2 hydrogenation over CoAl based catalysts: Effects of cobalt-metal oxide interaction. Appl. Catal. A Gen. 2024, 678, 119720. [Google Scholar] [CrossRef]
  28. O’cOnnell, G.E.; Tan, T.H.; Yuwono, J.A.; Wang, Y.; Kheradmand, A.; Jiang, Y.; Kumar, P.V.; Amal, R.; Scott, J.; Lovell, E.C. Seeing the light: The role of cobalt in light-assisted CO2 methanation. Appl. Catal. B Environ. 2023, 343, 123507. [Google Scholar] [CrossRef]
  29. Mehrabi, M.; Ashjari, M.; Meshkani, F. Cobalt supported on silica–alumina nanocomposite for use in CO2 methanation process: Effects of Si/Al molar ratio and Co loading on catalytic activity. Res. Chem. Intermed. 2023, 50, 219–238. [Google Scholar] [CrossRef]
  30. Prabhakar, J.K.; Apte, P.A.; Deo, G. Exploring optimal total metal loading of Ni3Fe/Al2O3 catalyst for CO2 methanation and its kinetic model. Fuel 2024, 367, 131447. [Google Scholar] [CrossRef]
  31. Li, J.; Xu, Q.; Han, Y.; Guo, Z.; Zhao, L.; Cheng, K.; Zhang, Q.; Wang, Y. Efficient photothermal CO2 methanation over NiFe alloy nanoparticles with enhanced localized surface plasmon resonance effect. Sci. China Chem. 2023, 66, 3518–3524. [Google Scholar] [CrossRef]
  32. Yan, P.; Peng, H.; Zhang, X.; Rabiee, H.; Ahmed, M.; Weng, Y.; Rozhkovskaya, A.; Vogrin, J.; Konarova, M.; Zhu, Z. Unlocking the role of Ni-Fe species in CO2 methanation. Fuel 2024, 374, 132373. [Google Scholar] [CrossRef]
  33. Aziz, M.A.; Jalil, A.A.; Hassan, N.S.; Bin Bahari, M.; Hatta, A.H.; Abdullah, T.A.T.; Jusoh, N.W.C.; Setiabudi, H.D.; Saravanan, R. Kinetic exploration of CO2 methanation over nickel loaded on fibrous mesoporous silica nanoparticles (CHE-SM). Process. Saf. Environ. Prot. 2024, 186, 1229–1241. [Google Scholar] [CrossRef]
  34. Ilsemann, J.; Murshed, M.M.; Gesing, T.M.; Kopyscinski, J.; Bäumer, M. On the support dependency of the CO2 methanation–decoupling size and support effects. Catal. Sci. Technol. 2021, 11, 4098–4114. [Google Scholar] [CrossRef]
  35. Zhou, L.; Guo, X.; Hu, X.; Zhang, Y.; Cheng, J.; Guo, Q. CO2 methanation reaction over La-modified NiAl catalysts derived from hydrotalcite-like precursors. Fuel 2024, 362, 130888. [Google Scholar] [CrossRef]
  36. Abahussain, A.A.M.; Al-Fatesh, A.S.; Rajput, Y.B.; Osman, A.I.; Alreshaidan, S.B.; Ahmed, H.; Fakeeha, A.H.; Al-Awadi, A.S.; El-Salamony, R.A.; Kumar, R. Impact of Sr Addition on Zirconia–Alumina-Supported Ni Catalyst for COx-Free CH4 Production via CO2 Methanation. ACS Omega 2024, 9, 9309–9320. [Google Scholar] [CrossRef] [PubMed]
  37. Guo, X.; Traitangwong, A.; Hu, M.; Zuo, C.; Meeyoo, V.; Peng, Z.; Li, C. Carbon Dioxide Methanation over Nickel-Based Catalysts Supported on Various Mesoporous Material. Energy Fuels 2018, 32, 3681–3689. [Google Scholar] [CrossRef]
  38. Bacariza, M.; Graça, I.; Lopes, J.; Henriques, C. Enhanced activity of CO2 hydrogenation to CH4 over Ni based zeolites through the optimization of the Si/Al ratio. Microporous Mesoporous Mater. 2018, 267, 9–19. [Google Scholar] [CrossRef]
  39. Quindimil, A.; De-La-Torre, U.; Pereda-Ayo, B.; González-Marcos, J.A.; González-Velasco, J.R. Ni catalysts with La as promoter supported over Y- and BETA- zeolites for CO2 methanation. Appl. Catal. B Environ. 2018, 238, 393–403. [Google Scholar] [CrossRef]
  40. Gao, B.; Wang, I.-W.; Ren, L.; Haines, T.; Hu, J. Catalytic Performance and Reproducibility of Ni/Al2O3 and Co/Al2O3 Mesoporous Aerogel Catalysts for Methane Decomposition. Ind. Eng. Chem. Res. 2018, 58, 798–807. [Google Scholar] [CrossRef]
  41. Hu, F.; Chen, X.; Tu, Z.; Lu, Z.-H.; Feng, G.; Zhang, R. Graphene Aerogel Supported Ni for CO2 Hydrogenation to Methane. Ind. Eng. Chem. Res. 2021, 60, 12235–12243. [Google Scholar] [CrossRef]
  42. Canales, R.; Barrio, V.L. Photo- and Thermocatalytic CO2 Methanation: A Comparison of Ni/Al2O3 and Ni–Ce Hydrotalcite-Derived Materials under UV and Visible Light. Materials 2023, 16, 5907. [Google Scholar] [CrossRef] [PubMed]
  43. Zhang, B.; Yao, P.; Li, F.; Pan, L.; Xiong, W.; Zhang, Y.; Li, X. Catalytic Pyrolysis of Waste Textiles for Hydrogen-Rich Syngas Production over NiO/Al2O3 Catalyst. Processes 2024, 13, 15. [Google Scholar] [CrossRef]
  44. Oliveira, G.V.; de Macedo, V.; Urquieta-González, E.A.; Magriotis, Z.M.; Pereira, C.A. Fe2O3/γ-Al2O3 and NiO/γ-Al2O3 catalysts for the selective catalytic oxidation of ammonia. Catal. Today 2024, 444, 114991. [Google Scholar] [CrossRef]
  45. Tian, T.; Zhao, Z.-M.; Zhang, Y.; Jing, J.; Li, W.-Y. Insight into the Nature of Nickel Active Sites on the NiAl2O4 Catalyst for Phenanthrene Hydrogenation Saturation. ACS Omega 2025, 10, 8303–8313. [Google Scholar] [CrossRef] [PubMed]
  46. Valecillos, J.; Iglesias-Vázquez, S.; Landa, L.; Remiro, A.; Bilbao, J.; Gayubo, A.G. Insights into the Reaction Routes for H2 Formation in the Ethanol Steam Reforming on a Catalyst Derived from NiAl2O4 Spinel. Energy Fuels 2021, 35, 17197–17211. [Google Scholar] [CrossRef] [PubMed]
  47. Garbarino, G.; Chitsazan, S.; Phung, T.K.; Riani, P.; Busca, G. Preparation of supported catalysts: A study of the effect of small amounts of silica on Ni/Al2O3 catalysts. Appl. Catal. A Gen. 2015, 505, 86–97. [Google Scholar] [CrossRef]
  48. Zhang, L.; Wang, X.; Chen, C.; Zou, X.; Shang, X.; Ding, W.; Lu, X. Investigation of mesoporous NiAl2O4/MOx (M = La, Ce, Ca, Mg)–γ-Al2O3 nanocomposites for dry reforming of methane. RSC Adv. 2017, 7, 33143–33154. [Google Scholar] [CrossRef]
  49. Garbarino, G.; Campodonico, S.; Perez, A.R.; Carnasciali, M.M.; Riani, P.; Finocchio, E.; Busca, G. Spectroscopic characterization of Ni/Al2O3 catalytic materials for the steam reforming of renewables. Appl. Catal. A Gen. 2013, 452, 163–173. [Google Scholar] [CrossRef]
  50. Kumar, S.; Ojha, A.K. Ni, Co and Ni–Co codoping induced modification in shape, optical band gap and enhanced photocatalytic activity of CeO2 nanostructures for photodegradation of methylene blue dye under visible light irradiation. RSC Adv. 2015, 6, 8651–8660. [Google Scholar] [CrossRef]
  51. Castillo-Saenz, J.; Nedev, N.; Valdez-Salas, B.; Curiel-Alvarez, M.; Mendivil-Palma, M.I.; Hernandez-Como, N.; Martinez-Puente, M.; Mateos, D.; Perez-Landeros, O.; Martinez-Guerra, E. Properties of Al2O3 Thin Films Grown by PE-ALD at Low Temperature Using H2O and O2 Plasma Oxidants. Coatings 2021, 11, 1266. [Google Scholar] [CrossRef]
  52. Roškarič, M.; Žerjav, G.; Zavašnik, J.; Pintar, A. The influence of synthesis conditions on the visible-light triggered photocatalytic activity of g-C3N4/TiO2 composites used in AOPs. J. Environ. Chem. Eng. 2022, 10, 107656. [Google Scholar] [CrossRef]
  53. Canales, R.; Gil-Calvo, M.; Barrio, V.L. UV- and visible-light photocatalysis using Ni–Co bimetallic and monometallic hydrotalcite-like materials for enhanced CO2 methanation in sabatier reaction. Heliyon 2023, 9, e18456. [Google Scholar] [CrossRef] [PubMed]
  54. Lin, J.; Ma, C.; Luo, J.; Kong, X.; Xu, Y.; Ma, G.; Wang, J.; Zhang, C.; Li, Z.; Ding, M. Preparation of Ni based mesoporous Al2O3 catalyst with enhanced CO2 methanation performance. RSC Adv. 2019, 9, 8684–8694. [Google Scholar] [CrossRef] [PubMed]
  55. Italiano, C.; Llorca, J.; Pino, L.; Ferraro, M.; Antonucci, V.; Vita, A. CO and CO2 methanation over Ni catalysts supported on CeO2, Al2O3 and Y2O3 oxides. Appl. Catal. B Environ. 2020, 264, 118494. [Google Scholar] [CrossRef]
  56. Gash, A.E.; Satcher, J.H.; Simpson, R.L. Monolithic nickel(II)-based aerogels using an organic epoxide: The importance of the counterion. J. Non-Crystalline Solids 2004, 350, 145–151. [Google Scholar] [CrossRef]
  57. Gil Seo, J.; Youn, M.H.; Bang, Y.; Song, I.K. Effect of Ni/Al atomic ratio of mesoporous Ni–Al2O3 aerogel catalysts on their catalytic activity for hydrogen production by steam reforming of liquefied natural gas (LNG). Int. J. Hydrogen Energy 2010, 35, 12174–12181. [Google Scholar] [CrossRef]
  58. Hao, Z.; Zhu, Q.; Lei, Z.; Li, H. CH4–CO2 reforming over Ni/Al2O3 aerogel catalysts in a fluidized bed reactor. Powder Technol. 2007, 182, 474–479. [Google Scholar] [CrossRef]
Figure 1. XRD profiles of calcined AGs. α signals correspond to NiO, while β signals correspond to NiAl2O4 spinel.
Figure 1. XRD profiles of calcined AGs. α signals correspond to NiO, while β signals correspond to NiAl2O4 spinel.
Catalysts 15 00686 g001
Figure 2. H2-TPR profiles of calcined AGs. The different lines are attributed to the contribution of different species: the green line corresponds to the weak NiO-Al2O3 interaction; the red line corresponds to the strong interaction between NiO and Al2O3 and the dark blue line corresponds to the contribution of the NiAl2O4 spinel.
Figure 2. H2-TPR profiles of calcined AGs. The different lines are attributed to the contribution of different species: the green line corresponds to the weak NiO-Al2O3 interaction; the red line corresponds to the strong interaction between NiO and Al2O3 and the dark blue line corresponds to the contribution of the NiAl2O4 spinel.
Catalysts 15 00686 g002
Figure 3. TEM micrographs (left) and EDX mapping (right) of (a) AG 1/1 V, (b) AG 1/1, and (c) AG 1/3.
Figure 3. TEM micrographs (left) and EDX mapping (right) of (a) AG 1/1 V, (b) AG 1/1, and (c) AG 1/3.
Catalysts 15 00686 g003
Figure 4. UV-visible DRS absorption spectra of the different Ni aerogels.
Figure 4. UV-visible DRS absorption spectra of the different Ni aerogels.
Catalysts 15 00686 g004
Figure 5. XPS spectra of the different calcined catalysts. The different lines correspond to different signals: dark blue and light green correspond to surface Ni; red and purple correspond to Ni2+ coming from NiO; orange and dark green are satellites of the red and purple ones.
Figure 5. XPS spectra of the different calcined catalysts. The different lines correspond to different signals: dark blue and light green correspond to surface Ni; red and purple correspond to Ni2+ coming from NiO; orange and dark green are satellites of the red and purple ones.
Catalysts 15 00686 g005
Figure 6. Thermo-photocatalytic activity results of aerogels under different conditions: (a) in the dark; (b) under visible light (λ = 470 nm); and (c) under UV light (λ = 365 nm).
Figure 6. Thermo-photocatalytic activity results of aerogels under different conditions: (a) in the dark; (b) under visible light (λ = 470 nm); and (c) under UV light (λ = 365 nm).
Catalysts 15 00686 g006
Table 1. Compositional and textural properties of the prepared catalysts.
Table 1. Compositional and textural properties of the prepared catalysts.
SampleNominal Al/Ni Molar RatioAl/Ni Molar RatioAl/Pt Molar Ratio (wt %)SBET (m2/g)Vt, pore (cm3/g)DBJH (nm)
AG 1/30.330.30---156 ± 130.567.5
AG 1/20.500.41---207 ± 180.9817.8
AG 1/11.000.98---215 ± 192.2824.6
AG-Pt1.000.971863 (0.1%)209 ± 190.846.8
AG 2/12.001.93---270 ± 253.2666.6
AG 3/13.002.94---492 ± 505.5067.0
AG 4/14.003.79---358 ± 353.6311.5
Table 2. Band gap energy values, calculated by the Tauc equation.
Table 2. Band gap energy values, calculated by the Tauc equation.
SampleBand Gap Energy (eV)
AG 1/33.38
AG 1/23.36
AG 1/13.40
AG-Pt3.39
AG 2/13.46
AG 3/13.52
AG 4/13.87
Table 3. Al/Ni ratio of the different catalysts obtained by XPS after calcination at 700 °C.
Table 3. Al/Ni ratio of the different catalysts obtained by XPS after calcination at 700 °C.
CatalystAl/Ni Atomic RatioAl/Pt Atomic Ratio
AG 1/31.00---
AG 1/21.07---
AG 1/11.67---
AG-Pt1.632068
AG 2/12.04---
AG 3/13.17---
AG 4/13.93---
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

Estevez, D.; Etxeberria, H.; Barrio, V.L. The Development of Ni-Al Aerogel-Based Catalysts via Supercritical CO2 Drying for Photocatalytic CO2 Methanation. Catalysts 2025, 15, 686. https://doi.org/10.3390/catal15070686

AMA Style

Estevez D, Etxeberria H, Barrio VL. The Development of Ni-Al Aerogel-Based Catalysts via Supercritical CO2 Drying for Photocatalytic CO2 Methanation. Catalysts. 2025; 15(7):686. https://doi.org/10.3390/catal15070686

Chicago/Turabian Style

Estevez, Daniel, Haritz Etxeberria, and Victoria Laura Barrio. 2025. "The Development of Ni-Al Aerogel-Based Catalysts via Supercritical CO2 Drying for Photocatalytic CO2 Methanation" Catalysts 15, no. 7: 686. https://doi.org/10.3390/catal15070686

APA Style

Estevez, D., Etxeberria, H., & Barrio, V. L. (2025). The Development of Ni-Al Aerogel-Based Catalysts via Supercritical CO2 Drying for Photocatalytic CO2 Methanation. Catalysts, 15(7), 686. https://doi.org/10.3390/catal15070686

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