Next Article in Journal
Maximizing Thermal Energy Recovery from Drinking Water for Cooling Purpose
Next Article in Special Issue
Syngas Production via CO2 Reforming of Methane over SrNiO3 and CeNiO3 Perovskites
Previous Article in Journal
Route Profile Dependent Tram Regenerative Braking Algorithm with Reduced Impact on the Supply Network
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Hydrogen Yield from CO2 Reforming of Methane: Impact of La2O3 Doping on Supported Ni Catalysts

1
Chemical Engineering Department, College of Engineering, King Saud University, P.O. Box 800, Riyadh 11421, Saudi Arabia
2
Department of Chemical and Process Engineering, University of Canterbury, 20 Kirkwood Avenue, Upper Riccarton, Christchurch 8041, New Zealand
3
King Abdullah City for Atomic & Renewable Energy, Energy Research & Innovation Center (ERIC) in Riyadh, Riyadh 11451, Saudi Arabia
*
Author to whom correspondence should be addressed.
Energies 2021, 14(9), 2412; https://doi.org/10.3390/en14092412
Submission received: 2 March 2021 / Revised: 16 April 2021 / Accepted: 20 April 2021 / Published: 23 April 2021
(This article belongs to the Special Issue Towards Greenhouse Gas Mitigation: Novelty in Heterogeneous Catalysis)

Abstract

:
Development of a transition metal based catalyst aiming at concomitant high activity and stability attributed to distinguished catalytic characteristics is considered as the bottleneck for dry reforming of methane (DRM). This work highlights the role of modifying zirconia (ZrO2) and alumina (Al2O3) supported nickel based catalysts using lanthanum oxide (La2O3) varying from 0 to 20 wt% during dry reforming of methane. The mesoporous catalysts with improved BET surface areas, improved dispersion, relatively lower reduction temperatures and enhanced surface basicity are identified after La2O3 doping. These factors have influenced the catalytic activity and higher hydrogen yields are found for La2O3 modified catalysts as compared to base catalysts (5 wt% Ni-ZrO2 and 5 wt% Ni-Al2O3). Post-reaction characterizations such as TGA have showed less coke formation over La2O3 modified samples. Raman spectra indicates decreased graphitization for La2O3 catalysts. The 5Ni-10La2O3-ZrO2 catalyst produced 80% hydrogen yields, 25% more than that of 5Ni-ZrO2. 5Ni-15La2O3-Al2O3 gave 84% hydrogen yields, 8% higher than that of 5Ni-Al2O3. Higher CO2 activity improved the surface carbon oxidation rate. From the study, the extent of La2O3 loading is dependent on the type of oxide support.

1. Introduction

The decrease of fossil fuel energy and the dilemma of environmental pollution urged a large number of researchers to maximize the conversions of methane and carbon dioxide into useful products such as hydrogen. Hydrogen is a benign source of energy. It is mainly obtained from biomass pyrolysis and thermal reforming. Methane, the main component of natural gas, can be obtained from various resources like shale gas and the fracking process, which has increased the availability of natural gas from infrequent deposits [1,2]. Moreover, the utilization of biogas is gaining momentum in recent years [3,4]. In the field of heterogeneous catalysis, particularly, in the latest decades, dry reforming of methane is regarded as one of the best prospective ways of conversion [5,6,7]. However, the dry reforming reaction as shown in Equation (1) is highly endothermic and thus requires high reaction temperatures. The process produces synthesis gas that has an appropriate ratio of H2 to CO suitable for Fischer–Tropsch synthesis [8]. Steam reforming of methane remains the best available industrial process for generating synthesis gas [9,10]. The requirement and the utilization of synthesis gas production are continuously increasing [11]. During methane dry reforming (DRM), CO2 is employed as an oxidant, which draws the interest and the likelihood of seizing and recycling CO2 from the exhaust flue gases of industrial and power plants. DRM is presently not industrially applied because of the heavy coking and sintering of the catalysts in high-temperature reforming reactions [12]. Thus, it is essential to find an innovative catalyst that endures sintering and coking. The sintering of the metallic phase and carbon formation on the surface of the catalyst, which causes the deactivation, originate from operating conditions that facilitate the side reactions, which involve the cracking of the CH4 Equation (2) and the reverse Boudouard reaction Equation (3) resulting from the combinations of CO [13].
CH 4 + CO 2   2 H 2 + 2 CO
CH 4   2 H 2 + C
CO 2 + C   CO + CO
H2 yield through DRM is significantly affected by dissociation of CH4 over Ni-supported surface and the gasification of carbon formed by CO2. The reverse water gas shift reaction and the action of the H2 spillover on the surface affect the H2 yield substantially. The hydrogen spill over enhances hydroxyl formation and catalytic activity toward CO oxidation at the metal/oxide interface. The hydroxyl groups at the metal/support interface react with CO to produce CO2. Similarly, the reverse water gas shift (CO2 + H2O → CO + H2) reduces the hydrogen; hence, the two phenomena involve the depletion of hydrogen and in turn influence the hydrogen yield. Noble metals like Pd, Ir, Pt, and Rh provide the extremely good performance of activity and stability but they are rare and expensive [14,15]. Transition metals like Ni are suitable alternatives for this reaction as they are stable and environmentally friendly [16,17]. Nonetheless, Ni-based catalysts are hampered by poor activity due to coke formations and sintering [18,19]. Hence, the major challenge is to come up with Ni catalysts capable of resisting carbon formation and sintering. Carbon generation may be opposed by controlling the reaction kinetics using suitable catalysts with proper constituents and supports. It has been established that metal–support interactions can alter both catalyst activity and activity maintenance. Bradford and Vannice had shown that support decoration of metal particle surfaces shattered big ensembles of metal atoms that served as active sites for carbon deposition and the sites in the metal–support interfacial region enhanced catalyst activity [20]. The choice of support type plays a vital role in DRM. Supports that possess O2 species at the surface of the catalyst help the carbon oxidation on the metal [21,22]. The balance between the rate of methane decomposition and the rate of carbon gasification regulates the catalyst stability [23]. The preparation of mesoporous oxides, like Al2O3, possessing high surface area and precise pores have revealed motivating results to disperse the metal over the support structure [24,25,26]. Bian et al., in their investigation of Ni supported home-made mesoporous alumina in methane dry reforming, indicated that the formation of NiAl2O4 spinel is advantageous to activity and stability towards DRM reaction [27].
Newnham et al. synthesized nanostructured Ni-incorporated mesoporous alumina with various Ni loadings by hydrothermal method and tested them as catalysts for CO2 reforming of methane [28]. The result displayed excellent stability when 10%Ni was used due to the fact that the Ni nanoparticles in these catalysts being highly stable towards migration/sintering under the reaction conditions. The presence of strong Ni–support interaction and/or active metal particles being confined to the mesoporous channels of the support. Al-Fatesh et al. studied dry reforming of methane using a series of nickel-based catalysts supported on γ-alumina promoted by B, Si, Ti, Zr, Mo, and W [5]. They concluded that the promoters enhanced the interaction between NiO and γ-alumina support and, hence, Ni dispersion and stability. On the other hand, ZrO2 is prominent support with high thermal stability able to go through alteration in their acid–basic sites [29]. Numerous studies have displayed that the presence of ZrO2 improves the thermal resistance, redox properties, oxygen storage capacity and gasification of deposited carbon [30,31]. Hu et al. examined the dry reforming of methane over Ni/ZrO2 catalysts prepared via decomposition of nickel precursor under the influence of dielectric barrier discharge (DBD) plasma at ~150 °C [32]. It was found to improve activities due to the exposition of Ni (111) facets, smaller metal particles, and more tetragonal zirconia with increased oxygen vacancies. In the course of dry reforming of methane, oxygen species over the catalyst surface affected the catalytic performance and carbon deposition. Zhang et al. studied the effects of the surface adsorbed oxygen species tuned by doping with metals like La, Ce, Sm, and Y on the catalytic behavior [33]. Their results confirmed that the surface adsorbed oxygen species promoted both CO2 activation and CH4 dissociation. Doping La2O3 in supported Ni catalysts favor the CO2 adsorption on the surface of the catalyst [34], alters the chemical and electronic state of Ni at the interface with the support and decreases the chemical interaction between Ni and the support causing the intensification of reducibility and higher dispersion of nickel [35]. Tran et al. studied the enhancement of La2O3 in the physicochemical features of cobalt supported over alumina for DRM using different temperatures and feed compositions [36]. Their results displayed that the La2O3 improved the H2 activation; enriched oxygen vacancy and lowered the apparent activation energy of CH4 consumption. The work of Lui et al. elaborated the promotional effects of La, Al, and Mn on Fe-modified clay supported by Ni catalysts used for dry reforming of methane [37]. The result of adding La, Al, and Mn altered the surface area, the basicity of the catalysts, and produced a smaller metallic Ni size. Moreover, Yabe et al. performed dry reforming of methane using several transition metals supported on ZrO2 catalysts [38]. Their results exhibited high activity and low carbon deposition upon using 1 wt%Ni/10 mol%La-ZrO2 catalysts.
In the present work, we assess the effect of the lanthanum oxide as a textual promoter of alumina and zirconia supports over Ni catalysts in the catalytic reforming of CH4 with CO2. The impact of different loadings of lanthanum oxide will be examined and their influence on the hydrogen yield. The output data will be further associated with the characterization results of BET, XRD, TPR, TEM, TPD, and TGA before and after the reaction. The difference in the basic supports originating from alumina and zirconia in terms of sintering and coking will be explored.

2. Materials and Methods

2.1. Materials

Mesoporous γ-Al2O3 (purity 99.99%, purchased from Norton Co (New York, NY, USA)), precursor of zirconium oxide (ZrOCl2·8H2O, purity >99.0%, purchased from Fluka Chemika (Washington, DC, USA)), precursors of La2O3 and Ni, respectively, (La(NO3)3·6H2O and Ni(NO3)2·6H2O purchased from Sigma Aldrich (St. Louis, MO, USA)), double distilled water.

2.2. Catalyst Preparation

In the case of zirconium-based support (for 5Ni-10La-ZrO2), the required amounts of zirconia (2.62 g) in the form of ZrOCl2·8H2O was ground completely and poured into the empty crucible. Then, the desired weights of La(NO3)3·6H2O (0.265 g) and that of Ni(NO3)2·6H2O (0.25 g) were added to the crucible containing the support to get a powder mixture. The mixture was ground well in the crucible to obtain a homogenous mixture. Purified water was poured slowly to the mixture to produce a paste while mixing. The paste was set to evaporate under room temperature condition until it dried. Thereafter, the dried sample was calcined at 700 °C for 3 h. The obtained catalysts were denoted as 5Ni-ZrO2 for 5 wt% Ni supported over ZrO2, and 5Ni-xLa-ZrO2, where x = 10, 15, 20 wt%.
In the case of alumina-based support, the above-mentioned procedure was adopted by replacing zirconia with the required amounts of mesoporous γ-Al2O3. The obtained catalysts were denoted as 5Ni-Al2O3 for 5 wt% Ni supported over Al2O3 and 5Ni-xLa-Al2O3, where x = 10, 15, 20 wt%.

2.3. Catalyst Characterization and Activity

The catalyst activity test and characterization are described in detail in the supplementary information.

3. Results and Discussion

The surface texture of the catalysts was assessed via the nitrogen adsorption–desorption isotherms. Figure 1A shows the nitrogen adsorption isotherms of the fresh catalysts (5Ni-xLa2O3-ZrO2, x = 0, 10, 15, and 20 wt%) and according to IUPAC labelling, catalysts are showing type IV isotherm with capillary condensation appearing at a relative pressure below the saturation pressure, and H1 hysteresis loop. These features are associated with mesoporous materials that have cylindrical pore geometry with narrow size distribution as well as relatively high uniformity [39]. Figure 1B exhibits the nitrogen adsorption isotherms of the fresh catalysts (5Ni-x%La-Al2O3). All samples indicate typical type IV adsorption/desorption isotherms with H1 hysteresis loop. Point of inflection at a relative pressure in the range of 0.6–0.75 corresponds to the capillary condensation which indicates the uniformity of the pores in mesoporous material [40,41].The textural properties of the catalyst are given in Table S1 of the supplementary.
The reducibility of 5%Ni-x% La2O3-ZrO2 and 5%Ni-x% La2O3-Al2O3 (x = 0, 10, 15, and 20) catalysts were examined by TPR and the patterns are shown in Figure 2. For the 5%Ni-x% La2O3-ZrO2 catalysts (Figure 2A), two prominent reduction maxima are detected over the entire temperature range, which may be attributed to the reduction of different NiO species (NiO→Ni0). The first reduction peak situated in region I at Tmax = 298 °C could be ascribed to the reduction of free NiO that is not attached to the support, and hence reduces easily at low temperature. Further, the peak in the region II at Tmax = 468 °C is allotted to the NiO reduction, attached to ZrO2 by a moderately strong link and its reduction requires higher thermal energy. After support modification by means of La2O3, substantial variations in reduction kinetics were detected. Reduction maxima illustrating NiO reduction for the higher temperature peaks has shifted towards lower temperatures [42]. In the case of Figure 2B, the 5% Ni-x% La2O3-Al2O3 catalysts (x = 0, 10, 15, and 20) display a single peak in region III at higher temperatures. The un-promoted 5%Ni-Al2O3 catalyst shows a broad peak at 750 °C indicating that the reduced NiO is strongly attached to the support. When the support is modified with the addition of different loadings of La2O3, peaks of relatively lower areas appear at high temperatures. The La2O3 modified support catalyst is shifted to higher temperatures [43]. The highest La2O3 loading catalyst gives the highest peak shift. This means that the addition of La2O3 increases further the interaction between the NiO and the modified support. The quantitative analysis of H2 consumption during H2-TPR is displayed in Table S2 of the supplementary.
Figure 3 presents the powder X-ray diffraction patterns of the calcined 5Ni-xLa2O3-ZrO2 and 5Ni-xLa2O3-Al2O3 catalysts (x = 0, 10, 15, and 20 wt%). Figure 3A displays the XRD patterns for 5Ni-xLa2O3-ZrO2. There are no peaks attributable to La2O3 in these patterns since similar patterns are obtained for La2O3 modified and un-modified catalysts denoting the homogenous distribution of La2O3. The peaks at 28.3° and 31.6° are attributed to NiO phase (JCPDS No. 47–1049). The ZrO2 support in Figure 3A has two crystalline phases, tetragonal zirconia (t-ZrO2) and monoclinic zirconia (m-ZrO2). The peaks at 25.5° and 34.3° are ascribed to m-ZrO2 [44,45], while the peaks at 40.9°, 50.2°, and 55.5° are credited to t-ZrO2 [46]. In Figure 3B, there is no peak ascribable to La2O3 in the patterns and therefore La2O3 is well dispersed in the alumina matrix. The characteristics peaks at 37.3° (311), 45.6° (400), and 67.0° (440), all are correspondingly allocated to the Al2O3 structure (JCPDS 10–0425) [47,48].
Figure 4 displays the hydrogen yield versus time on stream for the dry reforming reaction at 700 °C. The impact of La2O3 addition on the DRM catalytic performance of Ni-ZrO2 is discussed in this section. The initial hydrogen yield of the 5Ni-ZrO2 catalyst in Figure 4 is lower than the La2O3 modified catalysts. An evident trend is noted when La2O3 is added causing a significant improvement in the DRM performance. The improvement profile escalates as the following: 5Ni-ZrO2 < 5Ni-20La2O3-ZrO2 < 5Ni-15La2O3-ZrO2 < 5Ni-10La2O3-ZrO2. The highest hydrogen yield of about 80% is recorded using 10% La2O3. The improvement due to the La2O3 addition is attributed to the fact that La2O3 increases the dispersion of Ni particles on the supports and reduces the agglomeration of Ni particles during the reforming reaction as depicted in Figure 2A. Moreover, La2O3 increases the basicity as shown in the TPD profiles and therefore adsorb and react with CO2 to form La2O2CO3 species on the surface of catalyst which can speed up the conversion [49].
In Figure 5, the hydrogen yield obtained is close for La2O3 doped and non-doped catalysts. The La2O3 addition improved marginally the hydrogen yield in the following manner 5Ni-Al2O3 < 5Ni-20La2O3-Al2O3 < 5Ni-10La2O3-Al2O3 < 5Ni-15La2O3-Al2O3. The 15% La2O3 gave the highest hydrogen yield of 84%. It can be inferred that the effect of La2O3 loading affects differently the hydrogen yield productivity depending on the type of the support. Table 1 describes the efficiency of the present work and some of the literature.
Figure 6 exhibits the CO2-TPD profiles of the (A) 5Ni-xLa2O3-ZrO2 and (B) 5Ni-xLa2O3-Al2O3 (x = 0, 10, 15, and 20 wt%) spent catalysts obtained at 700 °C reaction temperature. This was performed to scan the surface basicity of the catalysts, which plays a vital role in the catalytic DRM reaction [54]. It is commonly understood that a greater desorption temperature of CO2 reveals a stronger basicity and a bigger amount of CO2 desorption although signifying that more basic sites are presented on the surface of catalyst [55]. In Figure 6A, three chief CO2 desorption peaks were identified in the experimented temperature range from 50 to 700 °C for higher loadings of La2O3 (5Ni-15La2O3-ZrO2 and 5Ni-20La2O3-ZrO2) modified catalysts and only two peaks for the un-modified (5Ni-ZrO2) and lower loading La2O3 (5Ni-10La2O3-ZrO2) modified catalysts, which denoted that three types of basic sites existed in the 5Ni-xLa2O3 + ZrO2 (x = 0, 10, 15, and 20 wt%) catalysts. The CO2 desorption peaks appeared at 80 °C, 260 °C and 550 °C. The peaks correspond to the weak adsorption of CO2 on OH groups, moderate adsorption of CO2 and strong CO2 adsorption on the metal–oxygen pairs and O2− anions, respectively [56,57]. It is clear that CO2 was favorably absorbed on the strong basic sites as the support was modified with La2O3, rather than bare zirconia support. This result showed that the adsorption of CO2 had altered from physical adsorption to chemical adsorption because of the addition of La2O3. Figure 6B shows two main CO2 desorption peaks at 80 °C and 250 °C corresponding to weak and moderate basic sites. Hence, the addition of La2O3 to the support did give significant variation in basicity of the 5Ni-xLa2O3 + Al2O3 catalysts. The 5Ni-15La2O3-Al2O3 displays larger peak intensity than the remaining catalysts, which is in accordance with the better activity observed. Table 2 shows a summary of a quantitative assessment of CO2 adsorption of the spent catalysts via CO2-TPD for ZrO2 and Al2O3 supported catalysts.
Figure 7A,B contain images obtained from Energy-dispersive X-ray spectroscopy (EDX) analysis of the fresh samples of both 5Ni-10La2O3-ZrO2 and 5Ni-15La2O3-Al2O3. These results show the elemental composition of the as-prepared catalyst samples. First, the analysis confirmed the presence of all the elemental constituents that were mixed together during the catalyst synthesis. Moreover, the percentage loadings revealed by the EDX analysis are virtually the same as intended in the calculation and catalyst preparation with error values of 10% for Ni and 6% for La.
Broad examination of the morphology of catalysts was performed via TEM. Typical TEM overviews of the fresh catalysts (Figure 8A,C) and spent catalyst samples (Figure 8B,D) were acquired after DRM at 700 °C for 420 min time on stream. There is a difference in the morphology of the fresh and used catalysts. The fresh catalyst seems to have particles clumped on the surface, while the spent catalysts depict rough and dispersed particles on the catalyst surface. In addition to the recognized metal particles, carbon in the form of nanotubes can be seen in the images of the used catalysts. The image of the fresh 5Ni-10La2O3-ZrO2 catalyst is shown in Figure 8A. It is evident from the TEM images that the Ni is homogeneously dispersed over the surface of the support. In contrast to its spent sample as shown in Figure 8B, big quantities of coke are formed commonly in the form of nanotubes. The TEM images showed well dispersed Ni species ((average particle diameter of 4 nm) over La2O3-Al2O3 support (Figure 8C). However, after the reaction (Figure 8D) size of Ni species has grown (average particle diameter of 7–8 nm).
The Fourier Transform Infrared Spectroscopy (FTIR) analysis of fresh 5Ni-ZrO2, 5Ni-10La2O3-ZrO2, 5Ni-Al2O3 and 5Ni-15La2O3-Al2O3 catalysts were done to investigate the existing bonds within the catalyst system. The infrared spectra of absorption for these samples are shown in Figure 9A,B.
In Figure 9A, Al2O3 is well known for its tendency to adsorb moisture from the atmosphere onto itself. Thus, the distinct band representing the stretching vibration of O-H, within the wavelengths 3430–3460 cm−1 for all the samples can be ascribed to [OH]−1 groups interaction and/or physisorbed moisture interaction that is adsorbed onto the Al2O3 support [58]. It is noticeable from the figure that the extent of hydration of the catalyst surface changed with metal addition as the OH band’s intensity decreased on adding active metal. Moreover, the vibration bands centered at around 1400, 1520, 1649, and 2366 cm−1 can be seen for the support and the other samples. This implies that these bands can be associated with the support, the Al–O bond stretching to be specific [59]. The small less noticeable peaks appearing at wavelengths 1237 and 1719 cm−1 for the 5Ni-Al2O3 sample can be said to be the stretching vibration of NiO and NiAl2O4 species, respectively. The latter is thought to have stronger interaction with the support, in light of that it appeared at a higher wavelength.
As for the ZrO2 support and ZrO2 supported catalysts (Figure 9B); the band within 3357–3440 cm−1 can be said to be hydrogen-bonded bending and stretching of the OH groups as a result of adsorbed moisture while the peaks at 1632–1640 cm−1 can be assigned to the vibration of the water molecules [60]. These peaks are seen to decrease in intensity with the addition of nickel and lanthanum oxide to the support. The peaks that are centered at around 457 and 752 cm−1 represent the asymmetric stretching of the Zr–O–Zr bond [61]. The vibration owing to the presence of La2O3 was not discovered. This further supports the results obtained from the XRD analysis of samples with La2O3.
Figure 10A displays TGA profiles of spent 5Ni-xLa2O3 + ZrO2 (x = 0, 10, 15, and 20 wt%) catalysts operated at 700 °C. The weight loss above 500 °C was due to the removal of deposited carbon. The extents of carbon deposition on the spent catalysts exhibit the following sequence: 5Ni-10La2O3-ZrO2 < 5Ni-15La2O3-ZrO2 < 5Ni-20La2O3-ZrO2 < 5Ni- ZrO2. The un-promoted ZrO2 catalyst gives the highest weight loss of 47.4%. The results are well established with the catalytic performance. Similarly, Figure 9B shows the TGA profiles of spent 5Ni-xLa2O3 + Al2O3 (x = 0, 10, 15, and 20 wt%) catalysts at 700 °C. The amounts of weight loss are close to each between the La2O3 promoted and non-promoted catalysts. Since values range between 9.0% for 5Ni-10La2O3-Al2O3 and 17.2% for 5Ni-ZrO2, this indicates the Al2O3 supported catalysts are more resistant to carbon deposition than ZrO2 supported catalysts.
Figure 11 shows the Raman analysis of the used catalysts (5Ni-ZrO2, 5Ni-10La2O3-ZrO2, 5Ni-Al2O3, and 5Ni-15La2O3-Al2O3). The D (deformation) and G (graphitic) bands appear at nearly 1342 and 1580 cm−1 respectively except for spent 5Ni-Al2O3 with D and G bands appearing at about 1468 and 1532 cm−1, respectively. The spent catalysts are characterized by carbon deposits of different degree of graphitization. It is established that carbon deposits having high IG to ID ratio show better extent of graphitization [62]. From the figure, it can be seen that 5Ni-ZrO2 had the highest degree of graphitization followed by 5Ni-Al2O3. Moreover, it can be inferred that the graphitization decreases with the addition of La2O3. Thus, La2O3 promotes the formation carbons that are defective.

4. Conclusions

This paper elucidates the role of La2O3 addition to two different supports (ZrO2 and Al2O3) of Ni-based catalysts for the hydrogen production via CO2 reforming of methane. The 5Ni-10La2O3-ZrO2 catalyst increased the hydrogen yield by 25% in comparison with the pristine 5Ni-ZrO2. Similarly, the 5Ni-15La2O3-Al2O3 catalyst showed improved efficiency of 8% of hydrogen yield. The La2O3 loading influenced differently the ZrO2 and the Al2O3 supports. The study displayed that the modified La2O3-Al2O3 support catalysts gave a higher hydrogen yield than La2O3-ZrO2 supported catalyst. The catalyst characterizations showed that La2O3 addition improved specific surface areas, dispersion, reducibility, metal–support interaction, and surface basic sites which contributed towards the enhanced hydrogen yield. The qualitative and quantitative analysis of carbon formed over the spent catalysts using TEM, TGA, and Raman spectroscopy showed presence of carbon nanotubes. This work provides an insight towards the role of support modification during DRM.

Supplementary Materials

The following are available online at https://www.mdpi.com/article/10.3390/en14092412/s1, Table S1: Textural properties of different catalysts supported Ni catalysts: BET specific surface area (SBET), pore volume (PV), and pore diameter (DP), Table S2: The quantitative analysis of H2 consumption during H2-TPR.

Author Contributions

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

Funding

This research was funded by Deanship of Scientific Research at King Saud University so the authors would like to express their sincere appreciation to the Deanship of Scientific Research at King Saud University for funding this research project (#RG-119).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The authors would like to express their sincere appreciation to the Deanship of Scientific Research at King Saud University for funding this research project (#RG-119).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Wang, Q.; Chen, X.; Jha, A.N.; Rogers, H. Natural gas from shale formation—The evolution, evidences and challenges of shale gas revolution in United States. Renew. Sustain. Energy Rev. 2014, 30, 1–28. [Google Scholar] [CrossRef]
  2. Montgomery, C.T.; Smith, M.B. Hydraulic Fracturing: History of an Enduring Technology. J. Pet. Technol. 2010, 62, 26–40. [Google Scholar] [CrossRef]
  3. Themelis, N.J.; Ulloa, P.A. Methane generation in landfills. Renew. Energy 2007, 32, 1243–1257. [Google Scholar] [CrossRef]
  4. Izquierdo, U.; Barrio, V.; Requies, J.; Cambra, J.; Güemez, M.; Arias, P. Tri-reforming: A new biogas process for synthesis gas and hydrogen production. Int. J. Hydrog. Energy 2013, 38, 7623–7631. [Google Scholar] [CrossRef]
  5. Al-Fatesh, A.S.; Kumar, R.; Kasim, S.O.; Ibrahim, A.A.; Fakeeha, A.H.; Abasaeed, A.E.; Alrasheed, R.; Bagabas, A.; Chaudhary, M.L.; Frusteri, F.; et al. The effect of modifier identity on the performance of Ni-based catalyst supported on γ-Al2O3 in dry reforming of methane. Catal. Today 2020, 348, 236–242. [Google Scholar] [CrossRef]
  6. Ibrahim, A.A.; Al-Fatesh, A.S.; Khan, W.U.; Kasim, S.O.; Abasaeed, A.E.; Fakeeha, A.H.; Bonura, G.; Frusteri, F. Enhanced coke suppression by using phosphate-zirconia supported nickel catalysts under dry methane reforming conditions. Int. J. Hydrog. Energy 2019, 44, 27784–27794. [Google Scholar] [CrossRef]
  7. Al-Fatesh, A.S.; Abu-Dahrieh, J.K.; Atia, H.; Armbruster, U.; Ibrahim, A.A.; Khan, W.U.; Abasaeed, A.E.; Fakeeha, A.H. Effect of pre-treatment and calcination temperature on Al2O3-ZrO2 supported Ni-Co catalysts for dry reforming of methane. Int. J. Hydrog. Energy 2019, 44, 21546–21558. [Google Scholar] [CrossRef] [Green Version]
  8. Zhang, S.; Muratsugu, S.; Ishiguro, N.; Tada, M. Ceria-Doped Ni/SBA-16 Catalysts for Dry Reforming of Methane. ACS Catal. 2013, 3, 1855–1864. [Google Scholar] [CrossRef]
  9. Boyano, A.; Morosuk, T.; Blanco-Marigorta, A.; Tsatsaronis, G. Conventional and advanced exergoenvironmental analysis of a steam methane reforming reactor for hydrogen production. J. Clean. Prod. 2012, 20, 152–160. [Google Scholar] [CrossRef]
  10. Kim, A.R.; Lee, H.Y.; Cho, J.M.; Choi, J.-H.; Bae, J.W. Ni/M-Al2O3 (M=Sm, Ce or Mg) for combined steam and CO2 reforming of CH4 from coke oven gas. J. CO2 Util. 2017, 21, 211–218. [Google Scholar] [CrossRef]
  11. Lougou, B.G.; Shuai, Y.; Chaffa, G.; Xing, H.; Tan, H.; Du, H. Analysis of CO2 utilization into synthesis gas based on solar thermochemical CH4-reforming. J. Energy Chem. 2019, 28, 61–72. [Google Scholar] [CrossRef] [Green Version]
  12. Wu, Z.; Yang, B.; Miao, S.; Liu, W.; Xie, J.; Lee, S.; Pellin, M.J.; Xiao, D.; Su, D.; Ma, D. Lattice Strained Ni-Co alloy as a High-Performance Catalyst for Catalytic Dry Reforming of Methane. ACS Catal. 2019, 9, 2693–2700. [Google Scholar] [CrossRef]
  13. Arora, S.; Prasad, R. An overview on dry reforming of methane: Strategies to reduce carbonaceous deactivation of catalysts. RSC Adv. 2016, 6, 108668–108688. [Google Scholar] [CrossRef]
  14. Varga, E.; Pusztai, P.; Steinrück, H.-P.; Óvári, L.; Papp, C.; Kónya, Z.; Oszkó, A.; Erdőhelyi, A.; Kiss, J. Probing the interaction of Rh, Co and bimetallic Rh–Co nanoparticles with the CeO2 support: Catalytic materials for alternative energy generation. Phys. Chem. Chem. Phys. 2015, 17, 27154–27166. [Google Scholar] [CrossRef] [Green Version]
  15. Pakhare, D.; Spivey, J. A review of dry (CO2) reforming of methane over noble metal catalysts. Chem. Soc. Rev. 2014, 43, 7813–7837. [Google Scholar] [CrossRef] [PubMed]
  16. Damyanova, S.; Shtereva, I.; Pawelec, B.; Mihaylov, L.; Fierro, J.L.G. Characterization of none and yttrium-modified Ni-based catalysts for dry reforming of methane. Appl. Catal. B Environ. 2020, 278, 119335. [Google Scholar] [CrossRef]
  17. Jalali, R.; Rezaei, M.; Nematollahi, B.; Baghalha, M. Preparation of Ni/MeAl2O4-MgAl2O4 (Me=Fe, Co, Ni, Cu, Zn, Mg) nanocatalysts for the syngas production via combined dry reforming and partial oxidation of methane. Renew. Energy 2020, 149, 1053–1067. [Google Scholar] [CrossRef]
  18. Liu, C.-J.; Ye, J.; Jiang, J.; Pan, Y. Progresses in the Preparation of Coke Resistant Ni-based Catalyst for Steam and CO2 Reforming of Methane. ChemCatChem 2011, 3, 529–541. [Google Scholar] [CrossRef]
  19. Wang, N.; Shen, K.; Yu, X.; Qian, W.; Chu, W. Preparation and characterization of a plasma treated NiMgSBA-15 catalyst for methane reforming with CO2 to produce syngas. Catal. Sci. Technol. 2013, 3, 2278–2287. [Google Scholar] [CrossRef]
  20. Bradford, M.C.; Vannice, M.A. The role of metal–support interactions in CO2 reforming of CH4. Catal. Today 1999, 50, 87–96. [Google Scholar] [CrossRef]
  21. Tang, M.; Liu, K.; Roddick, D.M.; Fan, M. Enhanced lattice oxygen reactivity over Fe2O3/Al2O3 redox catalyst for chemical-looping dry (CO2) reforming of CH4: Synergistic La-Ce effect. J. Catal. 2018, 368, 38–52. [Google Scholar] [CrossRef]
  22. Faria, E.; Neto, R.; Colman, R.; Noronha, F. Hydrogen production through CO2 reforming of methane over Ni/CeZrO2/Al2O3 catalysts. Catal. Today 2014, 228, 138–144. [Google Scholar] [CrossRef]
  23. Stagg-Williams, S.M.; Noronha, F.B.; Fendley, G.; Resasco, D.E. CO2 Reforming of CH4 over Pt/ZrO2 Catalysts Promoted with La and Ce Oxides. J. Catal. 2000, 194, 240–249. [Google Scholar] [CrossRef] [Green Version]
  24. Li, Z.-X.; Shi, F.-B.; Li, L.-L.; Zhang, T.; Yan, C.-H. A facile route to ordered mesoporous-alumina-supported catalysts, and their catalytic activities for CO oxidation. Phys. Chem. Chem. Phys. 2010, 13, 2488–2491. [Google Scholar] [CrossRef] [PubMed]
  25. Dehimi, L.; Benguerba, Y.; Virginie, M.; Hijazi, H. Microkinetic modelling of methane dry reforming over Ni/Al2O3 catalyst. Int. J. Hydrog. Energy 2017, 42, 18930–18940. [Google Scholar] [CrossRef]
  26. Lašič Jurković, D.; Liu, J.-L.; Pohar, A.; Likozar, B. Methane Dry Reforming over Ni/Al2O3 Catalyst in Spark Plasma Reactor: Linking Computational Fluid Dynamics (CFD) with Reaction Kinetic Modelling. Catal. Today 2021, 362, 11–21. [Google Scholar] [CrossRef]
  27. Bian, Z.; Zhong, W.; Yu, Y.; Wang, Z.; Jiang, B.; Kawi, S. Dry reforming of methane on Ni/mesoporous-Al2O3 catalysts: Effect of calcination temperature. Int. J. Hydrog. Energy 2021. [Google Scholar] [CrossRef]
  28. Newnham, J.; Mantri, K.; Amin, M.H.; Tardio, J.; Bhargava, S.K. Highly stable and active Ni-mesoporous alumina catalysts for dry reforming of methane. Int. J. Hydrog. Energy 2012, 37, 1454–1464. [Google Scholar] [CrossRef]
  29. Montoya, J.; Romero-Pascual, E.; Gimon, C.; Del Angel, P.; Monzón, A. Methane reforming with CO2 over Ni/ZrO2–CeO2 catalysts prepared by sol–gel. Catal. Today 2000, 63, 71–85. [Google Scholar] [CrossRef]
  30. Li, W.; Zhao, Z.; Jiao, Y. Dry reforming of methane towards CO-rich hydrogen production over robust supported Ni catalyst on hierarchically structured monoclinic zirconia nanosheets. Int. J. Hydrog. Energy 2016, 41, 17907–17921. [Google Scholar] [CrossRef]
  31. Jang, J.T.; Yoon, K.J.; Bae, J.W.; Han, G.Y. Cyclic production of syngas and hydrogen through methane-reforming and water-splitting by using ceria–zirconia solid solutions in a solar volumetric receiver–reactor. Sol. Energy 2014, 109, 70–81. [Google Scholar] [CrossRef]
  32. Hu, X.; Jia, X.; Zhang, X.; Liu, Y.; Liu, C.-J. Improvement in the activity of Ni/ZrO2 by cold plasma decomposition for dry reforming of methane. Catal. Commun. 2019, 128, 105720. [Google Scholar] [CrossRef]
  33. Zhang, M.; Zhang, J.; Zhou, Z.; Chen, S.; Zhang, T.; Song, F.; Zhang, Q.; Tsubaki, N.; Tan, Y.; Han, Y. Effects of the surface adsorbed oxygen species tuned by rare-earth metal doping on dry reforming of methane over Ni/ZrO2 catalyst. Appl. Catal. B Environ. 2020, 264, 118522. [Google Scholar] [CrossRef]
  34. Zhang, W.; Liu, B.; Zhu, C.; Tian, Y. Preparation of La2NiO4/ZSM-5 catalyst and catalytic performance in CO2/CH4 reforming to syngas. Appl. Catal. A Gen. 2005, 292, 138–143. [Google Scholar] [CrossRef]
  35. Horváth, A.; Stefler, G.; Geszti, O.; Kienneman, A.; Pietraszek, A.; Guczi, L. Methane dry reforming with CO2 on CeZr-oxide supported Ni, NiRh and NiCo catalysts prepared by sol–gel technique: Relationship between activity and coke formation. Catal. Today 2011, 169, 102–111. [Google Scholar] [CrossRef]
  36. Tran, N.T.; Van Le, Q.; Van Cuong, N.; Nguyen, T.D.; Phuc, N.H.H.; Phuong, P.T.; Monir, M.U.; Aziz, A.A.; Truong, Q.D.; Abidin, S.Z.; et al. La-doped cobalt supported on mesoporous alumina catalysts for improved methane dry reforming and coke mitigation. J. Energy Inst. 2020, 93, 1571–1580. [Google Scholar] [CrossRef]
  37. Liu, H.; Hadjltaief, H.B.; Benzina, M.; Gálvez, M.E.; Da Costa, P. Natural clay based nickel catalysts for dry reforming of methane: On the effect of support promotion (La, Al, Mn). Int. J. Hydrog. Energy 2019, 44, 246–255. [Google Scholar] [CrossRef]
  38. Yabe, T.; Mitarai, K.; Oshima, K.; Ogo, S.; Sekine, Y. Low-temperature dry reforming of methane to produce syngas in an electric field over La-doped Ni/ZrO2 catalysts. Fuel Process. Technol. 2017, 158, 96–103. [Google Scholar] [CrossRef]
  39. 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. [Google Scholar] [CrossRef] [Green Version]
  40. Mandal, S.; Santra, C.; Kumar, R.; Pramanik, M.; Rahman, S.; Bhaumik, A.; Maity, S.; Sen, D.; Chowdhury, B. Niobium doped hexagonal mesoporous silica (HMS-X) catalyst for vapor phase Beckmann rearrangement reaction. RSC Adv. 2014, 4, 845–854. [Google Scholar] [CrossRef]
  41. Rahman, S.; Santra, C.; Kumar, R.; Bahadur, J.; Sultana, A.; Schweins, R.; Sen, D.; Maity, S.; Mazumdar, S.; Chowdhury, B. Highly active Ga promoted Co-HMS-X catalyst towards styrene epoxidation reaction using molecular O2. Appl. Catal. A Gen. 2014, 482, 61–68. [Google Scholar] [CrossRef]
  42. Jiang, X.-Y.; Zhou, R.-X.; Pan, P.; Zhu, B.; Yuan, X.-X.; Zheng, X.-M. Effect of the addition of La2O3 on TPR and TPD of CuOγ-Al2O3 catalysts. Appl. Catal. A Gen. 1997, 150, 131–141. [Google Scholar] [CrossRef]
  43. Mo, W.; Ma, F.; Ma, Y.; Fan, X. The optimization of Ni–Al2O3 catalyst with the addition of La2O3 for CO2–CH4 reforming to produce syngas. Int. J. Hydrog. Energy 2019, 44, 24510–24524. [Google Scholar] [CrossRef]
  44. Goula, M.; Charisiou, N.; Siakavelas, G.; Tzounis, L.; Tsiaoussis, I.; Panagiotopoulou, P.; Goula, G.; Yentekakis, I. Syngas production via the biogas dry reforming reaction over Ni supported on zirconia modified with CeO2 or La2O3 catalysts. Int. J. Hydrog. Energy 2017, 42, 13724–13740. [Google Scholar] [CrossRef]
  45. Zhang, J.; Gao, Y.; Jia, X.; Wang, J.; Chen, Z.; Xu, Y. Oxygen vacancy-rich mesoporous ZrO2 with remarkably enhanced visible-light photocatalytic performance. Sol. Energy Mater. Sol. Cells 2018, 182, 113–120. [Google Scholar] [CrossRef]
  46. Titus, J.; Roussière, 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]
  47. Alipour, Z.; Rezaei, M.; Meshkani, F. Effects of support modifiers on the catalytic performance of Ni/Al2O3 catalyst in CO2 reforming of methane. Fuel 2014, 129, 197–203. [Google Scholar] [CrossRef]
  48. Mierczynski, P.; Mierczynska, A.; Ciesielski, R.; Mosinska, M.; Nowosielska, M.; Czylkowska, A.; Maniukiewicz, W.; Szynkowska, M.I.; Vasilev, K. High Active and Selective Ni/CeO2–Al2O3 and Pd–Ni/CeO2–Al2O3 Catalysts for Oxy-Steam Reforming of Methanol. Catalysts 2018, 8, 380. [Google Scholar] [CrossRef] [Green Version]
  49. Li, K.; Pei, C.; Li, X.; Chen, S.; Zhang, X.; Liu, R.; Gong, J. Dry reforming of methane over La2O2CO3-modified Ni/Al2O3 catalysts with moderate metal support interaction. Appl. Catal. B Environ. 2020, 264, 118448. [Google Scholar] [CrossRef]
  50. Khani, Y.; Bahadoran, F.; Shariatinia, Z.; Varmazyari, M.; Safari, N. Synthesis of highly efficient and stable Ni/CexZr1-xGdxO4 and Ni/X-Al2O3 (X = Ce, Zr, Gd, Ce-Zr-Gd) nanocatalysts applied in methane reforming reactions. Ceram. Int. 2020, 46, 25122–25135. [Google Scholar] [CrossRef]
  51. Lim, Z.-Y.; Tu, J.; Xu, Y.; Chen, B. Ni@ZrO2 yolk-shell catalyst for CO2 methane reforming: Effect of Ni@SiO2 size as the hard-template. J. Colloid Interface Sci. 2021, 590, 641–651. [Google Scholar] [CrossRef]
  52. Aghaali, M.H.; Firoozi, S. Enhancing the catalytic performance of Co substituted NiAl2O4 spinel by ultrasonic spray pyrolysis method for steam and dry reforming of methane. Int. J. Hydrog. Energy 2021, 46, 357–373. [Google Scholar] [CrossRef]
  53. Medeiros, R.L.; Macedo, H.P.; Melo, V.R.; Oliveira, Â.A.S.; Barros, J.M.; Melo, M.A.; Melo, D.M. Ni supported on Fe-doped MgAl2O4 for dry reforming of methane: Use of factorial design to optimize H2 yield. Int. J. Hydrog. Energy 2016, 41, 14047–14057. [Google Scholar] [CrossRef]
  54. Lavoie, J.-M. Review on dry reforming of methane, a potentially more environmentally-friendly approach to the increasing natural gas exploitation. Front. Chem. 2014, 2, 81. [Google Scholar] [CrossRef] [Green Version]
  55. Zhang, L.; Zhang, Q.; Liu, Y.; Zhang, Y. Dry reforming of methane over Ni/MgO-Al2O3 catalysts prepared by two-step hydrothermal method. Appl. Surf. Sci. 2016, 389, 25–33. [Google Scholar] [CrossRef]
  56. Wang, H.; Zhao, B.; Qin, L.; Wang, Y.; Yu, F.; Han, J. Non-thermal plasma-enhanced dry reforming of methane and CO2 over Ce-promoted Ni/C catalysts. Mol. Catal. 2020, 485, 110821. [Google Scholar] [CrossRef]
  57. Zheng, X.; Li, Y.; Zhang, L.; Shen, L.; Xiao, Y.; Zhang, Y.; Au, C.; Jiang, L. Insight into the effect of morphology on catalytic performance of porous CeO2 nanocrystals for H2S selective oxidation. Appl. Catal. B Environ. 2019, 252, 98–110. [Google Scholar] [CrossRef]
  58. Lei, H.; Zhang, P. Preparation of alumina/silica core-shell abrasives and their CMP behavior. Appl. Surf. Sci. 2007, 253, 8754–8761. [Google Scholar] [CrossRef]
  59. Sanchez, E.A.; Comelli, R.A. Hydrogen by glycerol steam reforming on a nickel–alumina catalyst: Deactivation processes and regeneration. Int. J. Hydrog. Energy 2012, 37, 14740–14746. [Google Scholar] [CrossRef]
  60. Goharshadi, E.K.; Hadadian, M. Effect of calcination temperature on structural, vibrational, optical, and rheological properties of zirconia nanoparticles. Ceram. Int. 2012, 38, 1771–1777. [Google Scholar] [CrossRef]
  61. Heshmatpour, F.; Aghakhanpour, R.B. Synthesis and characterization of superfine pure tetragonal nanocrystalline sulfated zirconia powder by a non-alkoxide sol–gel route. Adv. Powder Technol. 2012, 23, 80–87. [Google Scholar] [CrossRef]
  62. Kim, S.; Crandall, B.S.; Lance, M.J.; Cordonnier, N.; Lauterbach, J.; Sasmaz, E. Activity and stability of NiCe@SiO multi–yolk–shell nanotube catalyst for tri-reforming of methane. Appl. Catal. B Environ. 2019, 259, 118037. [Google Scholar] [CrossRef]
Figure 1. N2 adsorption–desorption isotherms for fresh (A) 5Ni-xLa2O3+ZrO2 and (B) 5Ni-x La2O3+Al2O3 catalyst calcined at 700 °C, (x = 0, 10, 15, and 20 wt%).
Figure 1. N2 adsorption–desorption isotherms for fresh (A) 5Ni-xLa2O3+ZrO2 and (B) 5Ni-x La2O3+Al2O3 catalyst calcined at 700 °C, (x = 0, 10, 15, and 20 wt%).
Energies 14 02412 g001
Figure 2. H2-TPR profiles of (A) for 5Ni-xLa2O3-ZrO2 and (B) 5Ni-xLa2O3-Al2O3 (x = 0, 10, 15, and 20 wt%) catalysts.
Figure 2. H2-TPR profiles of (A) for 5Ni-xLa2O3-ZrO2 and (B) 5Ni-xLa2O3-Al2O3 (x = 0, 10, 15, and 20 wt%) catalysts.
Energies 14 02412 g002
Figure 3. XRD patterns of the (A) 5Ni-xLa2O3+ZrO2 (B) 5Ni-xLa2O3+Al2O3 catalysts (x = 0, 10, 15, and 20 wt%) calcined at 700 °C).
Figure 3. XRD patterns of the (A) 5Ni-xLa2O3+ZrO2 (B) 5Ni-xLa2O3+Al2O3 catalysts (x = 0, 10, 15, and 20 wt%) calcined at 700 °C).
Energies 14 02412 g003
Figure 4. H2 yield vs. time on stream over the 5Ni-xLa2O3-ZrO2 (x = 0, 10, 15, and 20 wt%) catalysts at 700 °C for 420 min.
Figure 4. H2 yield vs. time on stream over the 5Ni-xLa2O3-ZrO2 (x = 0, 10, 15, and 20 wt%) catalysts at 700 °C for 420 min.
Energies 14 02412 g004
Figure 5. H2 yield vs. time on stream over the 5Ni-xLa2O3-Al2O3 (x = 0, 10, 15, and 20 wt%) catalysts at 700 °C for 420 min.
Figure 5. H2 yield vs. time on stream over the 5Ni-xLa2O3-Al2O3 (x = 0, 10, 15, and 20 wt%) catalysts at 700 °C for 420 min.
Energies 14 02412 g005
Figure 6. CO2-TPD profiles of the spent (A) 5Ni-xLa2O3-ZrO2 and (B) 5Ni-xLa2O3-Al2O3 (x = 0, 10, 15, and 20 wt%) catalysts obtained at 700 °C reaction temperature.
Figure 6. CO2-TPD profiles of the spent (A) 5Ni-xLa2O3-ZrO2 and (B) 5Ni-xLa2O3-Al2O3 (x = 0, 10, 15, and 20 wt%) catalysts obtained at 700 °C reaction temperature.
Energies 14 02412 g006
Figure 7. EDX analysis of fresh (A) 5Ni-15La2O3-Al2O3 and (B) 5Ni-10La2O3- ZrO2 showing the elemental composition of prepared catalysts.
Figure 7. EDX analysis of fresh (A) 5Ni-15La2O3-Al2O3 and (B) 5Ni-10La2O3- ZrO2 showing the elemental composition of prepared catalysts.
Energies 14 02412 g007
Figure 8. TEM image of (A) Fresh, (B) spent of 5Ni/10La2O3+ZrO2, (C) Fresh, (D) spent of 5Ni/15La2O3 + Al2O3 catalysts.
Figure 8. TEM image of (A) Fresh, (B) spent of 5Ni/10La2O3+ZrO2, (C) Fresh, (D) spent of 5Ni/15La2O3 + Al2O3 catalysts.
Energies 14 02412 g008
Figure 9. The FTIR spectra of fresh (A) Al2O3, 5Ni-Al2O3, 5Ni-15La2O3-Al2O3 and (B) ZrO2, 5Ni-ZrO2, 5Ni-10La2O3-ZrO2 showing the existing stretching vibration.
Figure 9. The FTIR spectra of fresh (A) Al2O3, 5Ni-Al2O3, 5Ni-15La2O3-Al2O3 and (B) ZrO2, 5Ni-ZrO2, 5Ni-10La2O3-ZrO2 showing the existing stretching vibration.
Energies 14 02412 g009
Figure 10. TGA profiles of spent (A) 5Ni-xLa2O3-ZrO2, (B) 5Ni-xLa2O3-Al2O3 (x = 0, 10, 15, and 20 wt%) catalysts at 700 °C.
Figure 10. TGA profiles of spent (A) 5Ni-xLa2O3-ZrO2, (B) 5Ni-xLa2O3-Al2O3 (x = 0, 10, 15, and 20 wt%) catalysts at 700 °C.
Energies 14 02412 g010
Figure 11. Raman spectra for the spent catalysts (5Ni-ZrO2, 5Ni-10La2O3-ZrO2, 5Ni-Al2O3, and 5Ni-15La2O3-Al2O3) showing the extent of graphitization of the carbon deposits.
Figure 11. Raman spectra for the spent catalysts (5Ni-ZrO2, 5Ni-10La2O3-ZrO2, 5Ni-Al2O3, and 5Ni-15La2O3-Al2O3) showing the extent of graphitization of the carbon deposits.
Energies 14 02412 g011
Table 1. Hydrogen yield performances obtained in CO2 reforming of methane of present and past work.
Table 1. Hydrogen yield performances obtained in CO2 reforming of methane of present and past work.
CatalystFeed (CH4:CO2: Inert)Reaction Temperature (°C)GHSV (mL/(g·h)Yield (%)Ref.
Ni/Zr-γAl2O31:1:375054,00063[50]
Ni@ZrO2-SiZr-7.7 1:1:180072,00081[51]
0.8% Ni+0.2% Co-MgAl2O41:1:0.570054,00051[52]
10Ni+1%Fe-MgAl2O41:1:175030,00078[53]
5Ni-10La2O3-ZrO21:1:0.3370042,00080This work
5Ni-15La2O3-Al2O31:1:0.3370042,00084This work
Table 2. Quantitative assessment of CO2 adsorption of the spent catalysts via CO2-TPD.
Table 2. Quantitative assessment of CO2 adsorption of the spent catalysts via CO2-TPD.
CatalystWeak-Basicity
(µmol/g)
Medium-Basicity
(µmol/g)
Strong-Basicity
(µmol/g)
Total-Basicity
(µmol/g)
5Ni-ZrO2 a0.520.480.001
5Ni-10La2O3-ZrO20.660.290.000.95
5Ni-15La2O3-ZrO20.630.490.121.24
5Ni-20La2O3-ZrO20.520.940.271.73
5Ni-Al2O3 a0.440.380.181
5Ni-10La2O3-Al2O30.420.110.000.53
5Ni-15La2O3-Al2O30.320.100.000.42
5Ni-20La2O3-Al2O30.350.060.000.41
a For comparison and basicity assessment of catalysts, the sum of all basic sites of 5Ni-ZrO2 is set to 1 and 5Ni-Al2O3 is set to 1.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Abasaeed, A.; Kasim, S.; Khan, W.; Sofiu, M.; Ibrahim, A.; Fakeeha, A.; Al-Fatesh, A. Hydrogen Yield from CO2 Reforming of Methane: Impact of La2O3 Doping on Supported Ni Catalysts. Energies 2021, 14, 2412. https://doi.org/10.3390/en14092412

AMA Style

Abasaeed A, Kasim S, Khan W, Sofiu M, Ibrahim A, Fakeeha A, Al-Fatesh A. Hydrogen Yield from CO2 Reforming of Methane: Impact of La2O3 Doping on Supported Ni Catalysts. Energies. 2021; 14(9):2412. https://doi.org/10.3390/en14092412

Chicago/Turabian Style

Abasaeed, Ahmed, Samsudeen Kasim, Wasim Khan, Mahmud Sofiu, Ahmed Ibrahim, Anis Fakeeha, and Ahmed Al-Fatesh. 2021. "Hydrogen Yield from CO2 Reforming of Methane: Impact of La2O3 Doping on Supported Ni Catalysts" Energies 14, no. 9: 2412. https://doi.org/10.3390/en14092412

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