Next Article in Journal
Enhanced Degradation of Phenol by a Fenton-Like System (Fe/EDTA/H2O2) at Circumneutral pH
Next Article in Special Issue
Thermal Deactivation of Rh/α-Al2O3 in the Catalytic Partial Oxidation of Iso-Octane: Effect of Flow Rate
Previous Article in Journal
Removal of Organic Micropollutants from a Municipal Wastewater Secondary Effluent by UVA-LED Photocatalytic Ozonation
Previous Article in Special Issue
A Case Study for the Deactivation and Regeneration of a V2O5-WO3/TiO2 Catalyst in a Tail-End SCR Unit of a Municipal Waste Incineration Plant
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Catalytic Behaviour of Ce-Doped Ni Systems Supported on Stabilized Zirconia under Dry Reforming Conditions

by
Ahmed Sadeq Al-Fatesh
1,*,
Yasir Arafat
1,*,
Ahmed Aidid Ibrahim
1,
Samsudeen Olajide Kasim
1,
Abdulrahman Alharthi
2,
Anis Hamza Fakeeha
1,
Ahmed Elhag Abasaeed
1,
Giuseppe Bonura
3 and
Francesco Frusteri
3
1
Chemical Engineering Department, College of Engineering, King Saud University, P.O. Box 11421 Riyadh, Saudi Arabia
2
College Sciences and Humanities, Prince Sattam Bin Abdulaziz University, Al-Kharj 11942, Saudi Arabia
3
CNR-ITAE, Istituto di Tecnologie Avanzate per l’Energia “Nicola Giordano”, Via S. Lucia sopra Contesse 5, 98126 Messina, Italy
*
Authors to whom correspondence should be addressed.
Catalysts 2019, 9(5), 473; https://doi.org/10.3390/catal9050473
Submission received: 19 April 2019 / Revised: 12 May 2019 / Accepted: 16 May 2019 / Published: 22 May 2019
(This article belongs to the Special Issue Catalysts Deactivation, Poisoning and Regeneration)

Abstract

:
Ni supported on bare and modified ZrO2 samples were synthesized using the incipient wet impregnation method. The t-ZrO2 phase was stabilized by incorporation of La2O3 into ZrO2. Moreover, the influence of CeO2-doping on the physico-chemical and catalytic properties under CO2 reforming conditions was probed. The characterization data of the investigated catalysts were obtained by using XRD, CO2/H2-TPD, BET, TPR, TPO, TGA, XPS and TEM characterization techniques. In the pristine Ni/Zr catalyst, the t-ZrO2 phase transformed into the monoclinic phase. However, upon support modification by La2O3, significant effects on the physicochemical properties were observed due to the monoclinic-to-tetragonal ZrO2 phase transformation also affecting the catalytic activity. As a result, superior activity on the La2O3 modified Ni/Zr catalyst was achieved, while no relevant change in the surface properties and activity of the catalysts was detected after doping by CeO2. The peculiar behavior of the Ni/La-ZrO2 sample was related to higher dispersion of the active phase, with a more pronounced stabilization of the t-ZrO2 phase.

Graphical Abstract

1. Introduction

Modern civilization is confronted with two major challenges: dwindling energy resources [1,2] and global warming caused by greenhouse gases (mainly CH4 and CO2) [3,4]. The dry reforming of the methane (DRM) reaction for the conversion of greenhouse gases into a valuable synthesis gas (H2 and CO) product is a potential contestant for confronting both these challenges simultaneously. Syngas is recognized as the building block for the production of H2 and liquid fuels such as olefins, paraffins, methanol, oxygenates and aromatics in the petrochemical industry, employing the Fischer–Tropsch synthesis process [5,6].
It is an acknowledged fact that DRM is hampered by some side reactions, such as the Boudouard reaction (Equation (2)), water gas shift reaction (Equation (3)), and methane decomposition (Equation (4)), which are considered the major reactions leading to the deactivation of catalysts as a result of coke deposition [7].
C H 4 +   C O 2 2 C O   +   2 H 2    Δ H 0 =   + 261   k J   m o l 1
2 C O C   +   C O 2         Δ H 0 =   171   k J   m o l 1
H 2 +   C O 2 H 2 O   +   C O      Δ H 0 =   + 41   k J   m o l 1
C H 4 2 H 2 +   C          Δ H 0 =   + 75   k J   m o l 1
Coke formation represents a serious problem in DRM. Therefore, noble metal catalysts like Ir, Pt, Rh, Ru and Pd, based on anti-coking properties, were used to constrain such phenomena [8]. In spite of their marvelous characteristics, noble metals are rare and expensive, and end up with a higher cost-benefit ratio. Therefore, the application of noble metals is no longer commercially feasible. On the other hand, Ni-based catalysts are cheap and abundant, and also demonstrate an interesting catalytic performance for the DRM reaction. Unfortunately, nickel particles undergo thermal sintering and tend to promote coke formation over the catalyst surface, leading to catalyst deactivation. Therefore, besides the relevance of active metal components, the role of support material is crucial in inhibiting (or at least limiting) the formation of coke. This can be accomplished by: modifying the electronic properties of the catalyst through a control of the metal-support interaction [9]; improving the oxygen storage capacity [10]; or controlling the size of active metal particles [11]. Consequently, several oxide supports, like Al2O3, SiO2, ZrO2, CeO2 and La2O3 have been employed to enhance the catalytic activity and stability of Ni catalysts. Among different oxides, ZrO2 is extensively employed as a functional support, possessing several unique properties like redox properties, acid-base bi-functional properties, thermal and mechanical stability, high ionic conductivity and oxygen transport properties, making it a favourable support for reforming reactions [12,13]. Moreover, the partial CO2 activation, as well as the strong anchoring effects endowed by Zr4+, may effectively boost the DRM reaction. However, it was found that Ni/ZrO2 catalysts undergo serious coke formation over their surface, leading to fast deactivation and, subsequently, reactor plugging [14,15].
The deactivation of the catalyst may occur for several reasons, for instance, commonly available ZrO2 possesses a rather smaller surface area (viz. <50 m2/g) in comparison to its counterparts [16,17]. Moreover, ZrO2 is characterized by three polymorph structures, named as the monoclinic phase (m-ZrO2, room temperature–1175 °C), tetragonal phase (t-ZrO2, 1175–2370 °C), and cubic phase (c-ZrO2, 2370–2680 °C) [18]. The c-ZrO2 is poorly stable at room temperature, which restricts its large scale application in catalysis as compared to m-ZrO2 and t-ZrO2. Furthermore, t-ZrO2 is found to have a better performance than m-ZrO2 [19]. In addition, t-ZrO2 is widely utilized as a support for Ni catalysts in CH4 reforming processes, even if its thermodynamic instability represents a problem to overcome. Rezaei et al. proposed that the surface energy difference between monoclinic and tetragonal phases may induce the t-ZrO2 phase to be thermodynamically stable for tiny crystals [20]. Consequently, in order to exploit the excellent properties associated with ZrO2 support, it is essential to find a way to prepare high surface area ZrO2 and stabilizing t-ZrO2, which will obviously contribute to enhancing the catalytic performance.
Recently, heteroatom oxide compounds (e.g., Y2O3, CeO2, TiO2 and K2O) have been reported to stabilize t-ZrO2 in catalytic processes [19,21]. When La2O3 is assigned to modify the support for the dry reforming reaction, the ability of support to activate CO2 was enhanced by means of La2O2CO3 formation. In this way, it discards the accumulated carbon through oxidation of coke at the Ni-La2O3 interface [22,23].
Furthermore, in order to promote the oxygen storage capacity of ZrO2, as well as to enhance the thermal stability of Ni-based catalysts, CeO2 was proposed as a promoter (in fact, ceria demonstrates outstanding redox properties as a result of a quick transition between Ce4+/Ce3+).
The purpose of this study was to exploit the aforementioned peculiar properties of ZrO2 support, trying to circumvent the limitations associated with its instability and low surface area. To achieve this objective, high surface area ZrO2 supported Ni catalysts were prepared using the incipient wet-impregnation method. Subsequently, ZrO2 support was modified by heteroatom oxides, like La2O3, with an intention that it might stabilize the t-ZrO2 phase. In addition, 1% CeO2 promoter was also incorporated to enhance the oxygen storage capacity of catalysts. Consequently, the unique properties achieved as an outcome of the synergistic effect of the ceria promoter and high surface area ZrO2 support modifiers (La2O3) were studied to stabilize t-ZrO2 and their eventual influence on the catalytic performance.

2. Results and Discussion

2.1. Physicochemical Features of the Ceria Promoted Ni/x-Zr (x = 0, La2O3) Catalysts

Figure 1a illustrates the X-ray diffraction patterns of fresh catalyst samples. The diffraction peaks of NiO can be featured on the diffractograms at 2θ = 37.3° and 43.3° corresponding to the [101] and [012] reflections, respectively.
According to the literature, monoclinic zirconia (m-ZrO2) was found at 2θ≈24.0°, 28.2°, 31.5°, 34.2°, 34.4°, 35.3° and 40.7°, while tetragonal zirconia (t-ZrO2) appeared at 2θ≈30.0°, 34.8°, 35.1°, 50.0° and 59.4° [24]. Based on this evidence, it can be observed that “pure” Ni/Zr catalyst predominantly consists of t-ZrO2. After incorporation of La2O3, the peaks related to t-ZrO2 became sharper and more pronounced. It implies that La2O3 contributed to enhance the stability of the t-ZrO2 phase of the Ni/Zr catalyst. Moreover, it is clear that the crystalline La2O3 phase (2θ = 28° and 49°, JCPDS: 01-089-4016) was not individually identified on the diffractograms, indicating that La2O3 was highly dispersed and incorporated into the ZrO2 lattice. Furthermore, when the Ni/Zr catalyst was promoted using 1% CeO2 (see Figure 1b), more intense diffraction patterns attributing to t-ZrO2 appeared for the Ni-Ce/Zr catalyst. However, when Ni/La-Zr was promoted by 1% CeO2, the intensity of diffraction peaks corresponding to t-ZrO2 was considerably reduced for the Ni-Ce/La-Zr catalyst.
The textural properties of Ni/x-Zr (x = 0, La2O3) and CeO2 promoted fresh catalysts are shown in Figure 2 and Table 1. It is evident that the incorporation of lanthanum as a ZrO2 modifier in the catalyst composition led to a significant enhancement of surface area and porosity with respect to the reference Ni/Zr sample.
In particular, the Ni/La-Zr sample presented the highest cumulative pore volume (P.V., 0.246 cm3/g), accounting for an average pore diameter (P.D.) just smaller than 16 nm. However, the undoped Ni/Zr sample exhibited the lowest extension of surface area (39 m2/g) along with the lowest porosity (0.111 cm3/g).
Moreover, the N2 adsorption–desorption branch of the Ni/Zr sample exemplifies a typical type IV isotherm, characteristic for mesoporous solids with a H4-type hysteresis loop. The doping of ZrO2 with La2O3 dramatically affected the hysteresis loop, transforming it into a H1-type hysteresis loop, associated with a cylindrical pore geometry as well as relatively high uniformity in pore size. Consequently, the La2O3 modified catalyst possessed a high surface area and a mesoporous structure. After promotion with 1% CeO2, no significant change in the textural properties of the catalysts was observed either in quantitative (see Table 1) or qualitative (see Figure 2) terms.
The reducibility and metal-support interaction of the Ni/Zr system after the support modification was evaluated by TPR analysis. Figure 3 illustrates the reduction profile of Ni/x-Zr (x = 0, La2O3) and Ni-Ce/x-Zr (x = 0, La2O3) samples.
It is evident from the TPR profile that the modification of ZrO2 support by the addition of La2O3 has a considerable influence on the reduction profiles, being in all cases characterized by the occurrence of multiple reduction peaks, spanning the range 100–1000 °C, and ascribed to different nickel species with different degrees of interaction with the support. A first reduction peak observed in the range 260–290 °C with a shoulder at some lower temperature in the case of CeO2 promoted catalysts, may be assigned to the reduction of surface oxygen species [25,26]. The second reduction peak can be referred to the reduction of NiO grafted onto the support by the weak interaction. The reduction peak at a higher temperature (>750 °C) may be attributed to the development of NiO–ZrO2 solid solutions [27]. It is interesting to observe that the modification of ZrO2 with La2O3 considerably improved the reduction kinetics, considering that NiO was completely reduced (NiO→Ni) below 550 °C owing to the inhibition of formation of NiO–ZrO2 solid solutions. It implies that the incorporation of La2O3 to ZrO2 had made it feasible to achieve the active metal at some lower temperature accompanied by the improved dispersion of metallic species. Furthermore, upon ZrO2 modification using lanthana, the peaks corresponding to NiO–ZrO2 solid solutions seemed to be completely suppressed. It implies that La2O3 modification also prevents the formation of NiO–ZrO2 solid solutions. On the other hand, La2O3 modification also contributed to enhancing the metal–support interaction, which is evident from the drop in H2 consumption from Ni/Zr to the Ni/La-Zr catalyst by 22% and from Ni-Ce/Zr to the Ni-Ce/La-Zr catalyst by 17% (Table 1). It is likely that the enhancement in the metal–support interaction is accompanied by the improvement in active phase dispersion. Likewise, Rotgerink et al. [28] also established that the incorporation of La2O3 to the Ni/Zr catalyst upgrades the reducibility and Ni dispersion. Some other studies have also found that the improvement in the reducibility of NiO is the outcome of the ability of La2O3 to disperse Ni metallic species on the ZrO2 support [29,30]. After doping by CeO2, the reduction patterns of catalysts had not significantly changed. Furthermore, in the case of the Ni-Ce/La-Zr catalyst, a peak centred at 225 °C, corresponding to surface oxygen species, became prominent, which may be the outcome of the synergistic effect of CeO2 and La2O3. Actually, the La2O3 lattice is recognized for surface and bulk oxygen vacancies, and their density is associated with the La2O3 concentration [31].
The quantitative data of H2-TPD measurements are listed in Table 2.
Given the same concentration of Ni in the catalysts, it was observed how the various modifiers of the ZrO2 support, as well as the promotion of the active phase with ceria, affect the surface characteristics of the prepared samples. In fact, the “unpromoted” Ni/Zr sample exhibited the lowest H2 uptake (107.3 mol/gcat), accounting for a minor metal surface area (8.4 m2/gcat) and dispersion (25.2%), as the result of particles not larger than 4 nm generated during preparation. Instead, promotion of carrier oxide by lanthanum oxide positively influenced the Ni/Zr system, favouring higher metal surface area and dispersion (16.3–19.8 m2/g and 49.2–59.6%, respectively), as the result of particles with a smaller diameter (1.7–2.1 nm). Regarding the effect of ceria, a promoting effect was markedly evident only on the Ni/Zr system, allowing an enhancement both in terms of metal surface area (13.6 m2/g) and nickel dispersion (40.8%). In contrast, it is clearly visible that the Ni-Ce/La-Zr sample showed that the surface properties dropped with respect to the sample without Ce (Ni/La-Zr). This suggests that ceria promotion has some inhibiting effect on the Ni/La-Zr catalyst.
The basic properties of Ni/x-Zr (x = 0, La2O3) catalysts were estimated by CO2 desorption measurements. Figure 4 illustrates the TPD profiles of the catalysts. It is recognized that the strength of basic sites is determined by the adsorption and desorption peak of CO2 in the corresponding temperature range: weak basic sites (50–200 °C), medium/Lewis basic sites (200–400 °C) and strong basic sites (400–650 °C).
The TPD profiles depicted the weak Lewis and moderate Lewis basicity of the reference Ni/Zr and La2O3-modified samples. It is clearly visible that the Ni/Zr catalyst experienced no significant variation in the basic character after CeO2 promotion except a rise in basicity in La2O3 modified catalysts.
XPS analyses were conducted to determine and interpret the chemical state of elements present in the catalysts and the correspondent spectra are illustrated in Figure 5.
Ni 2p peaks of all the catalysts Ni/Zr, Ni-Ce/Zr and Ni/La-Zr are presented in Figure 5a. The binding energy of the main Ni 2p peak may be found at 855.6 eV, while the rest of the peaks and satellites are visible at other binding energies. It is apparent from the shape of the main peak (Ni 2p3/2) that it is the combination of Ni0 and Ni2+/Ni3+ oxides [32,33]. The Ni 2p spectra of Ni/La-Zr and Ni-Ce/Zr catalysts becomes complicated due to the overlapping of La 3d, Ce 3d, and La MNN Auger peaks. As a result, it gives rise to uncertainty in the quantification of the XPS spectra. Generally, the largest uncertainty is confronted as an outcome of the overlapping of Ni 2p3/2 and La 3d3/2. However, in our case, the Ni/La-Zr catalyst was found to have the highest Ni concentration. The Zr 3d peaks of Ni/Zr, Ni/La-Zr and Ni-Ce/Zr catalysts are shown in Figure 5b. Zr 3d3/2 and Zr 3d5/2 can be found at 190.1 eV and 188.2 eV, respectively, for the Ni/Zr catalyst. Upon support modification using La2O3, the intensity of the Zr 3d peak significantly increased and it shifted to a higher binding energy. Likewise, CeO2 promoted peaks also shifted toward a higher binding energy. The shifting of binding energy to higher values may be attributed to the enhancement of oxygen vacancy as a result of an oxygen deficient state. The phenomenon of shifting of Zr 3d binding energy is related to the large quantity of lattice defects (oxygen vacancies) [34]. The O 1s data for Ni/Zr, Ni-Ce/Zr, Ni/La-Zr, used Ni/Zr and Ni/La-Zr catalysts are illustrated in Figure 5c. The binding energy of O 1s located at 529.6 eV, corresponding to low energy primary oxygen, is consistent for all the catalysts. This peak may be attributed to lattice oxygen (O2–lattice). The O 1s XPS data also displayed a small amount of COx species, which implies that COx species are present as reaction intermediates on the catalyst surface. Consequently, significant variation in chemical states of the catalyst after La incorporation into Ni/Zr implies that support modification using La2O3 has substantial influence.

2.2. Catalytic Activity and Stability

The catalytic behavior of the prepared samples was investigated under dry reforming conditions, in terms of carbon dioxide and methane conversion as a function of time on stream for Ni/x-Zr and Ni-Ce/x-Zr (x = 0, La2O3) catalysts (Figure 6). The effects of ZrO2 modification, as well as Ni doping with cerium oxide on the Ni/Zr system, were analyzed. A maximum reaction temperature of 700 °C was chosen, considering that at higher temperature, the formation of encapsulated carbon leads to a rapid deactivation of the system [35,36].
It can be seen that not only La modification affects Ni/Zr activity, but also ceria significantly influences the catalytic behavior of the investigated samples. The “reference” Ni/Zr sample exhibited the worst performance, with an initial CH4 and CO2 conversion of 35 and 46%, respectively, which decreased to 20–30% after only 400 min. After the incorporation of La, the conversion rate of the catalyst samples was significantly enhanced. Interestingly, the best catalytic performance was obtained by using the La-modified sample, with similar initial values of CH4 and CO2 conversion (ca. 70%) and final conversion of 61 (CH4) and 67% (CO2), accounting for a slightly higher stability in respect to the other samples. Regarding the influence of the CeO2 promoter, from Figure 6 a net improvement in the activity of the Ni-Ce/Zr sample can be observed, not only in terms of initial conversion values of CH4 and CO2 (54 and 64% respectively), but also in terms of stability, considering that after 400 min the CH4 conversion decreased from 54 to 45%, and CO2 conversion decreased from 64 to 55%, which were lower with respect to the values recorded in the correspondent catalyst without ceria. Instead, 1% CeO2 promoter on the Ni phase showed an inhibiting effect on the Ni/La-Zr catalyst, since the activity of the Ni-Ce/La-Zr sample dropped to lower conversion values with respect to the correspondent unpromoted catalyst, by exhibiting a more pronounced decay trend of CH4 conversion from 61 to 52%. In terms of the H2/CO ratio (Figure 6c), a decreasing trend over time on the stream is visible for Ni/Zr, progressively favouring a bit greater formation of CO (Equation (3)) with a final H2/CO value of 0.6. However, upon La2O3 modification, an equimolar formation of hydrogen and carbon monoxide with a H2 / CO ≥ 1 was achieved. On the whole, looking at the properties of the investigated samples, an almost straight-line relationship was found between the initial CO2 conversion rate and the nickel surface area (see Figure 7). It seems evident that a controlled modification of the carrier structure determines a better surface exposure of the active phase (Ni), and therefore inducing higher CO2 activation.
The catalyst activity must also be put in direct relation to the catalyst structure, considering that, depending on modification by oxides, ZrO2 can become a prominently more stable phase. In general, a substantially higher activity is observed when ZrO2 exists as a tetragonal phase, resulting in appreciably suppressed m-ZrO2 phase even in the used catalysts, like the Ni/La-Zr and Ni-Ce/La-Zr samples (see Figure 1a,b). On the other hand, initially, Ni/Zr and Ni-Ce/Zr had a prominent t-ZrO2 phase, however, upon thermal treatment, tetragonal-to-monoclinic phase transformation occurred.
To better analyze the obtained results, a deactivation factor (DF) was also deduced for all the samples as the ratio between the loss of activity calculated between the initial (20 min on stream) and final (400 min on stream) time and the initial conversion value of CH4. From Figure 8, it is evident that the incorporation of La2O3 considerably improves the catalyst stability of the Ni/Zr system, as the deactivation factor decreased from 44% to 22%, with a lower net drop of the activity in respect of the initial time under the same experimental conditions. Moreover, the addition of ceria leads to a further improvement of the catalytic stability for both the undoped Ni/Zr and Ni/La-Zr samples, yielding the smallest loss of activity during time (< 12%) in the Ni-Ce/La-Zr sample.
To investigate the phase analysis following the catalytic treatment, post reaction XRD of the catalysts was performed (Figure 1). It was revealed that the thermal treatment had a profound influence on the phase transformation. For instance, used Ni/Zr catalyst found to have severe lattice distortion and t-ZrO2 had significantly diminished and transformed into m-ZrO2. Likewise, even CeO2 promotion to Ni/Zr catalyst was not able to completely preserve the t-ZrO2. However, upon support modification using La2O3, the t-ZrO2 phase was perfectly retained with no appearance of m-ZrO2. It implies that La2O3 incorporation has a significant contribution in stabilizing the t-ZrO2 and achieving the 100% rise in catalytic activity. Similarly, used Ni-Ce/La-Zr also prevented the tetragonal-to-monoclinic transformation and pure t-ZrO2 was preserved.
As a main possible reason for the decay of activity during time, it was obviously necessary to consider the coke formation over the different samples, as it is known that the amount and typology of coke is also influenced by the catalyst properties. On this account, temperature programmed oxidation (TPO) ceria promoted and unpromoted Ni/x-ZrO2 (0, La2O3) catalysts were conducted to determine the type of the deposited carbon during the CO2 reforming reactions. The TPO peaks are illustrated in Figure 9a.
The major peak in each profile is positioned around 585 °C, suggesting the presence of carbon nanotubes (CNTs). TGA analysis of spent catalysts was performed on reference and promoted Ni/Zr catalysts (Figure 9b) to quantify the amount of deposited carbon. It was found that no significant weight loss occurs in the case of a pure catalyst, indicating that a low amount of carbon was deposited during the reaction. On the contrary, La-modified catalysts presented a higher weight loss. After CeO2 promotion, TGA profiles illustrate that the amount of accumulated carbon on the catalyst surface had considerably reduced. Moreover, the corresponding peaks negatively shifted, implying that the type of carbon material formed, is less stable, and burnt at a relatively lower temperature. It implies that carbon nanotubes (CNTs) mainly formed after CeO2 promotion, and are easily burnt and affect catalyst deactivation less [7,37,38,39].
Figure 10 and Figure 11 show the TEM images of ceria promoted and unpromoted Ni/x-ZrO2 (0, La2O3) catalysts.
It can be seen that the majority of the deposited carbon on the Ni/ZrO2 surface, after the catalytic reaction, is of a filamentous nature, and a tangle of carbon nanotubes (CNTs) is clearly visible (Figure 10A), while the rest of the carbon is of the encapsulating type. It is clear that several particles were encapsulated due to the carbon formation on Ni surface (Figure 10B). Ni/ZrO2 catalysts also undergo sintering, therefore accumulation of metallic species can be seen (Figure 10C). Conspicuously, metal particles varying in size in the range 2–9 nm were entrapped inside the CNTs (Figure 10D,E), which may be another reason for the lower activity of the pristine catalyst.
When the pristine catalyst Ni/Zr was modified with La2O3, morphological investigation revealed the formation of multiwall carbon nanotubes (MWCNTs). The average external diameter of MWCNTs was found to be 28 nm, whereas the metallic species were positioned together in the interior as well as on the exterior of MWCNTs (Figure 11A). It seems that the incorporation of La2O3 had no influence in controlling the sintering of metallic particles (Figure 11B). The Ni particles at the tip of the CNTs clearly demonstrated that Ni particles not strongly anchored to the catalyst support formed filaments with a mechanism that included the diffusion of elementary C through the Ni and precipitation on the back side of the particle with consequent formation of the filament. In this case, Ni continues to be active for dry reforming [40,41,42]. Furthermore, when the Ni/ZrO2 catalyst was promoted using CeO2, the situation escalated because the rate of sintering was significantly enhanced upon CeO2 promotion (Figure 11C), which may justify the deteriorating activity of the Ni-Ce/La-Zr catalyst.

3. Materials and Methods

3.1. Catalyst Preparation

In this experimental work an incipient wet-impregnation technique was used to prepare the desired catalysts for the dry reforming of methane. The supports (zirconium oxide and modified zirconium oxide with La2O3) along with the active metal obtained from the precursor of Ni nitrate [Ni(NO3)2.6H2O; 99.9% purity] were dissolved in distilled water. The amount of Ni used in the catalyst was 5 wt%. The solution was kept heated at 90 °C under stirring for 3 h. Later, impregnated catalysts were heated up to 120 °C overnight for drying and then subjected to calcination at 600 °C for 3 h. The Ce(NO3)3.6H2O promoted catalysts with 1 wt% were prepared by co-impregnation of the nitrate salts of the promoter and active metal with support using the same procedure mentioned above. The solution was kept heated at 90 °C under stirring for 3 h. Later, impregnated catalysts were heated up to 120 °C overnight for drying and then subjected to calcination at 600 °C for 3 h. The commercial samples of the support were obtained from KIGAKU DAIICHI KOGYO CO.; LTD, Osaka, Japan. We are grateful to the company for providing us with free samples.

3.2. Catalytic Testing

Dry reforming of CH4 experiments over Ni/ZrO2 catalysts were performed at 700 °C and at atmospheric pressure in a vertical stainless steel fixed-bed tubular (i.d., 9.1 mm; length, 0.3 m) micro-reactor (PID Eng& Tech micro activity reference, Madrid, Spain). Activity tests were performed using 0.1 g of a catalyst placed in a quartz reactor between two quartz wool beds. A K-type stainless steel sheathed thermocouple, placed axially at the center of the catalyst bed, measured the temperature during the reaction. Prior to each test, the catalysts were activated under a continuous flow of H2 (20 mL/min) for 1 h at 600 °C. Experiments were carried out using a feed gas mixture (CH4, CO2 and N2) at the ratio 6/6/1 and overall gas flow rate of 65 mL/min (space velocity: 39,000 mL/)h.gcat)). The outlet gas composition was analyzed by on-line gas chromatography (Shimadzu 2014, Kyoto, Japan) fitted out with a thermal conductivity detector (TCD). The CH4 conversion and hydrogen yield were determined using the following formulae:
C H 4   c o n v e r s i o n ( % ) = ( C H 4 , i n C H 4 , o u t ) C H 4 , i n * 100
H 2   y i e l d ( % ) = ( H 2 , o u t ) 2 * C H 4 , i n * 100

3.3. Catalyst Characterization

3.3.1. XRD Characterization

Rigaku (Miniflex) diffractometer (Rigaku Corporation, Dover, DE, USA), with a Cu Kα X-ray radiation working at 40 kV and 40 mA, was used to investigate the structure of the catalysts before and after the reaction. The scanning 2θ range and steps were 10–85° and 0.02° respectively. The raw data file of the instrument was evaluated by X’Pert high score plus software (version 2.1, Panalytical, Malvern, UK). Different phases with their scores were corresponded with the JCPDS data bank.

3.3.2. N2 Physisorption

The specific surface area and distribution of pore size of the catalysts were obtained using Micromeritics Tristar II 3020 surface area and porosity analyzer. The pore size distribution was calculated by BJH method.

3.3.3. Temperature Programmed Reduction (TPR)

The TPR measurements were performed using Micromeritics Auto Chem II apparatus (Micromeritics, Atlanta, GA, USA). The sample (70 mg) was charged in the TPR cell and flushed with argon at 150 °C for 30 min. Then the sample temperature was reduced to 25 °C. Finally, the furnace temperature was heated to 1000 °C at 10 °C/min ramp at a 40 mL/min flow rate with a H2/Ar mixture (10:90 vol.%). A thermal conductivity detector (TCD) was used to follow the signals of H2 consumption.

3.3.4. H2 Temperature-Programmed Desorption

To evaluate metal surface area (MSA), and surface average nickel particle size (dNi) of the investigated catalysts, H2-TPD measurements in the range 20–900 °C (β = 10 °C/min) were carried out at atmospheric pressure by using a linear quartz micro-reactor (i.d., 4 mm) flowing Ar as carrier gas at 50 stp cm3/min. Before measurements, a catalyst sample (50–100 mg) was reduced for half-an hour in flowing H2 (25 stp cm3/min) at 600 °C. Thereafter, the sample was cooled in flowing H2 to room temperature and then hydrogen was shut off and the sample was purged by the carrier stream until baseline stabilization (≈20 min). Assuming a chemisorption stoichiometry H2:Nisurf = 1:2 and a spherical shape of the metal particle, the following equations were applied:
MSA (m2/gcat) = 2·XH2·NAvNi
dNi (nm) = 6000·(CNi/100)/(ρNi·MSA)
where “XH2” is the H2 uptake (mol/gcat), “NAv” is Avogadro’s number, “σNi” represents the concentration of surface atoms (1.54 × 1019 at/m2 for Ni), “CNi” is the metal loading (wt%), while “ρNi” is the metal density (8.9 g/cm3 for Ni).

3.3.5. CO2 Temperature-Programmed Desorption (CO2-TPD)

The CO2 temperature-programmed desorption (TPD) was carried out using a Chemisorption Analyzer (Micromeritics Autochem II apparatus, Micromeritics, Atlanta, GA, USA). The catalyst (50 mg) was reduced at 600 °C for 1 h under He flow (30 mL/min) and then cooled to 50 °C. The CO2 flow was kept for 60 min, and the sample was then flushed with He to take away any physisorbed CO2. The desorption profile of the catalysts were recorded by ramping the temperature at a rate of 10 °C/min, while temperature was then linearly increased up to 800 °C. The CO2 concentration in the output stream was measured with a thermal conductivity detector, and the areas under the peaks were used to determine the amount of desorbed CO2 during TPD.

3.3.6. Thermo-Gravimetric Analysis (TGA)

The analysis of carbonaceous material deposition post reaction over the catalyst’s surface was quantitatively performed using the TGA-15 SHIMADZU analyzer (Shimadzu, Kyoto, Japan) under an air atmosphere. The spent catalyst (10–15 mg) was heated from room temperature to 1000 °C at a heating rate of 20 °C/min and the loss of weight was measured.

3.3.7. TEM Characterization

TEM measurements of the used samples were accomplished on a 120 kV JEOL JEM-2100F transmission electron microscope (JEOL, Peabody, MA, USA).

3.3.8. XPS Analysis

X-ray photoelectron spectroscopy (XPS) spectra were recorded on an Omicron Nanotechnology (ELS5000) spectrometer with a monochromatic Al source. X-ray was employed using a flood gun with a spot size of radius ~ 400 μm. Survey scanning was acquired from −10 to 1350 eV with energy steps of 1 eV, while employing a pass energy of 200 eV. The number of scans was four with a dwell time of 10 ms. Similarly, a slow scan was conducted for the element using a pass energy of a 50 eV and the number of scans was four with a dwell time of 10 ms. The charging effects were corrected by adjusting the binding energy of the C 1s peak from adventitious carbon to 285 eV. The characterization experiments were carried out for fresh (before reaction) and used (after reaction) samples.

4. Conclusions

A high surface area ZrO2 supported Ni system was prepared using an incipient wet-impregnation method and then modified by heteroatom oxides, like La2O3, and the influence of ceria doping in dry reforming of CH4 at 700 °C was also assessed. It is worthwhile highlighting that the reduction properties of the catalysts were significantly enhanced upon La2O3 modification. La2O3 modification suppressed the formation of NiO–ZrO2 solid solutions, with a complete reduction of the active phase at temperatures below 750 °C. Furthermore, La2O3 modification also contributed to enhancing the metal-support interaction and active phase dispersion. However, the weak-medium basicity of all modified catalyst samples were substantially unchanged after Ce-promotion. Phase analysis revealed that the t-ZrO2 phase of the Ni/Zr catalyst had almost completely transformed into the m-ZrO2 phase after thermal treatment. However, when La2O3 incorporated into the ZrO2 support, the t-ZrO2 phase was protected and prominently stabilized, leading to a superior catalytic performance. Interestingly, CH4 conversion increased to 2x, and CO2 conversion to 1.5x, that of pristine Ni/ZrO2. However, when the modified catalysts were promoted by CeO2, a decline in catalytic activity was observed in the case of the Ni-Ce/La-Zr catalyst. Eventually, it was revealed that the high activity of La2O3 modified catalysts was achieved by protecting the stability of the t-ZrO2 phase.

Author Contributions

A.S.-F., Y.A., A.A.I., S.O.K. and A.H.F. synthesized the catalysts, performed all the experiments and characterization tests and wrote the manuscript. F.F., A.A. and G.B. performed the characterization tests for CO2, H2-TPD, XPS and TEM, and proofread the manuscript. A.E.A. contributed to the analysis of the data and proofread the manuscript.

Funding

The authors would like to express their sincere appreciation to the Deanship of Scientific Research at King Saud University for funding this research project (No. RGP-1435-078).

Acknowledgments

The authors would like to extend their sincere appreciation to the Deanship of Scientific Research at King Saud University for its funding this research group No. (RGP-1435-078).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Pan, X.; Fan, Z.; Chen, W.; Ding, Y.; Luo, H.; Bao, X. Enhanced ethanol production inside carbon-nanotube reactors containing catalytic particles. Nat. Mat. 2007, 6, 507–511. [Google Scholar] [CrossRef] [PubMed]
  2. Hamza Fakeeha, A.; Arafat, Y.; Aidid Ibrahim, A.; Shaikh, H.; Atia, H.; Elhag Abasaeed, A.; Armbruster, U.; Sadeq Al-Fatesh, A. Highly Selective Syngas/H2 Production via Partial Oxidation of CH4 Using (Ni, Co and Ni–Co)/ZrO2–Al2O3 Catalysts: Influence of Calcination Temperature. Processes 2019, 7, 141. [Google Scholar] [CrossRef]
  3. Gao, X.Y.; Hidajat, K.; Kawi, S. Facile synthesis of Ni/SiO2 catalyst by sequential hydrogen/air treatment: A superior anti-coking catalyst for dry reforming of methane. J. CO2 Util. 2016, 15, 146–153. [Google Scholar] [CrossRef]
  4. Xu, Y.; Lin, Q.; Liu, B.; Jiang, F.; Xu, Y.; Liu, X. A Facile Fabrication of Supported Ni/SiO2 Catalysts for Dry Reforming of Methane with Remarkably Enhanced Catalytic Performance. Catalysts 2019, 9, 183. [Google Scholar] [CrossRef]
  5. Karam, L.; Casale, S.; El Zakhem, H.; El Hassan, N. Tuning the properties of nickel nanoparticles inside SBA-15 mesopores for enhanced stability in methane reforming. J. CO2 Util. 2017, 17, 119–124. [Google Scholar] [CrossRef]
  6. El Hassan, N.; Kaydouh, M.; Geagea, H.; El Zein, H.; Jabbour, K.; Casale, S.; El Zakhem, H.; Massiani, P. Low temperature dry reforming of methane on rhodium and cobalt based catalysts: Active phase stabilization by confinement in mesoporous SBA-15. Appl. Catal. A Gen. 2016, 520, 114–121. [Google Scholar] [CrossRef] [Green Version]
  7. Al-Fatesh, A.S.; Arafat, Y.; Atia, H.; Ibrahim, A.A.; Ha, Q.L.M.; Schneider, M.; M-Pohl, M.; Fakeeha, A.H. CO2-reforming of methane to produce syngas over Co-Ni/SBA-15 catalyst: Effect of support modifiers (Mg, La and Sc) on catalytic stability. J. CO2 Util. 2017, 21, 395–404. [Google Scholar] [CrossRef]
  8. Moradi, G.R.; Rahmanzadeh, M.; Khosravian, F. The effects of partial substitution of Ni by Zn in LaNiO3 perovskite catalyst form ethane dry reforming. J. CO2 Util. 2014, 6, 7–11. [Google Scholar] [CrossRef]
  9. Kambolis, A.; Matralis, H.; Trovarelli, A.; Papadopoulou, C. Ni/CeO2-ZrO2 catalysts for the dry reforming of methane. Appl. Catal. A 2010, 377, 16–26. [Google Scholar] [CrossRef]
  10. Lemonidou, A.A.; Vasalos, I.A. Carbon dioxide reforming of methane over 5 wt.% Ni/CaO-Al2O3 catalyst. Appl. Catal. A 2002, 228, 227–235. [Google Scholar] [CrossRef]
  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. Świrk, K.; Gálvez, M.E.; Motak, M.; Grzybek, T.; Rønning, M.; Da Costa, P. Dry reforming of methane over Zr- and Y-modified Ni/Mg/Al double-layered hydroxides. Catal. Commun. 2018, 117, 26–32. [Google Scholar]
  13. Sarkar, D.; Adak, S.; Chu, M.; Cho, S.; Mitra, N. Influence of ZrO2 on the thermo-mechanical response of nano-ZTA. Ceram. Int. 2007, 33, 255–261. [Google Scholar] [CrossRef]
  14. Lercher, J.; Bitter, J.; Hally, W.; Niessen, W.; Seshan, K. Design of stable catalysts for methane-carbon dioxide reforming. Studies Surf. Sci. Catal. 1996, 101, 463–472. [Google Scholar] [Green Version]
  15. Li, X.; Chang, J.; Park, S. Carbon as an intermediate during the carbon dioxide reforming of methane over zirconia-supported high nickel loading catalysts. Chem. Lett. 1999, 10, 1099–1100. [Google Scholar] [CrossRef]
  16. Chuah, G.K.; Liu, S.H.; Jaenicke, S.; Li, J. High surface area zirconia by digestion of zirconium propoxide at different pH. Microporous Mesoporous Mat. 2000, 39, 381–392. [Google Scholar] [CrossRef]
  17. Jaenicke, S.; Chuah, G.K.; Raju, V.; Nie, Y.T. Structural and Morphological Control in the Preparation of High Surface Area Zirconia. Catal. Surv. Asia 2008, 12, 153–169. [Google Scholar] [CrossRef]
  18. Fan, M.-S.; Abdullah, A.Z.; Bhatia, S. Utilization of greenhouse gases through carbon dioxide reforming of methane over Ni–Co/MgO–ZrO2: preparation, characterization and activity studies. Appl. Catal. B 2010, 100, 365–377. [Google Scholar] [CrossRef]
  19. Yamasaki, M.; Habazaki, H.; Asami, K.; Izumiya, K.; Hashimoto, K. Effect of tetragonal ZrO2 on the catalytic activity of Ni/ZrO2 catalyst prepared from amorphous Ni–Zr alloys. Catal. Commun. 2006, 7, 24–28. [Google Scholar] [CrossRef]
  20. Rezaei, M.; Alavi, S.M.; Sahebdelfar, S.; Yan, Z.-F. A highly stable catalyst in methane reforming with carbon dioxide. Scripta Materialia 2009, 61, 173–176. [Google Scholar] [CrossRef]
  21. Choque, V.; de la Piscina, P.R.; Molyneux, D.; Homs, N. Ruthenium supported on new TiO2–ZrO2 systems as catalysts for the partial oxidation of methane. Catal. Today 2010, 149, 248–253. [Google Scholar] [CrossRef]
  22. Verykios, X.E. Catalytic dry reforming of natural gas for the production of chemicals and hydrogen. Int. J. Hydrogen Energy 2003, 28, 1045–1063. [Google Scholar] [CrossRef]
  23. Slagtern, A.; Schuurman, Y.; Leclercq, C.; Verykios, X.; Mirodatos, C. Specific features concerning the mechanism of methane reforming by carbon dioxide over Ni/La2O3 catalyst. J. Catal. 1997, 172, 118–126. [Google Scholar] [CrossRef]
  24. Titus, J.; Roussiere, T.; Wasserschaff, G.; Schunk, S.; Milanov, A.; Schwab, E.; Wagner, G.; Oeckler, O.; Gläser, R. Dry reforming of methane with carbon dioxide over NiO–MgO–ZrO2. Catal. Today 2016, 270, 68–75. [Google Scholar] [CrossRef]
  25. Ay, H.; Üner, D. Dry reforming of methane over CeO2 supported Ni, Co and Ni–Co catalysts. Appl. Catal. B 2015, 179, 128–138. [Google Scholar] [CrossRef]
  26. Chen, J.; Zhang, X.; Arandiyan, H.; Peng, Y.; Chang, H.; Li, J. Low temperature complete combustion of methane over cobalt chromium oxides catalysts. Catal. Today 2013, 201, 12–18. [Google Scholar] [CrossRef]
  27. Goula, M.A.; Charisiou, N.D.; Siakavelas, G.; Tzounis, L.; Tsiaoussis, I.; Panagiotopoulou, P.; Goula, G.; Yentekakis, I.V. Syngas production via the biogas dry reforming reaction over Ni supported on zirconia modified with CeO2 or La2O3 catalysts. Int. J. Hydrogen Energy 2017, 42, 13724–13740. [Google Scholar]
  28. Rotgerink, H.L.; Paalman, R.; Van Ommen, J.; Ross, J. Studies on the promotion of nickel—alumina coprecipitated catalysts: II. Lanthanum oxide. Appl. Catal. 1988, 45, 257–280. [Google Scholar] [CrossRef]
  29. Wang, S.; Lu, G.; Millar, G.J. Carbon dioxide reforming of methane to produce synthesis gas over metal-supported catalysts: state of the art. Energy Fuels 1996, 10, 896–904. [Google Scholar] [CrossRef]
  30. Kumar, P.; Sun, Y.; Idem, R.O. Comparative Study of Ni-based Mixed Oxide Catalyst for Carbon Dioxide Reforming of Methane. Energy Fuels 2008, 22, 3575–3582. [Google Scholar] [CrossRef]
  31. Huang, S.J.; Walters, A.B.; Vannice, M.A. TPD, TPR and DRIFTS studies of adsorption and reduction of NO on La2O3 dispersed on Al2O3. Appl. Catal. B 2000, 26, 101–118. [Google Scholar] [CrossRef]
  32. Grosvenor, A.P.; Biesinger, M.C.; Smart, R.S.C.; McIntyre, N.S. New interpretations of XPS spectra of nickel metal and oxides. Surf.Sci. 2006, 600, 1771–1779. [Google Scholar] [CrossRef]
  33. Charisiou, N.D.; Siakavelas, G.; Tzounis, L.; Sebastian, V.; Monzon, A.; Baker, M.A.; Hinder, S.J.; Polychronopoulou, K.; Yentekakis, I.V.; Goula, M.A. An in depth investigation of deactivation through carbon formation during the biogas dry reforming reaction for Ni supported on modified with CeO2 and La2O3 zirconia catalysts. Int. J. Hydrogen Energy 2018, 43, 18955–18976. [Google Scholar] [CrossRef]
  34. Liu, G.; Liu, A.; Meng, Y.; Shan, F.; Shin, B.; Lee, W.; Cho, C. Annealing dependence of solution-processed ultra-thin ZrOx films for gate dielectric applications. J. Nanosci. Nanotechnol. 2015, 15, 2185–2191. [Google Scholar] [CrossRef]
  35. Italiano, G.; Delia, A.; Espro, C.; Bonura, G.; Frusteri, F. Methane decomposition over Co thin layer supported catalysts to produce hydrogen for fuel cell. Int. J. Hydrogen Energy 2010, 35, 11568–11575. [Google Scholar] [CrossRef]
  36. Frusteri, F.; Italiano, G.; Espro, C.; Cannilla, C.; Bonura, G. H2 production by methane decomposition: Catalytic and technological aspects. Int. J. Hydrogen Energy 2012, 37, 16367–16374. [Google Scholar] [CrossRef]
  37. Frusteri, F.; Frontera, P.; Modafferi, V.; Bonura, G.; Bottari, M.; Siracusano, S.; Antonucci, P.L. Catalytic features of Ni/Ba–Ce0. 9–Y0. 1 catalyst to produce hydrogen for PCFCs by methane reforming. Int. J. Hydrogen Energy 2010, 35, 11661–11668. [Google Scholar]
  38. Frontera, P.; Macario, A.; Monforte, G.; Bonura, G.; Ferraro, M.; Dispenza, G.; Antonucci, V.; Aricò, A.S.; Antonucci, P.L. The role of Gadolinia Doped Ceria support on the promotion of CO2methanation over Ni and NiFe catalysts. Int. J. Hydrogen Energy 2017, 42, 26828–26842. [Google Scholar] [CrossRef]
  39. Al-Fatesh, A.S.; Arafat, Y.; Ibrahim, A.A.; Atia, H.; Fakeeha, A.H.; Armbruster, U.; Abasaeed, A.E.; Frusteri, F. Evaluation of Co-Ni/Sc-SBA–15 as a novel coke resistant catalyst for syngas production via CO2 reforming of methane. Appl. Catal. A 2018, 567, 102–111. [Google Scholar] [CrossRef]
  40. Bonura, G.; Di Blasi, O.; Spadaro, L.; Arena, F.; Frusteri, F. A basic assessment of the reactivity of Ni catalysts in the decomposition of methane for the production of “COx-free” hydrogen for fuel cells application. Catal. Today 2006, 116, 298–303. [Google Scholar] [CrossRef]
  41. Frusteri, F.; Spadaro, L.; Arena, F.; Chuvilin, A. TEM evidence for factors affecting the genesis of carbon species on bare and K-promoted Ni/MgO catalysts during the dry reforming of methane. Carbon 2002, 40, 1063–1070. [Google Scholar] [CrossRef]
  42. Branca, C.; Frusteri, F.; Magazzù, V.; Mangione, A. Characterization of carbon nanotubes by TEM and infrared spectroscopy. J. Phys. Chem. B 2004, 108, 3469–3473. [Google Scholar] [CrossRef]
Figure 1. X-ray diffraction patterns of fresh and used (a) Ni/x-Zr (x = 0, La2O3) and (b) Ni-Ce/x-Zr catalysts.
Figure 1. X-ray diffraction patterns of fresh and used (a) Ni/x-Zr (x = 0, La2O3) and (b) Ni-Ce/x-Zr catalysts.
Catalysts 09 00473 g001
Figure 2. N2 adsorption–desorption isotherms of fresh Ni/x-Zr and Ni-Ce/x-Zr (x = 0, La2O3) catalysts.
Figure 2. N2 adsorption–desorption isotherms of fresh Ni/x-Zr and Ni-Ce/x-Zr (x = 0, La2O3) catalysts.
Catalysts 09 00473 g002
Figure 3. TPR profiles of fresh Ni/x-Zr and Ni-Ce/x-Zr (x = 0, La2O3) catalysts.
Figure 3. TPR profiles of fresh Ni/x-Zr and Ni-Ce/x-Zr (x = 0, La2O3) catalysts.
Catalysts 09 00473 g003
Figure 4. Patterns of fresh Ni/x-Zr and Ni-Ce/x-Zr (x = 0, La2O3) catalysts.
Figure 4. Patterns of fresh Ni/x-Zr and Ni-Ce/x-Zr (x = 0, La2O3) catalysts.
Catalysts 09 00473 g004
Figure 5. Spectra in the “fresh” (a) Ni 2p, (b) Zr 3d and “used” (c) O 1s regions of Ni/La-Zr, Ni/Zr and Ni-Ce/Zr catalysts.
Figure 5. Spectra in the “fresh” (a) Ni 2p, (b) Zr 3d and “used” (c) O 1s regions of Ni/La-Zr, Ni/Zr and Ni-Ce/Zr catalysts.
Catalysts 09 00473 g005
Figure 6. (a) CH4 and (b) CO2 conversion and (c) H2:CO ratio versus time on stream (TOS) for Ni/x-Zr (x = 0, La2O3) catalysts at TR = 700 °C; F/W = 133.33 mL/min.gcat). Total flow rate = 40 mL/min (CH4 = 17 mL/min, CO2 = 17mL/min, N2 = 6 mL/min).
Figure 6. (a) CH4 and (b) CO2 conversion and (c) H2:CO ratio versus time on stream (TOS) for Ni/x-Zr (x = 0, La2O3) catalysts at TR = 700 °C; F/W = 133.33 mL/min.gcat). Total flow rate = 40 mL/min (CH4 = 17 mL/min, CO2 = 17mL/min, N2 = 6 mL/min).
Catalysts 09 00473 g006
Figure 7. Initial rate of CO2 conversion (TR = 700 °C @ 20 min) as a function of MSA.
Figure 7. Initial rate of CO2 conversion (TR = 700 °C @ 20 min) as a function of MSA.
Catalysts 09 00473 g007
Figure 8. Deactivation factor (DF) of Ni/x-Zr (x = 0, La2O3) catalysts, calculated as: [(Initial CH4 conversion – final CH4 conversion)/initial conversion of CH4] (Initial = 20 min; Final = 400 min at TR = 700 °C).
Figure 8. Deactivation factor (DF) of Ni/x-Zr (x = 0, La2O3) catalysts, calculated as: [(Initial CH4 conversion – final CH4 conversion)/initial conversion of CH4] (Initial = 20 min; Final = 400 min at TR = 700 °C).
Catalysts 09 00473 g008
Figure 9. (a) TPO profiles of “spent” Ni/x-Zr and Ni-Ce/x-Zr (x = 0, La2O3) catalysts after reaction at TR = 700 °C and (b) TGA profiles of “spent” Ni/x-Zr and Ni-Ce/x-Zr (x = 0, La2O3) catalysts after reaction at TR = 700 °C.
Figure 9. (a) TPO profiles of “spent” Ni/x-Zr and Ni-Ce/x-Zr (x = 0, La2O3) catalysts after reaction at TR = 700 °C and (b) TGA profiles of “spent” Ni/x-Zr and Ni-Ce/x-Zr (x = 0, La2O3) catalysts after reaction at TR = 700 °C.
Catalysts 09 00473 g009
Figure 10. Micrographs of the “spent” Ni/Zr sample (TR = 700 °C; reaction time = 400 min): (A) cluster of CNTs; (B) metal sintering; (C) encapsulation of metal species; (D) and (E) metal particles captured inside CNTs.
Figure 10. Micrographs of the “spent” Ni/Zr sample (TR = 700 °C; reaction time = 400 min): (A) cluster of CNTs; (B) metal sintering; (C) encapsulation of metal species; (D) and (E) metal particles captured inside CNTs.
Catalysts 09 00473 g010
Figure 11. Micrographs of the “spent” Ni/La-Zr sample (TR = 700 °C; reaction time = 400 min): (A) CNTs over catalyst surface; (B) metal particles sintering; (C) influence of Ce addition.
Figure 11. Micrographs of the “spent” Ni/La-Zr sample (TR = 700 °C; reaction time = 400 min): (A) CNTs over catalyst surface; (B) metal particles sintering; (C) influence of Ce addition.
Catalysts 09 00473 g011
Table 1. Properties and H2 consumption during reduction of NiO species of fresh Ni/x-Zr and Ni-Ce/x-Zr (x = 0, La2O3) catalysts (calcined at 600 °C).
Table 1. Properties and H2 consumption during reduction of NiO species of fresh Ni/x-Zr and Ni-Ce/x-Zr (x = 0, La2O3) catalysts (calcined at 600 °C).
SAMPLEBET
(m2/g)
P.V.
(cm3/g)
P.D.
(nm)
H2 Consumption
(µmol/g)
La-Zr
Ni/Zr
67
39
0.247
0.111
4.0
11.3
-
1049.3
Ni/La-Zr620.24615.8814.2
Ni-Ce/Zr410.13913.3945.5
Ni-Ce/La-Zr650.24215.0782.2
Table 2. Data of H2-TPD measurements.
Table 2. Data of H2-TPD measurements.
SAMPLEH2 uptake
(µmol/gcat)
MSA a
(m2/gcat)
DNib
(%)
dNic
(nm)
Ni/Zr107.38.425.24.0
Ni/La-Zr253.319.859.61.7
Ni-Ce/Zr173.513.640.82.5
Ni-Ce/La-Zr182.514.342.92.4
a Metal surface area. b Nickel dispersion. c Average Ni particle diameter.

Share and Cite

MDPI and ACS Style

Al-Fatesh, A.S.; Arafat, Y.; Ibrahim, A.A.; Kasim, S.O.; Alharthi, A.; Fakeeha, A.H.; Abasaeed, A.E.; Bonura, G.; Frusteri, F. Catalytic Behaviour of Ce-Doped Ni Systems Supported on Stabilized Zirconia under Dry Reforming Conditions. Catalysts 2019, 9, 473. https://doi.org/10.3390/catal9050473

AMA Style

Al-Fatesh AS, Arafat Y, Ibrahim AA, Kasim SO, Alharthi A, Fakeeha AH, Abasaeed AE, Bonura G, Frusteri F. Catalytic Behaviour of Ce-Doped Ni Systems Supported on Stabilized Zirconia under Dry Reforming Conditions. Catalysts. 2019; 9(5):473. https://doi.org/10.3390/catal9050473

Chicago/Turabian Style

Al-Fatesh, Ahmed Sadeq, Yasir Arafat, Ahmed Aidid Ibrahim, Samsudeen Olajide Kasim, Abdulrahman Alharthi, Anis Hamza Fakeeha, Ahmed Elhag Abasaeed, Giuseppe Bonura, and Francesco Frusteri. 2019. "Catalytic Behaviour of Ce-Doped Ni Systems Supported on Stabilized Zirconia under Dry Reforming Conditions" Catalysts 9, no. 5: 473. https://doi.org/10.3390/catal9050473

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