Next Article in Journal
Exploring the Potential of Black Soldier Fly Larval Proteins as Bioactive Peptide Sources through in Silico Gastrointestinal Proteolysis: A Cheminformatic Investigation
Next Article in Special Issue
Gold(I)-Catalyzed Direct Alkyne Hydroarylation in Ionic Liquids: Mechanistic Insights
Previous Article in Journal
Methane Oxidation over the Zeolites-Based Catalysts
Previous Article in Special Issue
The Impact of Functionality and Porous System of Nanostructured Carriers Based on Metal–Organic Frameworks of UiO-66-Type on Catalytic Performance of Embedded Au Nanoparticles in Hydroamination Reaction
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Gold and Ceria Modified NiAl Hydrotalcite Materials as Catalyst Precursors for Dry Reforming of Methane

by
Valeria La Parola
1,
Giuseppe Pantaleo
1,*,
Leonarda Francesca Liotta
1,
Anna Maria Venezia
1,*,
Margarita Gabrovska
2,
Dimitrinka Nikolova
2 and
Tatyana Tabakova
2,*
1
Istituto per lo Studio dei Materiali Nanostrutturati, CNR, 90146 Palermo, Italy
2
Institute of Catalysis, Bulgarian Academy of Sciences, 1113 Sofia, Bulgaria
*
Authors to whom correspondence should be addressed.
Catalysts 2023, 13(3), 606; https://doi.org/10.3390/catal13030606
Submission received: 17 February 2023 / Revised: 10 March 2023 / Accepted: 13 March 2023 / Published: 16 March 2023
(This article belongs to the Collection Gold Catalysts)

Abstract

:
Structured hydrotalcite NiAl-HT material with Ni/Al atomic ratio of 2.5 was prepared by co-precipitation of Ni and Al nitrate precursors and then modified by the addition of 1 wt% Ce and/or 3 wt% Au species. The obtained materials, after calcination at 600 °C, were characterized by XRD, XPS and TPR. Their catalytic performance was tested through dry reforming of methane (DRM) and by the temperature-programmed surface reaction of methane (TPSR-CH4). Thermal gravimetry analysis (TGA) of the spent catalysts was performed to determine the amount of carbon accumulated during the reaction. The effects of the addition of cerium as a support promoter and gold as nickel promoter and the sequential addition of cerium and gold on the structural properties and on the catalytic efficiency were investigated. Under the severe condition of high space velocity (600,000 mL g−1 h−1), all the catalysts were quite active, with values of CH4 conversion between 67% and 74% at 700 °C. In particular, the combination of cerium and gold enhanced the CH4 conversion up to 74%. Both additives, individually and simultaneously, enhanced the nickel dispersion with respect to the unpromoted NiAl and favored the reducibility of the nickel. During DRM all the catalysts formed graphitic carbon, contributing to their deactivation. The lower carbon gasification temperature of the promoted catalysts confirmed a positive effect played by Ce and Au in assisting the formation of an easier-to-remove carbon. The positive effect was testified by the better stability of the Ce/NiAl with respect to the other catalysts. In the gold-containing samples, this effect was neutralized by Au diffusing towards the catalyst surface during DRM, masking the nickel active sites. TPSR-CH4 test highlighted different CH4 activation capability of the catalysts. Furthermore, the comparison of the deposited carbon features (amount and removal temperature) of the DRM and TPSR spent catalysts indicated a superior activation of CO2 by the Au/Ce/NiAl, to be related to the close interaction of gold and ceria enhancing the oxygen mobility in the catalyst lattice.

Graphical Abstract

1. Introduction

The energetic crisis connected to geopolitical changes and to the continuous depletion of fossil fuels requires the rapid development of efficient renewable energy sources, such as solar, wind and tidal power, green hydrogen and efficient chemical processes, to produce synthetic fuels and chemicals [1]. Among the relevant chemical processes, dry reforming of methane with CO2 (DRM) is currently the subject of increased attention from the research and industrial community as a prospective method for obtaining fuels and chemicals through the conversion of the syngas produced according to the following equation.
CH4 + CO2 ↔ 2CO + 2H2           (ΔH 298K = + 247 kJ/mol)
The obtained syngas is a mixture of H2 and CO in a 1:1 molecular ratio, ideal for the Fischer-Tropsch synthesis of fuels and for the synthesis of oxygenates. Furthermore, the conversion of CH4 and CO2 into a mixture of H2 and CO, two important chemical precursor molecules, would conceivably contribute to the recycling of the two most green housing effect gases [2,3,4,5,6]. However, the effect of this process in global warming mitigation depends on different factors, such as the source of CO2, (whether from fossil fuel combustion or from organic waste), or the temperature of the reaction [5,7]. The development of this process at the industrial level is still hampered by the lack of appropriate catalysts with high efficiency and low cost. It is well-recognized that nickel-based catalysts are the most suitable in terms of costs over benefits [6]. Nevertheless, nickel suffers from easy deactivation. The loss of efficiency with time is mainly ascribable to coking and to the sintering of the active nickel sites. Coking is the formation of carbon, which, depending on its morphology and structure, may affect the catalyst activity differently [8]. The formation of carbon graphite, favored by larger nickel particles, is highly detrimental, being rather inert and difficult to gasify, differently from the amorphous carbon, which could be easily removed by subsequent treatment with water (C + H2O = CO + H2) or with CO2 [5,9,10]. Limiting the nickel particle size or using supports with a certain degree of basicity would help to mitigate the coking effects [6]. An inverse relationship between nickel particle sizes and DRM activity was observed with an optimal particle size around 2–3 nm for reaction temperature of 500–600 °C [11,12]. According to literature, the size of the nickel particles plays an important role in the nucleation and growth of carbon filament. Generally, the initiation step for carbon formation is more difficult in small particles, and increasing the size beyond 6 nm generates more carbon [13]. The sintering effect occurring during the pretreatment reduction and during the reaction at high temperature is generally avoided using appropriate support design, suitable catalyst synthesis and the addition of proper promoters. The use of different supports and different synthetic procedures is intended to increase the interaction between the nickel and the carrier to limit the otherwise free particle growth [5]. SiO2 and Al2O3 are the most used supports for this type of catalysts, each having characteristics beneficial to the catalytic activity of the supported nickel [14]. SiO2, being a non-interacting support, allows an easy reduction of the nickel, but it does not restrain its particle size. On the other hand, Al2O3, which is a much more interactive oxide, limits the nickel particle growth but at the same time inhibits the nickel reduction, with a consequent loss of activity. Different types of promoters have been investigated, aiming to favor the reducibility of the catalyst without enhancing metal sintering, or to modify the basicity/acidity property of the carrier to increase the catalytic stability without reducing the activity [15]. The addition of alkali/alkaline metals, increasing the support basicity, would favor the adsorption and activation of the acidic CO2 molecule, with the consequent gasification of the deposited carbon through the reverse of the Boudouard reaction according to the following equation [6,8].
CO2 + C ↔ 2CO         (ΔH 298K = 171 kJ/mol)
However, a good compromise between acidity and basicity must be considered, since a certain degree of support acidity is known to favor the CH4 decomposition and consequently its conversion [6]. The addition of rare earth metals, such as Ce, Zr and La, was demonstrated to improve the catalytic performance by preventing the formation of coke. Ce, with its high oxygen storage capacity (OSC) related to the rapid interchange between the Ce3+ and Ce4+ redox couple, would provide oxygen vacancies promoting CO2 adsorption and dissociation, increasing the activity of the catalyst and the gasification of the deposited carbon during the DRM reaction [16,17]. La also promoted the formation of small nickel particles and gasification of amorphous carbon deposits [18]. Zr, by a moderate interaction with the support, favored the formation of small nickel particles [19]. Other metal promoters, such as Sn, In and Au, have been investigated. Their role in nickel activity promotion was mostly related to their capability of alloying with nickel and therefore restraining the active crystallite size [20,21,22]. It is worth noting that discrepancies sometimes observed in literature on the role of each promoter are strictly related to the different experimental procedures used for the catalyst preparation and for the reactivity tests. Therefore, any general statement must be carefully considered. Besides single oxide carriers, mixed oxides of different structural and morphology properties have also been considered as DRM catalyst supports [15]. Catalysts derived from perovskite precursors and from layered double hydroxides (LDH) have been used in the DRM process [23,24]. These solid structures have been claimed to contribute to the formation of well-dispersed and active metal particles. In particular, the LDH compounds, also known as hydrotalcite-type (HT) compounds, are lamellar metal hydroxides consisting of main layers of divalent and trivalent metal cations and anion interlayers [24]. Recently, a special issue of Catalysts was entirely dedicated to the application of these compounds as supports or as catalysts for many different sustainable chemical reactions [25]. The suitability of LDH materials as such and as catalyst precursors lies in the active component dispersion, in their high degree of alkalinity and the possibility of forming oxides with homogeneous mesoporous textures. Moreover, the mixed oxides obtained after calcination are characterized by a large specific surface area due to the removal of anion groups such as CO32− and NO32−, and better resistance to sintering compared to other supported catalysts [26]. Indeed, the mixed oxides derived from calcined hydrotalcite-like precursors have been shown to stabilize the Ni-active centers [27]. The effect of the Ni/Al atomic ratio, of the calcination temperature and of different promoters on the DRM catalytic performance of HT-derived catalysts was investigated [28,29,30]. The first pioneer study on the hydrotalcite-derived NiAl oxides confirmed a higher methane-reforming activity of these oxides compared to traditional Ni catalysts [31]. Aiming to enhance the nickel reducibility and the stability of the HT-derived catalysts, the effects of different promoters, such as Ce, La and Rh, were investigated [29,30]. Ceria was investigated because of its peculiar oxygen mobility, favoring gasification of the deposited carbon [29]. Lanthania was found to form oxycarbonate La2O2CO3, which, reacting with adsorbed CHx species, contributed to carbon removal [30]. The same authors demonstrated that rhodium increased the reducibility of the nickel, maintaining the activity for longer time [30]. Based on the above grounds, the present study focused on the combined effect of CeO2 and Au on the DRM catalytic performance of NiAl HT-derived catalysts. Ceria was selected by virtue of its high oxygen-storage capacity and the related effects described above. On the other hand, gold was chosen because of its high electronegativity and alloying capability, favoring nickel reduction and restraining particle growth [22]. In fact, to the best of our knowledge, the combined effect of these two promoters on a nickel HT-derived catalyst has not yet been reported. Recently, some coauthors of the present study published a study on the catalytic performance of NiAl-HT catalysts modified by CeO2 and Au in the water–gas shift reaction (WGS) [32]. The innovative approach of the direct deposition of ceria on the NiAl-layered structure was successful for the attainment of well-dispersed CeO2 particles, which effectively constrained the particle size of the deposited gold. Based on these outcomes, a similar procedure was adopted to prepare NiAl hydrotalcite doped with ceria and gold. To avoid aggregation of ceria and to maintain the HT structure of the precursor, only 1 wt% Ce was considered. The DRM catalytic activity of the obtained samples after high-temperature calcination was evaluated and correlated with structural and electronic properties.

2. Results and Discussions

2.1. Catalytic Results

The NiAl-HT-derived catalysts tested in the DRM reaction are listed in Table 1 along with their elemental composition and their textural and structural parameters, which are discussed later.
The CH4 and CO2 conversions obtained with the different catalysts, plotted as a function of temperature in the temperature range of 450–700 °C, are shown in Figure 1 and Figure 2, respectively.
Although a higher reaction temperature favors the DRM reaction and inhibits side reactions, a relatively low temperature range up to 700 °C was selected, being more appealing for economic and technical reasons [4,24]. In the considered temperature range, the values of CH4 conversion were lower than the thermodynamic equilibrium conversions calculated considering the same conditions of the present study, i.e., at atmospheric pressure and with CH4/CO2 ratio of 1, as reported in literature [33]. Conversions of CH4 between 67% and 74% were obtained at 700 °C under the severe condition of high space velocity (600,000 mL g−1 h−1). These results compared quite positively with those from catalysts of similar composition tested at a much lower-space velocity [24].
A definite improvement in the CH4 conversion was observed for the Au- and Ce-promoted catalysts, particularly for Au/Ce/NiAl containing both promoter elements as compared to the NiAl. As observed with other nickel-supported catalysts, CO2 conversions larger than the CH4 conversions were obtained [34]. The differences, more evident at the lower temperatures, were due to the reverse water–gas shift reaction (RWGS)
CO2 + H2 ↔ CO + H2O      (ΔH 298K = + 41 kJ/mol)
competing with the DRM reaction [33,34]. However, the opposite effect of CH4 conversions larger than the CO2 conversions was reported with some Ce-promoted hydrotalcite-derived catalysts, attributed to the preferential occurrence of a CH4 decomposition reaction [24]. Different experimental conditions, such as gas feed dilution or space velocity, might explain the discrepancy between the results. The H2/CO molecular ratios, plotted as a function of temperature, are shown in the inset of Figure 1. The ratio decreased with temperature, ranging from values above 1.5 to a stable 0.7 for temperatures above 600 °C, passing through the stoichiometric value of 1 at 550 °C. The changes of the H2/CO ratio and the discrepancy between CH4 and CO2 conversions with temperature were intrinsically related to the side reactions taking place during the DRM test, particularly CH4 decomposition and RWGS, depending differently on the temperatures [15,33]. Generally, a H2/CO ratio less than one is due to an excess of CO, which may imply less coke deposition [9,35]. The largest H2/CO ratio obtained at low temperatures with the unpromoted NiAl catalyst was also in accordance with its superior activity in WGSR tested at temperatures up to 300 °C, as reported in the previous study [32].
At the temperature of 700 °C, the DRM activity of the catalysts was monitored as a function of time on stream (TOS) for 6 h. The obtained conversion plots for CH4 are displayed in Figure 3.
During the analysis, the H2/CO atomic ratio maintained a stable value of 0.7. With some differences, all the catalysts deactivated during the 6 h reaction. However, the best stability was exhibited by the Ce/NiAl, and the worst by the two gold-containing catalysts. The deleterious effect of gold relative to the catalyst deactivation was due to the surface migration of gold, covering part of the nickel-active sites. Indeed, the tendency of gold to segregate at the surface of a nickel catalyst upon exposure to the reducing environment at elevated temperature, as in the DRM process, was reported in a previous study on nickel catalysts supported on MgAl2O4 [22].
For better clarity, the catalytic results at 700 °C in terms of conversion and deactivation percentages, H2 yield, H2/CO ratio and percentage of carbon formation discussed below are summarized in Table 2. The hydrogen yield, calculated according to the equation given in the experimental section, ranged from 39% for the unpromoted NiAl to 44% for the gold- and cerium-promoted Au/Ce/NiAl catalyst.
The amount of carbon built up over the samples at the end of the catalytic test was determined by TGA analyses. The profiles are given in Figure 4.
At the beginning of the curves, a slight increase in the weight due to the oxidation of nickel was registered in a temperature range of 200–400 °C. Such an increase, equal for all samples, was hardly detectable, being much smaller compared to the large weight losses due to the carbon combustion. The carbon loss occurred within the wide temperature range of 500–750 °C, which corresponded to the reactivity of different types of carbon [14,27]. According to the profiles of Figure 4 and the calculated percentages of deposited carbon listed in Table 2, a similar amount of carbon formed upon DRM reaction with the unpromoted and promoted samples, except for the Ce-promoted one, which exhibited more carbon formation. Increased carbon deposition for Ce-doped samples was reported on similar types of catalysts and the weight loss observed at temperatures above 600 °C was attributed to graphitic carbon combustion [36]. The first derivative curves (DTG) allowed the pinpointing of the temperatures of the maximum loss, which are also reported in Table 2. Downward temperature shifts of 38 °C for the Ce/NiAl, 22 °C for the Au/NiAl and 52 °C for the Au/Ce/NiAl were observed in comparison to the unpromoted NiAl. The decrease in the temperature confirmed a positive effect exerted by both promoters, favoring the formation of an easier-to-remove carbon [36].
To investigate the activation of methane in the absence of CO2 by the different samples, the temperature-programmed surface reaction of methane (TPSR-CH4) method was carried out. The profiles of the gas evolution as a function of temperature are given in Figure 5.
The start (light off) of the curve with each sample is zoomed in the inset. It is worth noting the presence of small peaks at the beginning of each profile. These peaks corresponded to the conversion of methane reducing the nickel oxide still present on the catalyst surface even after the H2 pretreatment described in the experimental section. Consequent to this process, CO and CO2 evolution peaks were registered. The main CH4 conversion peaks were due to the decomposition of methane into solid carbon material C(s) and hydrogen gas H2(g), stoichiometrically evolved with respect to the decomposition reaction of methane [37]. The light-off temperatures, the conversion peak positions, and the decomposition temperature ranges, Tdec, summarized in Table 3, reflected the different activation capability of each sample.
As reported in the literature, the temperature window for TPSR-CH4 activity, corresponding to Tdec in Table 3, is indicative of the anti-coking ability in DRM [21,38]. Therefore, according to the Tdec values, the Au/Ce/NiAl catalyst should have the best anti-coking aptitude during the dry reforming process in accordance with the TGA results in Table 2 [38]. Simultaneously to the methane decomposition, there is a small but continuous production of CO arising from the reaction of the carbon, stemming from CH4 dissociation and diffusing through the nickel particles, with the oxygen of the sample lattice [37]. The methane conversion taking place above 800 °C was due to the homogeneous thermal decomposition of CH4 [39].
To investigate the carbon formed during the methane decomposition, TG analyses of the spent samples were performed. The results in terms of percentage of carbon and temperature of the DTG peak are listed also in Table 3. As expected, in the absence of CO2, much more carbon was accumulated over the catalysts during a shorter reaction time (about 120 min) compared to the DRM test. The Au/NiAl catalyst represented the only exception. In agreement with the TPSR curve, which showed CO continuously evolving up to 800 °C, the amount of accumulated carbon on this sample was indeed much lower than on the other samples. In all cases, the DTG peaks were in accord with the presence of graphitic carbon. The effect of the CO2 on carbon accumulation could be inferred by a comparison of the corresponding TGA data reported in Table 2 and Table 3. Indeed, the sample exhibiting a higher carbon removal temperature and higher carbon accumulation during the TPSR, i.e., Au/Ce/NiAl, was the one exhibiting a lower carbon removal temperature and a lower amount of accumulated carbon during the DRM test. Therefore, in this case, the presence of CO2 during the DRM test contributed to the formation of an easier-to-remove carbon, favored by the close interaction between ceria and gold, enhancing the oxygen mobility of the support and the adsorption of CO2 [40]. The Ce/NiAl maintained the same temperature for the carbon removal after TPSR and DRM, confirming the main role of ceria in favoring the formation of a more reactive carbon, regardless of the presence of CO2. For the other two samples, NiAl and Au/NiAl, which were characterized by lower carbon removal temperatures in TPSR compared to DRM, the presence of CO2 would not contribute to the formation of a more active carbon.

2.2. Characterization

2.2.1. BET and XRD Analyses

In Table 1, along with the elemental composition, the BET-specific surface areas and the pore volumes of the calcined HT-derived samples are listed. The particle sizes of NiO, Au and Ni as derived from the XRD analyses to be discussed later, are also included. The N2 adsorption–desorption isotherms are plotted in Figure 6, with the pore size distributions given in the inset.
According to the IUPAC classification, the isotherms were of type IV, characteristic of mesoporous materials, with clear hysteresis loops of H1 type [41]. With respect to the surface area of the parental hydrotalcite samples reported in the previous study, a ~70% increase in the specific surface area was observed in all the HT-derived samples [32]. The increase in the surface area was in fact expected, due to the removal of interlayer water and decomposition of the interlayer anions with the creation of additional porosity [24]. The surface areas of the gold-containing samples were slightly lower.
The structure of the parental NiAl hydrotalcites was confirmed by the X-ray diffraction patterns displayed in Figure 7, where all the reflections typical of the layered structure of takovite mineral and those of metallic gold are indicated [32].
The X-ray diffraction patterns of the catalysts, which were calcined at 600 °C, changed considerably. The diffractograms of these catalysts before and after the catalytic tests are given in Figure 8. As expected, after calcination, the takovite structure disappeared, and instead each sample before reaction exhibited a pattern characterized by the reflections at 2θ = 37.2°, 43.3°, 62.9°, 75.4° and 79.4° indexed as (111), (200), (220), (311) and (222) crystalline planes of cubic NiO (JCDD-PDF File 00-047-1049) [42]. No peaks related to the spinel NiAl2O4 were detected. Due to the low amount of ceria, no ceria-related reflections were present in the patterns of the samples containing 1 wt% Ce. By contrast, the patterns of the samples with 3 wt% Au contained peaks at 2θ = 38.2° and 64.6° very close to the NiO reflections, and a peak at 2θ = 77.5°, attributed respectively to the (111), (220) and (311) crystalline planes of the cubic Au (JCDD-PDF File 00-004-0784). Deconvolution of the NiO (111) and Au (111) peaks is shown as a zoomed-in region in the insets of Figure 8.
The diffractograms of the spent catalysts still contained some of the strongest NiO reflections, but clearly indicated the presence of metallic nickel characterized by (111), (200) and (220) reflections at 2θ = 44.5°, 51.8° and 76.4°, respectively (JCDD-PDF File 04-0850). Relative to the gold-containing catalysts, after the catalytic test, no gold reflections were discernible in the corresponding patterns. This could be attributed to the surface migration of gold induced by the DRM process, explaining the deactivation of the gold-containing catalysts, as discussed ahead in relation to Figure 3 [22]. From the Scherrer analyses of the NiO (200), Ni (200) and Au (111) reflections, the crystallite sizes of nickel oxide, metallic nickel and metallic gold were estimated [43]. According to the obtained values reported in Table 1, the fresh samples contained NiO particles of ~4 nm. In the gold-containing samples, gold crystallites of ~20 nm sizes were estimated. After reaction, NiO particles of larger size and Ni particles between 16 and 25 nm in size were obtained. It is worth noting that the presence of ceria limited the growth of the NiO and the Ni crystallites, whereas gold alone did not significantly affect them. A similar size-constraining effect on nickel particles exerted by ceria was reported in a WGSR investigation of similar catalysts [41]. The diffractograms of all the spent catalysts contained a peak at 2θ~26° corresponding to graphitic carbon in accord with the TGA analyses.

2.2.2. TPR Analyses

Reforming of methane requires active sites consisting of metallic nickel aggregates, which should be readily formed and maintained during the reaction. The effect of the promoters on the reducibility of the catalysts was then investigated through H2-TPR analyses. The TPR profiles of the calcined catalysts are shown in Figure 9.
All the profiles exhibited one broad asymmetric peak between 380 °C–800 °C in accord with what is reported in literature for similar calcined NiAl HT samples [36]. The peak was attributed to the reduction of NiO in strong interaction with the alumina, whereas the asymmetry on the low-temperature side of the peak was due to the reduction of NiO particles of different sizes and differently interacting with the supports. In particular, the shoulder observed at 400 °C in the profile of the NiAl could be due to larger NiO particles less interacting with the alumina. The position of the main peak shifted from a higher temperature of 660 °C, exhibited by NiAl and CeNiAl, to 640 °C, exhibited by the Au-containing catalysts. The downward temperature shift observed in the gold-containing samples, excluding any significant particle size effect (Table 1), could be attributed to the tendency of gold to intimately interact with nickel and therefore enhancing the nickel reducibility [22]. The hydrogen consumption was different for each sample, and in all cases a little lower than the uptake of 224 mlH2/gcat expected from the Ni loading. With respect to the hydrogen consumption, the following order was obtained: NiAl (179 mlH2/gcat) < Au/NiAl = Ce/NiAl (193 mlH2/gcat) < Au/Ce/NiAl (210 mlH2/gcat). The smaller amount of hydrogen consumption, especially by the unpromoted sample, suggested the presence of some difficult-to-reduce nickel species.

2.2.3. XPS Analyses

XPS measurements of the fresh (calcined) catalysts were carried out to investigate the distribution and the chemical state of each catalyst component. The Ni 2p spectra of all the catalysts were typical of Ni2+, characterized by the main Ni 2p3/2 and Ni 2p1/2 spin-orbit components, each accompanied by the shakeup features [44]. As an example, the Ni 2p spectra of the Au/Ce/NiAl catalyst before (fresh) and after the DRM reaction (aged) are given in Figure 10.
The main Ni 2p peaks were fitted with the two spin-orbit components Ni 2p3/2 and Ni 2p1/2 centered at 855.7 ± 0.1 eV and 873. 6 ± 0.1 eV, respectively. Both peaks were accompanied by the shakeup satellites about 6 eV apart from the main photoelectron peaks. The binding energy and the presence of these satellites were typical of oxidized nickel as Ni++. The Ni++ in pure NiO is characterized by a Ni 2p3/2 binding energy of 854.5 eV, more than 1 eV smaller than the value obtained in the present case, which was intermediate between bulk NiO and NiAl2O4, suggesting the formation of a thin surface layer of spinel not detectable by XRD analyses. Both spectra in the figure contained extra features attributed to ceria, however the weak signals did not allow a proper fitting. After catalytic test, the intensity of the Ni 2p spectrum decreased significantly, as shown by the decreased signal-to-noise ratio. Within the uncertainty of the fitting procedure due to the poor quality of the spectrum, it was possible to fit small components attributed to metallic nickel characterized by Ni 2p3/2 binding energy at ~851.8 eV.
In Figure 11, the spectral regions of the fitted Ni 3p, Al 2p and Au 4f are displayed. The binding energy of Al 2p was equal to 74.5 eV ± 0.1 eV, which is typical of aluminum oxide. The binding energy of Au 4f7/2 of 84.0 eV obtained in both gold-containing samples was typical of metallic gold [22]. The extra peak overlapping with the Au 4f5/2 component was due to an Al 2p plasmon satellite. The fitting allowed a proper quantification of the elements in terms of atomic ratios. The results for all the fresh catalysts in terms of binding energies and atomic ratios are summarized in Table 4.
From the comparison of the XPS-derived surface atomic ratios Ni/Al XPS and Au/Al XPS with the corresponding bulk ratios, a surface depletion of nickel and gold was evident. This observation could have been a direct consequence of the migration of both metals inside the hydrotalcite structure during the precursor’s preparation procedure and during the calcination. It is worth noting the surface enhancement of nickel in the ceria- and gold-promoted catalysts, which explains their better initial activity with respect to the unpromoted one. Moreover, an enrichment of the surface gold concentration due to the presence of ceria was observed in the Au/Ce/NiAl catalyst. As shown in Figure 12 for the Au/Ce/NiAl sample, in accordance with the XRD patterns and the TGA analyses, a strong and asymmetric C 1s peak at ~284 eV typical of graphitic carbon was observed in all the C 1s regions of the spent catalysts [44]. Unfortunately, after the catalytic test, the extensive deposition of carbon over the samples prevented the XP analyses of the Au 4f peaks.

3. Experimental

3.1. Synthesis of Materials

The materials were prepared according to the procedure already described [32]. In brief, the carbonate form of NiAl-HT with a Ni2+/Al3+ molar ratio of 2.5 was obtained by coprecipitating at 80 °C and at constant pH = 8 the mixed aqueous solution of 0.5M Ni(NO3)2 × 6H2O, and Al(NO3)3 × 9H2O with a 0.9M Na2CO3 as precipitant agent. The obtained slurry was aged for 60 min under controlled stirring conditions, filtered off and washed with hot distilled water. Then, the precipitate was dried at 105 °C for 20 h.
The CeO2-modified NiAl-HT sample, containing 1wt.% CeO2, was prepared by the direct deposition of ceria over the NiAl-HT suspended in distilled water, through precipitation of Ce3+ ions from Ce(NO3)3 × 6H2O with 1M NaOH. The washed sample (pH of filtrate ~6–7) was dried at 105 °C.
Gold-containing samples were obtained by deposition–precipitation of gold over NiAl-HT and Ce/NiAl-HT. Portions of the HT powders were homogeneously suspended in distilled water with the help of ultrasound. Deposition of gold (3 wt.%) was performed by simultaneous addition of aqueous solutions of 0.06 M HAuCl4 x 3H2O and 0.2 M Na2CO3 into the reactor at 60 °C and pH = 7.0. After aging under the same conditions, for 60 min, the samples were filtered and carefully washed with distilled water until the removal of the Cl ions.
The elemental composition of each sample, as obtained from ICP measurements, is listed in Table 1. The molar ratio of Ni2+/Al3+ of the promoted and unpromoted samples satisfied the range of 0.2 < M2+/(M2+ + M3+) < 0.33 required for the formation of pure hydrotalcites [28]. Formation of the layered double-oxide structure NiAl-HT was confirmed by the X-ray diffraction pattern typical of the mineral takovite, as previously reported [32]. For the catalytic application, all the samples were calcined at 600 °C for 2 h with a heating rate of 5 °C/min, and they were designated as NiAl, Ce/NiAl, Au/NiAl and Au/Ce/NiAl.

3.2. Characterization

Specific surface areas and pore volumes of the samples were measured at −196 °C with a nitrogen sorption technique using the ASAP 2020 equipment (Micromeritics, Norcross, GA, USA). Before the measurements, the powder samples were degassed at 250 °C for 2 h. The specific surface area was calculated via the Brunauer-Emmett-Teller (BET) method in the standard pressure range 0.05–0.3 P/P0. The pore volume and pore size distribution were obtained by analysis of the desorption branch, using the Barrett, Joyner and Halenda (BJH) calculation method [41].
X-ray diffraction (XRD) patterns were recorded with a D5000 diffractometer (Bruker AXS, Karlsruhe, Germany) employing Ni-filtered Cu Kα radiation in the 2θ range between 20° and 90° with 0.05° step size. The assignment of the crystalline phases was based on the JPDS powder diffraction file cards [42]. Crystallite sizes were calculated from the Scherrer analyses of selected peaks [43].
H2 temperature-programmed reduction (TPR) measurements were carried out with Autochem 2910 HP system (Micromeritics, Norcross, GA, USA) equipped with a TCD detector. Before analysis, 0.1 g of the sample was treated with a mixture of 5 % O2/He (v/v, 50 mL/min) while heating up (10 °C/min) to 400 °C, and was left at this temperature for 30 min. After cooling to room temperature, the gas mixture of 5 % H2/Ar (v/v, 30 mL/min) was introduced into the sample tube. The hydrogen consumption was measured as a function of the temperature, increased up to 1000 °C at a rate of 10 °C/min.
The X-ray photoelectron spectroscopy (XPS) analyses were carried out with a VG Microtech ESCA 3000 Multilab (VG Scientific, Sussex, UK), equipped with a dual Mg/Al anode. Al Kα radiation was used as an excitation source (1486.6 eV). All the binding energies were referred to the C 1s energy, previously calibrated at 285.1 eV, arising from adventitious carbon. Qualitative and quantitative analyses of the peaks were performed using CasaXPS software (version 2.3.17, Casa Software Ltd. Wilmslow, Cheshire, UK, 2009). A precision of ± 0.15 eV on the binding energy values and of ± 10%. on the atomic percentages was considered.
The thermogravimetric analyses (TGA) of the samples after the catalytic reactions were performed using the TGA 1 Star System (Mettler Toledo, Scwerzenbach, Switzerland). About 10 mg of the sample were heated from room temperature to 100 °C, left at this temperature for 1 h and then heated to 1100 °C at the rate of 10 °C/min under flowing air at 30 mL/min.

3.3. Catalytic Measurements

The DRM reaction was carried out in a fixed-bed quartz reactor with an inner diameter of 12 mm. The reaction was conducted under atmospheric pressure in the temperature range of 450 °C–700 °C, with increasing temperature steps of 50 °C. Typically, 10 mg of catalyst powder (sieved fraction between 180 and 250 mm) diluted 1:2 with inert SiC were placed inside the reactor. Then, prior to the catalytic testing, to clean up the catalyst surface and make consistent comparisons between the different catalysts, the samples were treated with a mixture of 5% O2/He (v/v, 50 mL/min) at 350 °C for 1/2 h. After cooling down to room temperature, the sample was heated up to 750 °C at a rate of 10 °C/min in a 5% H2/He (v/v, 30 mL/min), and held at this temperature for 1 h. The feed gas composed of 10% of CH4, 10% of CO2 and He as balance was introduced into the reactor with a flow rate of 100 mL/min (STP), equivalent to a gas hourly space velocity (GHSV) of 600,000 mL g−1 h−1. The inlet and outlet gas compositions were analyzed online using the Agilent 7890B Gas Chromatograph (Agilent Technology, Santa Clara, CA, USA) equipped with a DB-1 capillary column and a molecular sieve, to follow the evolution of CH4, CO, CO2 and H2, using FID and TCD detectors. Water was condensed at the outlet of the reactor. CH4 and CO2 conversion, XCH4 and XCO2 respectively, and hydrogen yield YH2 (%) were calculated according to the following equations, where CH4in, CO2in and CH4out, CO2out, H2out refer to the concentration (ppm) of the species entering and exiting, respectively, the catalytic reactor.
XCH4 = 100 × (CH4in–CH4out)/CH4in
XCO2 =100 × (CO2in–CO2out)/ CO2in
YH2 (%) = 100 × H2out / (2 × CH4in)
The activation of methane by the catalyst surface was investigated by temperature-programmed surface reaction of methane (TPSR-CH4), using Autochem 2950 HP system (Micromeritics, Norcross, GA, USA). For this reaction, 10 mg of catalyst mixed with 1 g of carborundum to disperse the active phase were placed in a U-shaped quartz tube with quartz wool on each side of the bed. Before the reaction, the sample underwent the same oxidation and reduction treatment used for the DRM reaction, as described above. After pretreatment, the sample was cooled down to room temperature in He. Then, a mixture of 5% CH4/He (v/v, 30 mL/min) was introduced into the reactor, and the temperature was increased up to 1100 °C, using a heating rate of 10 °C/min. The evolutions of CH4, CO2 and CO were registered using UV/IR analyzers (ABB S.p.a. Milano, Italy).

4. Conclusions

Ni containing hydrotalcite-derived catalysts with a molar ratio Ni++/Al+++ ratio = 2.5 were used in the dry reforming of methane. The samples were modified by the addition of cerium and gold to improve the resistance of the catalyst toward carbon formation and toward sintering of the nickel. The unpromoted and promoted catalysts were tested at a relatively low temperature range between 450 °C and 700 °C. Values of CH4 conversion between 67% and 74% obtained at 700 °C under the harsh experimental condition of a high space velocity (600,000 mL g−1 h−1) were indicative of active catalysts. During the temperature-dependent DRM test, definite improvement in the CH4 conversion of the promoted catalysts compared to the unpromoted NiAl was observed, particularly of the Au/Ce/NiAl containing both promoter elements. The H2/CO molecular ratio decreased with temperature, ranging from values above 1.5 to a stable 0.7 for temperatures above 600 °C. The changes in the H2/CO ratio along with the discrepancies between the CH4 and CO2 conversions were related to side reactions taking place during the DRM test. During 6 h of time on stream at 700 °C, all catalysts deactivated, the Ce/NiAl to a lesser degree and the two gold-containing samples to a greater degree. Accumulation of graphitic carbon, observed in all the spent catalysts, contributed to the deactivation.
The present study confirmed that the formation of carbon, acknowledged as one of the main causes of the nickel catalyst deactivation, was not always deleterious. The presence of ceria, which in the present case did not prevent the formation of carbon, contributed to an enhancement of the catalytic activity stability due to a particle-size-constraining effect. The gold-containing catalysts, in spite of similar or even lower amounts of registered carbon after DRM reaction, deactivated the most, due to the gold surface segregation driven by the reaction conditions. Moreover, a lower carbon-removal temperature observed in the TGA analyses of the spent promoted catalysts must be considered beneficial from the perspective of the catalyst regeneration procedure. Particularly in the case of the Au/Ce/NiAl catalyst, the close interaction between gold and ceria enhancing the oxygen mobility of the support and the adsorption of CO2 would favor the removal of the formed carbon. Therefore, any preparation method that allows proper anchoring of the gold in the catalyst structure, inhibiting its surface segregation, could be promising for obtaining a catalyst with better DRM performance.

Author Contributions

Conceptualization, A.M.V. and T.T.; methodology, V.L.P. and A.M.V.; catalytic activity, V.L.P. and G.P.; characterization and TPR/TPO measurements, L.F.L. and G.P.; XPS analyses, A.M.V. and V.L.P.; synthesis, M.G., D.N. and T.T.; writing and editing, A.M.V. and G.P. All the authors contributed to the data discussion and manuscript preparation. All authors have read and agreed to the published version of the manuscript.

Funding

The authors kindly acknowledge financial support from the Bilateral Project between the CNR and the Bulgarian Academy of Science (BAS). Authors affiliated with the ISMN-CNR acknowledge funding from the National Project NAUSICA (ARS01_00334-CUP B45F21000680005) and from the PNRR—Missione 2 “Rivoluzione Verde e Transizione Ecologica”—Componente 2 “Energia Rinnovabile, Idrogeno, Rete e Mobilità Sostenibile”—Investimento 3.5 “Ricerca e Sviluppo sull’Idrogeno”—Decreto MITE n. 545 del 23/12/2021(AdC ENEA-CNR) (CUP B93C22000630006). Authors affiliated with the Institute of Catalysis-BAS gratefully acknowledge the support of the European Regional Development Fund within the Operational Programme “Science and Education for Smart Growth 2014–2020”, Project CoE “National Center of Mechatronics and Clean Technologies “BG05M2OP001-1.001-0008”.

Data Availability Statement

The data presented in this study are available in the article.

Acknowledgments

Administrative support by Giovanna Bellanti, Giuseppe Napoli and Salvatore Romeo is acknowledged. The authors are thankful to Francesco Giordano for performing XRD analyses and to Nunzio Galli for performing surface area measurements.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Wittich, K.; Krämer, M.; Bottke, N.; Schunk, S.A. Catalytic Dry Reforming of Methane: Insights from Model Systems. ChemCatChem 2020, 12, 2130–2147. [Google Scholar] [CrossRef]
  2. Seo, H.O. Recent Scientific Progress on Developing Supported Ni catalysts for Dry (CO2) Reforming of Methane. Catalysts 2018, 8, 11. [Google Scholar] [CrossRef] [Green Version]
  3. Cai, X.; Hu, Y.H. Advances in catalytic conversion of methane and carbon dioxide to highly valuable products. Energy Sci. Eng. 2019, 7, 4–29. [Google Scholar] [CrossRef] [Green Version]
  4. Gao, X.; Ge, Z.; Zhu, G.; Wang, Z.; Ashok, J.; Kawi, S. Anti-Coking and Anti-Sintering Ni/Al2O3 Catalysts in the dry reforming of methane: Recent Progress and Prospects. Catalysts 2021, 11, 1003. [Google Scholar] [CrossRef]
  5. Sandoval-Diaz, L.E.; Schlögl, R.; Lunkenbein, T. Quo Vadis Dry Reforming of Methane? A Review on Its Chemical, Environmental, and Industrial Prospects. Catalysts 2022, 12, 465. [Google Scholar] [CrossRef]
  6. Touma, J.G.; Tarboush, B.A.; Zeaiter, J.; Ahmad, M.N. Catalyst design for dry reforming of methane: Analysis Reviews. Renew. Sustain. Energy Rev. 2018, 82, 2570–2583. [Google Scholar]
  7. Parsapur, R.K.; Chatterjee, S.; Huang, K.-W. The Insignificant Role of Dry Reforming of Methane in CO2 Emission Relief. ACS Energy Lett. 2020, 5, 2881–2885. [Google Scholar] [CrossRef]
  8. Arora, S.; Prasad, R. An overview on dry reforming of methane: Strategies to reduce carbonaceous deactivation of catalysts. RSC Adv. 2016, 6, 108688. [Google Scholar] [CrossRef]
  9. Muraza, O.; Galadima, A. A review on coke management during dry reforming of methane. Int. J. Energy Res. 2015, 39, 1196–1216. [Google Scholar] [CrossRef]
  10. Bermúdez, J.; Fidalgo, B.; Arenillas, A.; Menéndez, J. Dry reforming of coke oven gases over activated carbon to produce syngas for methanol synthesis. Fuel 2010, 89, 2897–2902. [Google Scholar] [CrossRef] [Green Version]
  11. Han, J.W.; Park, J.S.; Choi, M.S.; Lee, H. Uncoupling the size and support effects of Ni catalysts for dry reforming of methane. Appl. Catal. B 2017, 203, 625–632. [Google Scholar] [CrossRef]
  12. Akri, M.; Zhao, S.; Li, X.; Zang, K.; Lee, A.F.; Isaacs, M.A.; Xi, W.; Gangarajula, Y.; Luo, J.; Ren, Y.; et al. Atomically dispersed nickel as coke-resistant active sites for methane dry reforming. Nat. Commun. 2019, 10, 5181. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Vogt, C.; Kranenborg, J.; Monti, M.; Weckhuysen, B.M. Structure Sensitivity in Steam and Dry Reforming over Nickel: Activity and Carbon formation. ACS Catal. 2020, 10, 1428–1438. [Google Scholar] [CrossRef] [Green Version]
  14. Xu, Y.; Du, X.-H.; Li, J.; Wang, P.; Zhu, J.; Ge, F.-J.; Zhou, J.; Song, M.; Zhu, W.-Y. A comparison of Al2O3 and SiO2 supported Ni-based catalysts on their performance for the dry reforming of methane. J. Fuel Chem. Technol. 2019, 47, 199–208. [Google Scholar] [CrossRef]
  15. Singh, R.; Dhir, A.; Mohapatra, S.K.; Mahia, S.K. Dry reforming of methane using various catalysts in the process: Review. Biomass Convers. Biorefinery 2020, 10, 567–587. [Google Scholar] [CrossRef]
  16. Emamdoust, A.; La Parola, V.; Pantaleo, G.; Testa, M.L.; Farjami Shayesteh, S.; Venezia, A.M. Partial oxidation of methane over SiO2 supported Ni and NiCe catalysts. J. Energy Chem. 2020, 47, 1–9. [Google Scholar] [CrossRef]
  17. Marinho, A.L.A.; Toniolo, F.S.; Noronha, F.B.; Epron, F.; Duprez, D.; Bion, N. Highly active and stable Ni dispersed on mesoporous CeO2-Al2O3 catalysts for production of syngas by dry reforming of methane. Appl. Catal. B 2021, 10, 1428–1438. [Google Scholar] [CrossRef]
  18. Yang, R.; Xing, C.; Lv, C.; Shi, L.; Tsubaki, N. Promotional effect of La2O3 and CeO2 on Ni/gAl2O3 catalysts for CO2 reforming of CH4. Appl. Catal. A 2010, 385, 92–100. [Google Scholar] [CrossRef]
  19. Wang, Y.; Yao, L.; Wang, Y.; Wang, S.; Zhao, Q.; Mao, D.; Hu, C. Low -Temperature Catalytic CO2 Dry Reforming of Methane on Ni-Si/ZrO2 catalyst. ACS Catal. 2018, 8, 6495–6506. [Google Scholar] [CrossRef]
  20. Pantaleo, G.; La Parola, V.; Testa, M.L.; Venezia, A.M. CO2 reforming of CH4 over SiO2-Supported Ni Catalyst: Effect of Sn as Support and metal Promoter. Ind. Eng. Chem. Res. 2021, 60, 18684–18694. [Google Scholar] [CrossRef]
  21. Horváth, A.; Beck, A.; Maróti, B.; Sáfrán, G.; Pantaleo, G.; Liotta, L.F.; Venezia, A.M.; La Parola, V. Strong impact of indium promoter on Ni/Al2O3 and Ni/CeO2-Al2O3 catalysts used in dry reforming of methane. Appl. Catal. A 2021, 621, 18174. [Google Scholar] [CrossRef]
  22. Horváth, A.; Guczi, L.; Kocsonya, A.; Sáfrán, G.; La Parola, V.; Liotta, L.F.; Pantaleo, G.; Venezia, A.M. Sol-derived AuNi/MgAl2O4 catalysts: Formation, structure and activity in dry reforming of methane. Appl. Catal. A 2013, 468, 250–259. [Google Scholar] [CrossRef] [Green Version]
  23. Bhatar, S.; Abedin, M.A.; Kanitkar, S.; Spivey, J.J. A review on dry reforming of methane over perovskite derived catalysts. Catal. Today 2021, 365, 2–23. [Google Scholar] [CrossRef]
  24. Debęk, R.; Motak, M.; Grzybek, T.; Galvez, M.E.; Da Costa, P. A short review on the Catalytic Activity of Hydrotalcite- Derived Materials for Dry Reforming of Methane. Catalysts 2017, 7, 32. [Google Scholar] [CrossRef] [Green Version]
  25. Marcu, I.-C.; Pavel, O.D. Layered Double Hydroxide-Based Catalytic Materials for Sustainable Processes. Catalysts 2022, 12, 816. [Google Scholar] [CrossRef]
  26. Abdelsadek, Z.; Holgado, J.P.; Halliche, D.; Caballero, A.; Cherifi, O.; Cortes, S.G.; Masset, P.J. Examination of the deactivation Cycle of NiAl and NiMgAl- Hydrotalcite derived catalysts in the Dry Reforming of Methane. Catal. Lett. 2021, 151, 2596–2715. [Google Scholar] [CrossRef]
  27. Dai, H.; Zhu, Y.; Xiong, S.; Xiao, X.; Huang, L.; Deng, J.; Zhou, C. Dry Reforming of Methane over Ni/MgO@Al Catalysts with Unique Feature of Sandwich Structure. Chem. Sel. 2021, 6, e202102788. [Google Scholar] [CrossRef]
  28. Touahra, F.; Sehailia, M.; Ketir, W.; Bachari, K.; Chebout, R.; Tarri, M.; Cherifi, O.; Halliche, D. Effect of the Ni/Al ratio of Hydrotalcite-type catalysts on their performance in the methane dry reforming process. Appl. Petrochem. Res. 2016, 6, 1–13. [Google Scholar] [CrossRef] [Green Version]
  29. Daza, C.E.; Gallego, J.; Mondragon, F.; Moreno, S.; Molina, R. High stability of Ce-promoted Ni/Mg-Al catalysts derived from Hydrotalcites in dry reforming of methane. Fuel 2010, 89, 592–603. [Google Scholar] [CrossRef]
  30. Lucredio, A.F.; Assaf, J.M.; Assaf, E.M. Reforming of a model sulfur-free biogas on Ni catalysts supported on Mg(Al)O derived from hydrotalcite precursors: Effect of La and Rh addition. Biomass Bioenergy 2014, 60, 8–17. [Google Scholar] [CrossRef]
  31. Bhattacharyya, A.; Chang, V.W.; Schumaker, D.J. CO2 reforming of methane to syngas; I: Evaluation of Hydrotalcite clay-derived catalysts. Appl. Clay Sci. 1998, 13, 317–328. [Google Scholar] [CrossRef]
  32. Gabrovska, M.; Ivanov, I.; Nikolova, D.; Krstic, J.; Venezia, A.M.; Crisan, D.; Crisan, M.; Tenchev, K.; Idakiev, V.; Tabakova, T. Improved Water–Gas Shift Performance of Au/NiAl LDHs Nanostructured Catalysts via CeO2 Addition. Nanomaterials 2021, 11, 366. [Google Scholar] [CrossRef] [PubMed]
  33. Ali, S.; Khader, M.M.; Almarri, M.; Abdelmoneim, A.G. Ni-Based Catalysts for the dry reforming of methane. Catal. Today 2020, 343, 26–37. [Google Scholar] [CrossRef]
  34. Nikoo, M.K.; Amin, N.A.S. Thermodynamic analysis of carbon dioxide reforming in view of solid carbon formation. Fuel Process. Technol. 2011, 92, 678–691. [Google Scholar] [CrossRef] [Green Version]
  35. Wang, Y.; Li, L.; Cui, C.; Da Costa, P.; Hu, C. The effect of absorbed oxygen species on carbon-resistance of Ni-Zr catalyst modified by Al and Mn for dry reforming of methane. Catal. Today 2022, 384–386, 257–264. [Google Scholar]
  36. Debęk, R.; Motak, M.; Grzybek, T.; Galvez, M.E.; Da Costa, P. Catalytic activity of hydrotalcite-derived catalysts in the dry reforming of methane: The effect of Ce promotion and feed gas composition. React. Kinet. Mech. Catal. 2017, 121, 185–208. [Google Scholar] [CrossRef] [Green Version]
  37. Guil-Lopez, R.; La Parola, V.; Pena, M.A.; Fierro, J.L.G. Evolution of the Ni-active centres into ex Hydrotalcite oxide catalysts during the COx-free hydrogen production by methane decomposition. Int. J. Hydrogen Energy 2012, 37, 7042–7705. [Google Scholar] [CrossRef]
  38. Zhang, M.; Zhang, J.; Wu, Y.; Pan, J.; Zhang, Q.; Tan, Y.; Han, Y. Insight into the effects of the oxygen species over Ni/ZrO2 catalyst surface on methane reforming with carbon dioxide. Appl. Catal. B 2019, 244, 427–437. [Google Scholar] [CrossRef]
  39. Pantaleo, G.; La Parola, V.; Deganello, F.; Singha, R.K.; Bal, R.; Venezia, A.M. Ni/CeO2 catalysts for methane partial oxidation: Synthesis, driven structural and catalytic effects. Appl. Catal. B 2016, 189, 233–241. [Google Scholar] [CrossRef]
  40. Tabakova, T.; Gabrovska, M.; Nikolova, D.; Ivanov, I.; Venezia, A.M.; Tenchev, K. Exploring the role of promoters (Au, Cu and Re) in the performance of Ni-Al layered double hydroxides for water-gas shift reaction. Int. J. Hydrog. Energy 2022, in press. [Google Scholar] [CrossRef]
  41. Thommes, M.; Kaneko, K.; Neimark, A.V.; Olivier, J.P.; Rodriguez-Reinoso, F.; Rouqueroi, J.; Sing, K.S.W. Physisorption of Gases, with Special Reference to the Evaluation of Surface Area and Pore Size Distribution (IUPAC Technical Report). Pure App. Chem. 2015, 87, 1051–1069. [Google Scholar] [CrossRef] [Green Version]
  42. Inorganic Crystal Structure Database (ICSD); FIZ Karlsruhe, GmbH: Eggenstein-Leopoldshafen, Germany, 2014.
  43. Klug, H.P.; Alexander, L.E. X-Ray Diffraction Procedures for Polycrystalline and Amorphous Materials, 2nd ed.; John Wiley and Sons: New York, NY, USA, 1974. [Google Scholar]
  44. La Parola, V.; Liotta, L.F.; Pantaleo, G.; Testa, M.L.; Venezia, A.M. CO2 reforming of CH4 over Ni supported on SiO2 modified by TiO2 and ZrO2: Effect of the support synthesis procedure. Appl. Catal. A 2022, 642, 118704. [Google Scholar] [CrossRef]
Figure 1. CH4 conversion of NiAl-HT-derived catalysts, with H2/CO molecular ratio plotted in the inset.
Figure 1. CH4 conversion of NiAl-HT-derived catalysts, with H2/CO molecular ratio plotted in the inset.
Catalysts 13 00606 g001
Figure 2. CO2 Conversion of NiAl-HT-derived catalysts.
Figure 2. CO2 Conversion of NiAl-HT-derived catalysts.
Catalysts 13 00606 g002
Figure 3. Deactivation test at 700 °C of NiAl-HT-derived catalysts.
Figure 3. Deactivation test at 700 °C of NiAl-HT-derived catalysts.
Catalysts 13 00606 g003
Figure 4. TGA profiles and corresponding DTG profiles of spent NiAl-HT-derived catalysts.
Figure 4. TGA profiles and corresponding DTG profiles of spent NiAl-HT-derived catalysts.
Catalysts 13 00606 g004
Figure 5. Gas evolution during TPSR-CH4 over NiAl-HT-derived catalysts.
Figure 5. Gas evolution during TPSR-CH4 over NiAl-HT-derived catalysts.
Catalysts 13 00606 g005
Figure 6. N2 adsorption–desorption isotherm plots with pore size distributions in the inset of the NiAl-HT-derived catalysts.
Figure 6. N2 adsorption–desorption isotherm plots with pore size distributions in the inset of the NiAl-HT-derived catalysts.
Catalysts 13 00606 g006
Figure 7. XRD patterns of precursor NiAl-HT samples. The diffraction lines of metallic gold are marked by asterisks.
Figure 7. XRD patterns of precursor NiAl-HT samples. The diffraction lines of metallic gold are marked by asterisks.
Catalysts 13 00606 g007
Figure 8. X-ray diffraction patterns of (a) NiAl, (b)Au/NiAl, (c) Ce/NiAl and (d) Au/Ce/NiAl HT-derived catalysts before and after DRM reaction.
Figure 8. X-ray diffraction patterns of (a) NiAl, (b)Au/NiAl, (c) Ce/NiAl and (d) Au/Ce/NiAl HT-derived catalysts before and after DRM reaction.
Catalysts 13 00606 g008
Figure 9. TPR profiles of NiAl-HT-derived catalysts.
Figure 9. TPR profiles of NiAl-HT-derived catalysts.
Catalysts 13 00606 g009
Figure 10. Ni2p XP spectra of fresh and aged AuCe/NiAl catalysts.
Figure 10. Ni2p XP spectra of fresh and aged AuCe/NiAl catalysts.
Catalysts 13 00606 g010
Figure 11. Background subtracted X-ray photoelectron regions of Ni 3p, Al2p and Au4f spectra for the fresh gold-containing samples.
Figure 11. Background subtracted X-ray photoelectron regions of Ni 3p, Al2p and Au4f spectra for the fresh gold-containing samples.
Catalysts 13 00606 g011
Figure 12. C 1s XP spectra of Au/Ce/NiAl; (a) before and (b) after DRM.
Figure 12. C 1s XP spectra of Au/Ce/NiAl; (a) before and (b) after DRM.
Catalysts 13 00606 g012
Table 1. Elemental composition and textural properties of the HT-derived samples.
Table 1. Elemental composition and textural properties of the HT-derived samples.
SamplesNi
wt%
Al
wt%
Ce
wt%
Au
wt%
S
(m2g−1)
Vtot
(cm3g−1)
dNiO *
(nm)
dAu
(nm)
dNi **
(nm)
NiAl61.711.3 2261.005(21) 25
Au/NiAl59.911.0 32150.974(21)2223
CeNiAl61.112.30.8 2240.924(12) 16
Au/CeNiAl59.310.90.832150.974(10)1820
* The values in parentheses refer to the samples after the DRM catalytic test. ** These values refer to the samples after the DRM catalytic test.
Table 2. Catalytic results of DRM at 700 °C for the HT-derived samples, with corresponding TGA-derived carbon information.
Table 2. Catalytic results of DRM at 700 °C for the HT-derived samples, with corresponding TGA-derived carbon information.
Samples CH4 Conversion * (%)CO2 Conversion * (%)H2 Yield * (%)H2/CO%C **TC *** (°C)
NiAl67 (19)83 (07)39(22)0.7244742
Au/NiAl72 (28)86 (08)43(26)0.7244720
Ce/NiAl73 (12)83 (08)43(10)0.7316704
Au/Ce/NiAl74 (22)88 (10)44(20)0.7240690
* In parentheses, the percentage of decrease after 6 h at 700 °C is given. ** The accumulated carbon during the DRM calculated from TGA weight loss as (Wi − Wf)/Wf (initial weight − final weight)/final weight. *** The DTG temperature peak.
Table 3. Catalytic results of CH4-TPSR test for the HT-derived samples with corresponding TGA-derived carbon information.
Table 3. Catalytic results of CH4-TPSR test for the HT-derived samples with corresponding TGA-derived carbon information.
SamplesT Light − off
(C°)
T peak
(°C)
Tdec range
(°C)
%C * TC **
(°C)
NiAl360562234300702
Au/NiAl450591194144702
Ce/NiAl422582208426704
Au/Ce/NiAl386584257566732
* The accumulated carbon during TPSR calculated from TGA weight loss as (Wi − Wf)/Wf (initial weight- final weight)/final weight. ** The DTG temperature peak.
Table 4. XPS binding energies (eV) and XPS-derived atomic ratios of the HT-derived samples.
Table 4. XPS binding energies (eV) and XPS-derived atomic ratios of the HT-derived samples.
SamplesNi 2p3/2
(eV)
Ni 3p
(eV)
Au 4f7/2
(eV)
Al 2p
(eV)
Ni/Al XPS * Au/Al XPS *
NiAl855.768.6 74.60.5 (2.5)-
Au/NiAl855.868.484.074.50.6 (2.5)0.004 (0.036)
Ce/NiAl855.668.4 74.50.7 (2.5)-
Au/Ce/NiAl855.668.584.074.60.6 (2.5)0.007 (0.037)
* The values in parentheses correspond to the bulk analytical ratios.
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

La Parola, V.; Pantaleo, G.; Liotta, L.F.; Venezia, A.M.; Gabrovska, M.; Nikolova, D.; Tabakova, T. Gold and Ceria Modified NiAl Hydrotalcite Materials as Catalyst Precursors for Dry Reforming of Methane. Catalysts 2023, 13, 606. https://doi.org/10.3390/catal13030606

AMA Style

La Parola V, Pantaleo G, Liotta LF, Venezia AM, Gabrovska M, Nikolova D, Tabakova T. Gold and Ceria Modified NiAl Hydrotalcite Materials as Catalyst Precursors for Dry Reforming of Methane. Catalysts. 2023; 13(3):606. https://doi.org/10.3390/catal13030606

Chicago/Turabian Style

La Parola, Valeria, Giuseppe Pantaleo, Leonarda Francesca Liotta, Anna Maria Venezia, Margarita Gabrovska, Dimitrinka Nikolova, and Tatyana Tabakova. 2023. "Gold and Ceria Modified NiAl Hydrotalcite Materials as Catalyst Precursors for Dry Reforming of Methane" Catalysts 13, no. 3: 606. https://doi.org/10.3390/catal13030606

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