Next Article in Journal
Microwaves and Heterogeneous Catalysis: A Review on Selected Catalytic Processes
Next Article in Special Issue
Effect of Ag-Addition on the Catalytic and Physicochemical Properties of Ni/ZrO2 Catalyst in Oxy-Steam Reforming of CH4 and LNG Processes
Previous Article in Journal
Co-Supported CeO2Nanoparticles for CO Catalytic Oxidation: Effects of Different Synthesis Methods on Catalytic Performance
Previous Article in Special Issue
Selective Hydrogenation of Acetylene Catalysed by a B12N12 Cluster Doped with a Single Nickel Atom: A DFT Study
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Dry Reforming of Methane Using Ce-modified Ni Supported on 8%PO4 + ZrO2 Catalysts

Chemical Engineering Department, College of Engineering, King Saud University, P.O. Box 800, Riyadh 11421, Saudi Arabia
*
Author to whom correspondence should be addressed.
Catalysts 2020, 10(2), 242; https://doi.org/10.3390/catal10020242
Submission received: 24 December 2019 / Revised: 23 January 2020 / Accepted: 28 January 2020 / Published: 18 February 2020
(This article belongs to the Special Issue Recent Advances in Nickel-Based Catalysts)

Abstract

:
Dry reforming of methane (DRM) was studied in the light of Ni supported on 8%PO4 + ZrO2 catalysts. Cerium was used to modify the Ni active metal. Different percentage loadings of Ce (1%, 1.5%, 2%, 2.5%, 3%, and 5%) were tested. The wet incipient impregnation method was used for the preparation of all catalysts. The catalysts were activated at 700 °C for ½ h. The reactions were performed at 800 °C using a gas hourly space velocity of 28,000 mL (h·gcat)−1. X-ray diffraction (XRD), N2 physisorption, hydrogen temperature programmed reduction (H2-TPR), temperature programmed oxidation (TPO), temperature programmed desorption (TPD), and thermogravimetric analysis (TGA) were used for characterizing the catalysts. The TGA analysis depicted minor amounts of carbon deposition. The CO2-TPD results showed that Ce enhanced the basicity of the catalysts. The 3% Ce loading possessed the highest surface area, the largest pore volume, and the greatest pore diameter. All the promoted catalysts enhanced the conversions of CH4 and CO2. Among the promoted catalysts tested, the 10Ni + 3%Ce/8%PO4 + ZrO2 catalyst system operated at 1 bar and at 800 °C gave the highest conversions of CH4 (95%) and CO2 (96%). The stability profile of Cerium-modified catalysts (10%Ni/8%PO4 + ZrO2) depicted steady CH4 and CO2 conversions during the 7.5 h time on stream.

1. Introduction

Dry reforming of methane (DRM) uses a favorable reaction that transforms two greenhouse gases, methane and carbon dioxide, into a synthesis gas, which is a beneficial product [1,2,3,4,5]. Moreover, DRM can be a potential process for the improvement of biogas composed largely by CO2 and CH4 [6,7]. The utilization of fossil fuels generates greenhouse gases, which contribute significantly to climate change and cause global warming [8,9,10]. The methane gas constitutes more than 80% of natural gas, and therefore, its conversion plays a vital role in the future energy supply, since it produces energy carriers such as carbonyl and dimethyl ether. The produced synthesis gas is composed of H2 and CO, which are versatile intermediates for many useful chemical products such as liquid fuels obtained via Fischer–Tropsch synthesis, light olefins, and hydrocarbons [9,11,12]. DRM can also be employed as chemical storage for renewable energy on account of its highly endothermic nature [13]. Consequently, broad studies are focused on developing a dry reforming process for syngas production. The absence of a catalyst that is stable under the reaction conditions hinders the commercialization of the dry reforming of methane process. For this reason, the utilization of an effective catalyst is essential. For instance, the use of noble metals such as Pt, Rh, Pd, Ru, and Ir distributed on support are competently used for dry reforming of CH4 [14,15,16]. Arandiyan et al. [17] investigated the effect of noble metals on the dry reforming of CH4. They tested the catalytic activity over La0.4M0.6Al0.2Ni0.8O3(M = Pt, Pd, Ru, Rh, Ir) perovskite-type oxides with a surface area of 3.26–4.14 m2/g. It was obtained that the La0.4Rh0.6Al0.2Ni0.8O3 catalyst revealed the best catalytic performance due to its high surface area, surface oxygen concentration, and good low-temperature reducibility. El Hassan et al. [18] studied the influence of Rh (0.2 and 0.5 wt %) to the Co/SBA-15 catalyst in the catalytic performance of dry reforming of CH4. It was found that Rh favored Co stabilization in the mesopores and was reduced at a much lower temperature. The nature of coke was affected and less γ-carbon formed on the Rh-containing sample. However, because of the high prices and limited availability of noble metals, the research has also focused on transition metals such as Co, Ni, Ti, which are broadly used for this process because of their high activity [19]. However, these catalysts are prone to quick deactivation due to coke deposition. Within the catalysts, Ni depicted the greatest performance due to its strong ability to break C–C/C–H bonds and activate CH4 [20]. AKri et al. [21] elaborated the dry reforming of CH4 employing atomically dispersed Ni atoms, stabilized by interaction with Ce-doped hydroxyapatite. Their results evidenced that the catalyst was very active and that isolated Ni atoms were inherently coke resistant. Ce doping of hydroxyapatite brought strong metal–support interactions which stabilized Ni single atoms toward sintering and favored the selective activation of only the first C–H bond in CH4. Goula et al. [22] investigated nickel supported on Zr modified with CeO2 or La2O3 catalysts for the dry reforming of a CH4 reaction. Although bare Zr support provided adequate activity and stability, however, the modified Zr support with Ce and La enhanced basicity and oxygen ion lability values and therefore gave better activity and stability performance. Abba et al. [23] used natural phosphate support for Ni in the reforming reaction of methane. They found that the support modification by means of an acid gave a very active and stable Ni-based catalyst. Ibrahim et al. [24] studied the dry reforming of CH4 using phosphate zirconia supported nickel catalysts (x%Ni/8%PO4-Zr, where x = 5, 10, 15 or 20). The result exhibited that activity and stability was intensely related to nickel loading, whereas the highest performance was recorded for a catalyst having a Ni loading of 10 wt %. Al-Fatesh et al. [5] investigated the dry reforming of CH4 over combined Mg and Ce on Ni-based catalysts supported on γ-Al2O3 support doped with 3.0 wt. % TiO2. Cerium and magnesium oxides strengthened the interaction between the nickel and the support, and hence the catalytic activity was substantially improved. Then again, Fatish et al. [4] reported the effect whereby Ce promoted Ni supported on lanthanum-modified zirconium oxides in the dry reforming of CH4. The result established that the reduction properties of the catalysts were considerably improved upon La2O3 modification. The lanthanum incorporated into the zirconium support and the t-ZrO2 phase was secure and highly stabilized, leading to greater catalytic performance. Remarkably, CH4 conversion increased twice, and CO2 conversion increased to 1.5 times that of pristine Ni/ZrO2. Deactivation arising from the coke deposition is the chief obstacle in the use of Ni-containing catalyst [25]; however, it can be overcome by the addition of promoters. Promoting elements play substantial roles in improving the catalyst, even though they are used in slight quantities [26]. The role of dopants on the catalytic activity for CO2 reforming of CH4 was examined on Ni/Me0.15Ce0.85O2−δ with Me = Zr4+, La3+ or Sm3+ catalysts [27]. Zr-doped catalyst gave the highest activity, while Sm and La-doped catalysts produced the lowest coke formation. The type of dopants affected the Ni–support interaction and the electronic state of the metal catalyst. Veen and Li modified the performance of Ni-catalyzed dry reforming of CH4 and carbon deposition by the use of Ni–CeO2−x interaction [28]. The result of the H2 reduction beyond 500 °C created a strong bonding between Ni and CeO2 that prevented Ni particle sintering. A reduction temperature of 600 °C and above brought the decoration/encapsulation of Ni nanoparticles by a thin layer of reduced CeO2 support with partial coverage of the Ni surface.
Kalai et al. investigated an Ni-based catalyst promoted with Ce in biogas reforming, and their results displayed that an active catalyst resistant to sintering and coke deposition was formed by a higher Ce/Ni ratio [29]. The promotional influence of various kinds of metal supported over Ni/SBA-15 was investigated via the dry reforming of CH4 [30]. It was found that the metal promoters could successfully improve Ni dispersion over the SBA-15’s siliceous framework and control the NiO crystallize sizes. Generally, the performance of a catalyst is determined not only by the active species but also by the type of support. Therefore, another method of reducing the coke generation is the employment of proper supports, which decreases sintering and carbon formation [31]. The supports are vital in improving activity and decreasing carbon production during the dry methane-reforming process. Basic supports help in the gasification of the carbon species and consequently control the carbon deposition. Moreover, their existence circumvents the destruction of the active phases by sintering [32].
Lino et al. investigated methane’s tri-reforming reaction performance using nickel catalysts supported on magnesium aluminate promoted with ZrO2, CeZrO2, and CeO2 [33]. Their results exhibited less coke formation and growth conversions when Zr and Ce-Zr promoted catalysts were used. Zirconia is used in many chemical reactions as a promoter and support because of its great thermal stability, medium basicity and acidity, and oxygen mobility [34]. The investigation of Li et al. found that adding ZrO2 in Ni/Al2O3 had a superb catalytic output owing to the evenly spread minor metallic nickel particles, which helped the reducibility of the nickel oxides and the metal–support interaction [35]. On the other hand, Li and co-authors reported the influence of PO4 upon the activity in perchlorate and stated enough endurance of Fe catalyst having a composition up to 5.0 M H3PO4 [36]. The anions of phosphate displayed an insignificant result in reducing the activity.
The aim of this work is to analyze the activity, stability, and efficiency of 10% Ni-containing catalyst prepared via the impregnation technique and operated at 1 bar and at 800 °C in DRM. An Ni catalyst is promoted with different loading of Ce (1%–5%) and supported upon 8% PO4-ZrO2.

2. Results and Discussion

2.1. Catalyst Characterization

Figure 1A depicts the N2 adsorption/desorption isotherms for the non-promoted and promoted catalysts, calcined at 600 °C. The N2 adsorption/desorption isotherms fall under the type-IV classification. In accordance with the International Union of Pure and Applied Chemistry (IUPAC) classification, mesoporous materials display the observed isotherm profiles. The first region of the isotherm can be ascribed to the adsorption of the monolayer–multilayer. The existence of mesoporous forms the hysteresis after an approximate P/Po = 0.4. Additionally, at the high P/Po, from 0.4 to 1.0, the H3 type of hysteresis loop is detected, which verifies the mesoporosity of the catalysts. Meanwhile, Figure 1B compares the distributions of pore size obtained by the Barrett, Joyner, and Halenda (BJH) method. It is evident that the distributions of pore size are narrow, showing that there existed only one type of pore, confirming the mesoporosity, as the pore diameter lies between 2 and 50 nm. Table 1 depicts the textural properties, which include the surface area, pore size, and pore diameters of 10% Ni catalysts supported by 8%PO4 and ZrO2 and promoted by different amounts of Ce.
XRD patterns of 10Ni+X%Ce/8%PO4+ZrO2 catalysts (X = 0, 1, 1.5, 2, 3, 5) are shown in Figure 2. The result indicated that the quantity of the promoter did not significantly influence the interaction between NiO and the support. The cubic phase of NiO is present in XRD patterns. The diffraction peaks at 2θ = 38.2, 43.8, and 61.5 were ascribed to NiO (JCPDS #47-1049), while the peaks at 2θ = 30.2 matched the monoclinic phase of ZrO2, and 38.2, 51.8, and 61.5 matched the tetragonal phase of ZrO2 (JCPDS #50-1089).
The H2−TPR profiles of 10Ni/8%PO4 + ZrO2, 10Ni + 1%Ce/8%PO4 + ZrO2, 10Ni +2%Ce/8%PO4 + ZrO2, 10Ni + 3%Ce/8%PO4 + ZrO2, and 10Ni + 5%Ce/8%PO4 + ZrO2 are presented in Figure 3. In all catalysts, wide reduction peaks related to the reduction of NiO were detected, but the mode of reduction of the catalysts was somewhat different. Peak maxima at 300–334, 564–593, and 732–780 °C were observed. The peak in the low temperature was ascribed to the non-associated NiO species, while the other peaks that appeared at the greater temperature were attributed to the NiOx complex species, where the support–metal interactions were high [37]. According to earlier literature, the reduction of catalysts at 300–334 °C was attributed to the free NiO species that had insufficient interaction with the support [38]. Alternatively, the peaks of catalyst at 588–593 and 732–780 °C were credited to the moderately strong and non-weakly interacting NiO species [39]. In general, there is no significant difference in the reduction temperature of the promoted and unpromoted catalysts. Table 2 illustrated the TPR results studied for the catalyst samples, in respect of amount of H2 uptake, temperature of peak maxima (TM), and percentage reduction degree. The 10%Ni + 3%Ce/8%PO4 + ZrO2 catalyst gave the highest value.
The CO2-TPD profile of 10Ni + x%Ce/8% PO4-ZrO2 (x = 0, 1, 3, and 5) catalyst samples are exhibited in Figure 4. The CO2-TPD profile represents the level of interaction of CO2 with the catalyst surface as well as the basic profile of the catalyst surface. In the low-temperature range (50–200 °C), CO2 reacts with the basic hydroxyl species on the surface, while at the medium temperature range (200–450 °C), it with reacts basic surface oxygen anion; lastly, it interacts with the lattice oxygen anions at the high-temperature range (450–800 °C) [40,41]. Commonly, CO2 adsorption capacity and dissociation are improved, raising the basicity of the catalyst. In turn, this reduces carbon deposition over the surface and hence lowers the catalyst deactivation [42,43]. Four principal adsorption peaks could be recognized with maxima at around 90 °C, 170 °C, 250 °C, and 640 °C, respectively for Ce-promoted catalysts. Meanwhile, the unpromoted catalyst gives three adsorption peaks and does not generate peaks beyond 250 °C. The peaks could be allocated to non-strong, Brønsted basic sites, adequate strength basic ‘Lewis acid-base sites’, and strong basic sites ‘Lewis basic sites’. Thus, the addition of the cerium enhanced the basicity. Table 3 displays the basicity strength of various catalysts. It is obvious that 3% of Ce gave the highest amount of total basicity.

2.2. Catalyst Activity

The experimental results of the reactor operated at the desired experimental conditions at 800 °C with support and in the absence of the active metals displayed the lack of activity. Figure 5A showed the CH4 conversion and stability for a reaction period of 7.5 h for the catalysts calcined at 600 °C. The activity of the non-promoted catalysts started at 77% and reduced and stabilized to about 72%. Virtually all the promoted catalysts exhibited similar patterns. The addition of the promoter enhanced the conversion from 15% to 24%. The catalyst promoted with 3% Ce generated the highest conversion and hence the best improvement. When the promoter loading was increased to 5%, the conversion was reduced.
Figure 5B illustrated the results of CO2 conversion versus time on stream, using the catalysts calcined at 600 °C and activated at 700 °C. The CO2 conversions were greater than the corresponding CH4 conversion. This was because the H2 was produced from the reforming participates in a side reaction by initiating a parallel reverse water gas shift reaction (RWGS) (CO2 + H2 → CO + H2O) [44]. The Figure 5B profile was similar to that of methane in Figure 5A. Here, the conversion of the non-promoted catalyst started at 81.5% and maintained a relatively stable trend with less reduction in time on stream. After 7.5 h, the conversion was about 80%. When the catalyst was promoted with Ce, the CO2 conversion was enhanced from 9% to 16%. The highest conversion was recorded using the 3% Ce loading catalyst. Loading higher than 3% such as 5% affected negatively affected the conversion due to similar reasons as those shown in Figure 5A. Table 4 summarized the performance comparison between the current work and the work of other investigators. The present catalysts offered superb methane activity.
Figure 6 exhibits the results obtained from the thermogravimetric (TGA) analysis of spent catalysts. The effect of the different loadings of the cerium promoter was seen in the weight loss. In the early period of the TGA, some catalysts showed a weight increase due to the formation of oxides in the ambient air. A sharp weight loss was observed for all promoted catalysts at 600 °C, ending at about 800 °C. It was observed that the weight loss reduced with the increase in the percentage loading of ceria: from 3% weight loss for the unpromoted catalyst down to about 1% weight loss for the 3%Ce promoted catalyst. There was an increase in the carbon deposits for the catalyst with ceria loading above 3%. It was observed that under this condition, the Ce promoter enhanced the gasification of the carbon deposits, and the percentage loading of 3% appeared to be the optimum.
Raman spectroscopy was done to study the degree of graphitization characteristics of the formed carbon on the used catalysts. Figure 7 displays the Raman spectra of used catalysts, and it could be noted that all catalysts exhibited two peaks with high intensities in the range between 1400 and 1600 cm−1. These peaks can be ascribed to the characteristic D and G bands, respectively [51]. The D band is attributed to the imperfections and disorders of the faulty structural carbons, and the G band is related to the vibrations of perfect graphite carbon. The peak areas ratio of the D and G (AD/AG) bands is used to assess the graphitic degree and the extent of defects in the carbon deposits over the used catalysts; a lower AD/AG ratio designates a well-ordered structure of the carbon deposits found on the used catalysts. There was an increase in the intensity of the peaks for 10 Ni + 5% Ce/8% PO4 + ZO2 relative to that of the 10 Ni + 3% Ce/8% PO4 + ZO2, indicating more carbon deposits as revealed by the TGA results. The ratio of AD/AG was computed, and this showed that the 10 Ni/8% PO4 + ZO2 (1.07) had the highest value relative to the ceria-promoted samples. This implies that the addition of cerium oxide slightly raised the extent of graphitization of the carbon deposits.

3. Materials and Methods

3.1. Preparation of Catalyst

The method of wet impregnation was employed for the preparation of catalysts used for the dry reforming of methane. The used procedure was described in our previous paper [4]. The phosphate zirconium was gifted by Daiichi Kigenso Kagaku Kogyo Co., Ltd., Osaka, Japan. The surface area of the support is 246 m2/g, while the particle size is 4.73 µm. The compositions of the catalysts and the textural properties such as pore volume (PV), surface area (BET), and pore diameter (PD) are shown in Table 1. Energy dispersive x-ray (EDX) analysis of two catalysts, 10%Ni + 1.5%Ce/8%PO4 + ZrO2 and 10%Ni + 1.5%Ce/8%PO4 + ZrO2, was performed. The composition of the total active metals prepared were 11.50% and 13.00%, respectively, while the EDX analysis gave compositions of the total active metals as 11.17 %and 12.96%, indicating close values.

3.2. Catalyst Performance Evaluation

The procedures for CO2 reforming of methane experiments were performed in a stainless steel tubular reactor (0.91 cm diameter and 0.30 m long). The reactor was obtained from PID Eng. & Tech Microactivity Reference Company. The particle size of the 10 Ni + x% Ce/8% PO4 + ZO2 ((x = 0–5) catalysts ranged between 19.0 and 19.6 nm. The operating pressure of 1 atm was applied during the reforming reactions. An amount of 150 mg of the catalyst was reduced with H2 flows of 20 mL/min for 30 min at 700 °C. Then, the system was treated with N2 for 15 min to desorb the physisorbed H2 from the bed. After that, the reactor temperature was adjusted to 800 °C by the flow of N2. In the usual test, the ratio of N2//CO2/CH4 was fixed to 2/6/6 at a total flow rate of 70 mL/min, giving a gas hourly space velocity of 28,000 mL (h·gcat)−1. The analysis of the catalyst compositions was carried out using a gas chromatograph, which was fitted with a conductivity detector (GC-2014 SHIMADZU, Kyoto, Japan). After the reaction, the reactor system was flushed with N2 gas to room temperature. Then, the catalysts were taken for characterization. Afterward, experimental reproducibility was performed by considering an average of triplicate experimental runs. The conversion formula is illustrated below:
% CH 4   conversion =   CH 4   in CH 4   out CH 4   in   ×   100
% CO 2   conversion =   CO 2   in CO 2   out CO 2   in   ×   100 .

3.3. Catalyst Characterization

3.3.1. N2 Physisorption

In this work, the textural characteristics of the catalysts were studied by N2 adsorption–desorption isotherms, which were computed at −197 °C with a Micromeritics Tristar II 3020 porosity and surface area analyzer. In each test, 0.2–0.3 g of catalyst was taken. The samples were previously degassed at 300 °C for 3 h to expel from the catalyst surface undesired adsorbed gases, organics, and water vapor. The BET technique was applied to obtain the surface areas by means of the N2 adsorption scale in the interval of 0.06–0.35 times the equilibrium pressure.

3.3.2. TPR

The reduction behavior of the catalysts was examined via an AutoChem-II Micromeritics device. The method used to analyze was described in our paper [5].

3.3.3. XRD

To investigate the diffraction profiles of the samples, X-ray diffraction (XRD) measurements of fresh catalysts were used. The XRD was accomplished using Rigaku (Miniflex), with Kα-Cu radiation at 40 kV and 40 mA. A 2θ range of 10–85° and a scanning step of 0.02° were used.

3.3.4. TGA

The quantification of the total deposited C on the spent catalysts was made employing thermogravimetric analysis (TGA) in air. The unit is a Shimadzu TGA analyzer. From the used catalysts, 10–15 mg was heated at the rate of 20 °C/min from 25 °C to 1000 °C, and the mass loss was registered.

3.3.5. CO2-TPD

The Micromeritics Autochem II apparatus, Micromeritics analyzer was used to perform the CO2-TPD. A helium gas stream of 30 mL/min was used to outgas 0.05 g of catalyst at 600 °C for 60 min. Then, the catalyst was brought to 50 °C. After that, a CO2 stream was admitted for 1 h, and the catalyst was then flushed with helium to remove the physically attached CO2. The desorption peak profile was registered, while the temperature was varied at 10 °C/min. The CO2 composition in the output stream was determined using a thermal conductivity detector. The peak areas gave the quantity of CO2 desorbed.

3.3.6. Raman Spectroscopy

Raman spectra were performed using an NMR-4500 Laser Raman Spectrometer. The excitation beam was adjusted to a wavelength of 532 nm. The measurement was carried out by the objective lens with 20× magnification. A 6 mW beam power and an exposure time of 3 min were used. The Raman shift of the spectra was measured in the range of 1000–3000 cm−1 while Spectra Manager Ver.2 software processed the profiles.

4. Conclusions

Various cerium loadings were used to promote 10% Ni/8% PO4+ZO2 catalysts. The cerium loadings constituted the values (1%, 1.5%, 2%, 2.5%, 3%, and 5%). These catalysts were used for CO2 reforming of methane in a fixed bed reactor. The Ce promoter improved the catalytic performance. The highest conversions of CH4 and CO2 at 800 °C reaction were obtained using 3% Ce loading. The study of the Raman spectroscopy denoted that the 3% Ce-promoted catalyst possessed the least values of D and G bands, indicating the highest structural perfection. Table 1 also displayed that the 3% Ce loading possesses the highest surface area, the largest pore volume, and the greatest pore diameter. The study of the CO2-TPD revealed the enhancement of basicity by the addition of the Ce promoter. Table 2 and Table 3 exhibited the highest percentage reduction degree and the highest total basicity were given by the 3% Ce loading catalyst, respectively.
The investigation of the TGA indicated insignificant amounts of carbon deposition and therefore did not pose stimulating effects of carbon deposition onto the surface of the catalyst. The promoted catalysts without exception enhanced the conversions of CH4 and CO2. Among the promoted catalysts tested, The 10Ni + 3%Ce/8%PO4 + ZrO2 catalyst system operated at 1 bar and at 800 °C gave the highest conversions of CH4 (95%) and CO2 (96%). Thus, steady stability profiles of CH4 and CO2 conversions during time on stream for all promoted catalysts were observed.

Author Contributions

Experiment and Writing—original draft preparation A.A.I., A.S.A.-F., and S.O.K.; XRD analysis N.S.K., Writing—review and editing, A.E.A. and A.H.F. All authors have read and agreed to the published version of the manuscript.

Funding

The work is supported by the Deanship of Scientific Research programs of King Saud University via (project (No. RG-1435-078).

Acknowledgments

The KSU authors would like to extend their sincere appreciation to the Deanship of Scientific Research at the King Saud University for its funding for this research group project # (RG-1435-078).

Conflicts of Interest

The authors affirm no conflict of interest.

References

  1. 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]
  2. Zoundi, Z. CO2 emissions, renewable energy and the Environmental Kuznets Curve, apanel cointegration approach. Renew. Sustain. Energy Rev. 2017, 72, 1067–1075. [Google Scholar] [CrossRef]
  3. Hoehne, C.G.; Chester, M.V. Greenhouse gas and air quality effects of auto first-last mile use with transit. Transp. Res. Part D Transp. Environ. 2017, 53, 306–320. [Google Scholar]
  4. Al-Fatesh, A.S.; Arafat, Y.; Ibrahim, A.A.; Kasim, S.O.; Alharthi, A.; Fakeeha, A.H.; Abasaeed, A.E.; Giuseppe Bonura, G.; Francesco 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]
  5. Al-Fatesh, A.S.; Kasim, S.O.; Ibrahim, A.A.; Fakeeha, A.H.; Abasaeed, A.E.; Alrasheed, R.; Ashamari, R.; Bagabas, A. Combined Magnesia, Ceria and Nickel catalyst supported over γ-Alumina Doped with Titania for Dry Reforming of Methane. Catalysts 2019, 9, 188. [Google Scholar] [CrossRef] [Green Version]
  6. Lau, C.; Tsolakis, A.; Wyszynski, M. Biogas upgrade to syn-gas (H2–CO) via dry and oxidative reforming. Int. J. Hydrogen Energy 2011, 36, 397–404. [Google Scholar] [CrossRef]
  7. Kwon, B.W.; Oh, J.H.; Kim, G.S.; Yoon, S.P.; Han, J.; Nam, S.W.; Ham, H.C. The novel perovskite-type Ni-doped Sr0.92Y0.08TiO3 as a reforming biogas (CH4 + CO2) for H2 production. Appl. Energy 2018, 227, 213–219. [Google Scholar]
  8. IEA. World Energy Outlook 2015; OECD Publishing: Paris, France, 2015; Available online: https://doi.org/10.1787/weo-2015-en (accessed on 10 November 2015).
  9. Jiao, F.; Pan, X.; Xiao, J.; Li, H.; Ma, H.; Wei, M.; Pan, Y.; Zhou, Z.; Li, M.; Miao, S.; et al. Selective conversion of syngas to light olefins. Science 2016, 351, 1065–1068. [Google Scholar] [CrossRef]
  10. Cui, G.; Liu, J.; Wei, M.; Feng, X.; Elsworth, D. Evolution of permeability during the process of shale gas extraction. J. Nat. Gas Sci. Eng. 2018, 49, 94–109. [Google Scholar] [CrossRef]
  11. Quéllar-Franca, R.M.; Azapagic, A. Carbon capture, storage and utilization technologies: A critical analysis and comparison of their life cycle environmental impacts. J. CO2 Util. 2015, 9, 82–102. [Google Scholar] [CrossRef]
  12. Vieira, M.O.; Aquino, A.S.; Schütz, M.K.; Vecchia, F.D.; Ligabue, R.; Seferin, M.; Einloft, S. Chemical Conversion of CO2: Evaluation of Different Ionic Liquids as Catalysts in Dimethyl Carbonate Synthesis. Energy Procedia 2017, 114, 7141–7149. [Google Scholar] [CrossRef]
  13. Sheu, E.J.; Mokheimer, E.M.; Ghoniem, A.F. A review of solar methane reforming systems. Int. J. Hydrogen Energy 2015, 40, 12929–12955. [Google Scholar] [CrossRef] [Green Version]
  14. García-Diéguez, M.; Finocchio, E.; Larrubia, M. Ángeles; Alemany, L.J.; Busca, G. Characterization of alumina-supported Pt, Ni and PtNi alloy catalysts for the dry reforming of methane. J. Catal. 2010, 274, 11–20. [Google Scholar] [CrossRef]
  15. Nawfal, M.; Gennequin, C.; Labaki, M.; Nsouli, B.; Aboukaıs, A.; AbiAad, E. Hydrogen production by methane steam reforming over Ru supported on NieMgeAl mixed oxides prepared via hydrotalcite route. Int. J. Hydrogen Energy 2015, 40, 1269–1277. [Google Scholar] [CrossRef]
  16. Al-Fatesh, A.; Singh, S.K.; Kanade, G.; Atia, H.; Fakeeha, A.H.; Ibrahim, A.A.; El-Toni, A.M.; Labhasetwar, N.K. Rh promoted and ZrO2/Al2O3 supported Ni/Co based catalysts: High activity for CO2 reforming, steam–CO2 reforming and oxy–CO2 reforming of CH4. Int. J. Hydrogen Energy 2018, 43, 12069–12080. [Google Scholar] [CrossRef]
  17. Arandiyan, H.; Peng, Y.; Liu, C.; Chang, H.; Li, J. Effects of noble metals doped on mesoporous LaAlNi mixed oxide catalyst and identification of carbon deposit for reforming CH4 with CO2. J. Chem. Tech. Biotechnol. 2014, 89, 372–381. [Google Scholar] [CrossRef]
  18. 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]
  19. Hou, Z.; Chen, P.; Fang, H.; Zheng, X.; Yashima, T. Production of synthesis gas via methane reforming with CO2 on noble metals and small amount of noble-(Rh-)promoted Ni catalysts, Int. J. Hydrogen Energy 2006, 31, 555–561. [Google Scholar] [CrossRef]
  20. Usman, M.; Daud, W.M.A.W.; Abbas, H.F. Dry reforming of methane: Influence of process parameters—A review. Renew. Sust. Energ. Rev. 2015, 45, 710–744. [Google Scholar]
  21. Akri, M.; Shu Zhao, S.; Xiaoyu Li, X.; Zang, K.; Lee, A.F.; Isaacs, M.A.; Xi, W.; Gangarajula, Y.; Luo, J.; Ren, Y.; et al. Atomically dispersed nickel as coke-resistant active sites for methane dry reforming. Nat. Commun. 2019, 10, 5181. [Google Scholar] [CrossRef] [Green Version]
  22. Goula, M.; Charisiou, N.; 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 CeO 2 or La 2 O 3 catalysts. Int. J. Hydrogen Energy 2017, 42, 13724–13740. [Google Scholar] [CrossRef]
  23. Abba, M.O.; Gonzalez-Delacruz, V.M.; Colón, G.; Sebti, S.; Caballero, A. In situ XAS study of an improved natural phosphate catalyst for hydrogen production by reforming of methane. Appl. Catal. B Environ. 2014, 150, 459–465. [Google Scholar] [CrossRef]
  24. 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. Hydrogen Energy 2019, 44, 27784–27794. [Google Scholar] [CrossRef]
  25. Rostrupnielsen, J.; Hansen, J. CO2-Reforming of Methane over Transition Metals. J. Catal. 1993, 144, 38–49. [Google Scholar] [CrossRef]
  26. Abdullah, B.; Ghani, N.A.A.; Vo, D.-V.N. Recent advances in dry reforming of methane over Ni-based catalysts. J. Clean. Prod. 2017, 162, 170–185. [Google Scholar] [CrossRef] [Green Version]
  27. Luisetto, I.; Tuti, S.; Romano, C.; Boaro, M.; Di Bartolomeo, E.; Kesavan, J.K.; Kumar, S.S.; Selvakumar, K. Dry reforming of methane over Ni supported on doped CeO2: New insight on the role of dopants for CO2 activation. J. CO2 Util. 2019, 30, 63–78. [Google Scholar] [CrossRef]
  28. Li, M.; Van Veen, A.C. Tuning the catalytic performance of Ni-catalysed dry reforming of methane and carbon deposition via Ni-CeO 2- x interaction. Appl. Catal. B Environ. 2018, 237, 641–648. [Google Scholar] [CrossRef] [Green Version]
  29. Kalai, D.Y.; Stangeland, K.; Tucho, W.M.; Jin, Y.; Yu, Z. Biogas reforming on hydrotalcite-derived Ni-Mg-Al catalysts: The effect of Ni loading and Ce promotion. J. CO2 Util. 2019, 33, 189–200. [Google Scholar] [CrossRef]
  30. Chong, C.C.; Teh, L.P.; Setiabudi, H.D. Syngas production via CO2 reforming of CH4 over Ni-based SBA-15: Promotional effect of promoters (Ce, Mg, and Zr). Mater. Today Energy 2019, 12, 408–417. [Google Scholar] [CrossRef]
  31. Pietraszek, A.; Koubaissy, B.; Roger, A.-C.; Kiennemann, A. The influence of the support modification over Ni-based catalysts for dry reforming of methane reaction. Catal. Today 2011, 176, 267–271. [Google Scholar] [CrossRef]
  32. Chein, R.-Y.; Fung, W.-Y. Syngas production via dry reforming of methane over CeO2 modified Ni/Al2O3 catalysts. Int. J. Hydrogen Energy 2019, 44, 14303–14315. [Google Scholar] [CrossRef]
  33. Lino, A.V.P.; Calderon, Y.N.C.; Mastelaro, V.R.; Assaf, E.M.; Assaf, J.M.; Colmenares, Y.N. Syngas for Fischer-Tropsch synthesis by methane tri-reforming using nickel supported on MgAl2O4 promoted with Zr, Ce and Ce-Zr. Appl. Surf. Sci. 2019, 481, 747–760. [Google Scholar] [CrossRef]
  34. Li, J.; Chen, J.; Song, W.; Liu, J.; Shen, W. Influence of zirconia crystal phase on the catalytic performance of Au/ZrO2 catalysts for low-temperature water gas shift reaction. Appl. Catal. A Gen. 2008, 334, 321–329. [Google Scholar] [CrossRef]
  35. Li, H.; Wang, J. Study on CO2 reforming of methane to syngas over Al2O3–ZrO2 supported Ni catalysts prepared via a direct sol–gel process. Chem. Eng. Sci. 2004, 59, 4861–4867. [Google Scholar] [CrossRef]
  36. Li, Q.; Wu, G.; Cullen, D.A.; More, K.L.; Mack, N.H.; Chung, H.T.; Zelenay, P. Phosphate-Tolerant Oxygen Reduction Catalysts. ACS Catal. 2014, 4, 3193–3200. [Google Scholar] [CrossRef] [Green Version]
  37. Hu, L.; Urakawa, A. Continuous CO2 capture and reduction in one process: CO2 methanation over unpromoted and promoted Ni/ZrO2. J. CO2 Util. 2018, 25, 323–329. [Google Scholar] [CrossRef] [Green Version]
  38. Pan, Q.; Peng, J.; Sun, T.; Wang, S.; Wang, S. Insight into the reaction route of CO2 methanation: Promotion effect of medium basic sites. Catal. Commun. 2014, 45, 74–78. [Google Scholar] [CrossRef]
  39. Salker, A.V.; Naik, S.J. Mechanistic study of acidic and basic sites for CO oxidation over nano based Co2xFexWO6 catalysts. Appl. Catal. B Environ. 2009, 89, 246–254. [Google Scholar] [CrossRef]
  40. Singh, H.; Rai, A.; Yadav, R.; Sinha, A.K. Glucose hydrogenation to sorbitol over unsupported mesoporous Ni/NiO catalyst. Mol. Catal. 2018, 451, 186–191. [Google Scholar]
  41. Wang, C.; Sun, N.; Zhao, N.; Wei, W.; Zhao, Y. Template-free preparation of bimetallic mesoporous Ni-Co-CaO-ZrO2 catalysts and their synergetic effect in dry reforming of methane. Catal. Today 2017, 281, 268–275. [Google Scholar] [CrossRef]
  42. Alipour, Z.; Rezaei, M.; Meshkani, F. Effect of Ni loadings on the activity and coke formation of MgO-modified Ni/Al2O3 nanocatalyst in dry reforming of methane. J. Energy Chem. 2014, 23, 633–638. [Google Scholar] [CrossRef]
  43. Ding, M.; Tu, J.; Zhang, Q.; Wang, M.; Tsubaki, N.; Wang, T.; Ma, L. Enhancement of methanation of bio-syngas over CeO 2 -modified Ni/Al 2 O 3 catalysts. Biomass-Bioenergy 2016, 85, 12–17. [Google Scholar] [CrossRef]
  44. Conner, W.C.; Falconer, J.L. Spillover in Heterogeneous Catalysis. Chem. Rev. 1995, 95, 759–788. [Google Scholar] [CrossRef]
  45. Monteiro, W.F.; Vieira, M.O.; Calgaro, C.O.; Perez-Lopez, O.W.; Ligabue, R.A. Dry reforming of methane using modified sodium and protonated titanate nanotube catalysts. Fuel 2019, 253, 713–721. [Google Scholar] [CrossRef]
  46. Cao, K.; Gong, M.; Yang, J.; Cai, J.; Chu, S.; Chen, Z.; Shan, B.; Chen, R. Nickel catalyst with atomically-thin meshed cobalt coating for improved durability in dry reforming of methane. J. Catal. 2019, 373, 351–360. [Google Scholar] [CrossRef]
  47. Sun, Y.; Zhang, G.; Xu, Y.; Zhang, R. Dry reforming of methane over Co-Ce-M/AC-N catalyst: Effect of promoters (Ca and Mg) and preparation method on catalytic activity and stability. Int. J. Hydrogen Energy 2019, 44, 22972–22982. [Google Scholar] [CrossRef]
  48. Fang, X.; Zhang, J.; Liu, J.; Wang, C.; Huang, Q.; Xu, X.; Peng, H.; Liu, W.; Wang, X.; Zhou, W. Methane dry reforming over Ni/Mg-Al-O: On the significant promotional effects of rare earth Ce and Nd metal oxides. J. CO2 Util. 2018, 25, 242–253. [Google Scholar] [CrossRef]
  49. Akri, M.; Pronier, S.; Chafik, T.; Achak, O.; Granger, P.; Simon, P.; Trentesaux, M.; Batiot-Dupeyrat, C. Development of nickel supported La and Ce-natural illite clay for autothermal dry reforming of methane: Toward a better resistance to deactivation. Appl. Catal. B Environ. 2017, 205, 519–531. [Google Scholar] [CrossRef]
  50. Kim, S.S.; Lee, S.M.; Won, J.M.; Yang, H.J.; Hong, S.C. Effect of Ce/Ti ratio on the catalytic activity and stability of Ni/CeO2–TiO2 catalyst for dry reforming of methane. Chem. Eng. J. 2015, 280, 433–440. [Google Scholar] [CrossRef]
  51. Chen, X.; Oh, W.-D.; Hu, Z.-T.; Sun, Y.-M.; Webster, R.D.; Li, S.-Z.; Lim, T.-T. Enhancing sulfacetamide degradation by peroxymonosulfate activation with N-doped graphene produced through delicately-controlled nitrogen functionalization via tweaking thermal annealing processes. Appl. Catal. B Environ. 2018, 225, 243–257. [Google Scholar] [CrossRef]
Figure 1. Nitrogen adsorption–desorption isotherms (A) and pore size distribution (B) diameter of fresh catalysts calcined at 600 °C.
Figure 1. Nitrogen adsorption–desorption isotherms (A) and pore size distribution (B) diameter of fresh catalysts calcined at 600 °C.
Catalysts 10 00242 g001
Figure 2. X-ray diffraction patterns of 10Ni + x%Ce/8%PO4 + ZrO2 catalysts (x = 0, 1, 1.5, 2, 3, 5) calcined at 600 °C.
Figure 2. X-ray diffraction patterns of 10Ni + x%Ce/8%PO4 + ZrO2 catalysts (x = 0, 1, 1.5, 2, 3, 5) calcined at 600 °C.
Catalysts 10 00242 g002
Figure 3. TPR-H2 profiles of the reduced catalysts calcined at 600 °C.
Figure 3. TPR-H2 profiles of the reduced catalysts calcined at 600 °C.
Catalysts 10 00242 g003
Figure 4. CO2-TPD profiles of 10Ni+z%Ce/8% PO4-ZrO2 (z = 0, 1, 3, and 5) catalysts.
Figure 4. CO2-TPD profiles of 10Ni+z%Ce/8% PO4-ZrO2 (z = 0, 1, 3, and 5) catalysts.
Catalysts 10 00242 g004
Figure 5. Catalytic activity of 10 Ni + x% Ce/8% PO4 + ZO2 (x = 0, 1, 1.5, 2, 2.5, 3.5, and 5) catalyst calcined at 600 °C, operated at 800 °C, GHSV = 28000 (mL/g/h) (A) CH4 conversion against time-on-stream (TOS) (B) CO2 conversion against TOS.
Figure 5. Catalytic activity of 10 Ni + x% Ce/8% PO4 + ZO2 (x = 0, 1, 1.5, 2, 2.5, 3.5, and 5) catalyst calcined at 600 °C, operated at 800 °C, GHSV = 28000 (mL/g/h) (A) CH4 conversion against time-on-stream (TOS) (B) CO2 conversion against TOS.
Catalysts 10 00242 g005
Figure 6. Thermogravimetric (TGA) results of the spent catalysts with 700, 600, and 800 °C activation, calcination, and reaction temperatures, respectively.
Figure 6. Thermogravimetric (TGA) results of the spent catalysts with 700, 600, and 800 °C activation, calcination, and reaction temperatures, respectively.
Catalysts 10 00242 g006
Figure 7. Raman spectra of used catalysts of 10 Ni + x% Ce/8% PO4 + ZO2 (x = 0, 3, and 5) catalyst calcined at 600 °C, operated at 800 °C, GHSV = 28,000 (mL/g/h).
Figure 7. Raman spectra of used catalysts of 10 Ni + x% Ce/8% PO4 + ZO2 (x = 0, 3, and 5) catalyst calcined at 600 °C, operated at 800 °C, GHSV = 28,000 (mL/g/h).
Catalysts 10 00242 g007
Table 1. Textural properties of 10% Ni catalysts supported by 8%PO4 and ZrO2 and promoted by different amounts of cerium.
Table 1. Textural properties of 10% Ni catalysts supported by 8%PO4 and ZrO2 and promoted by different amounts of cerium.
CatalystBET1 Surface Area (m2/g)Av. Pore Diameter (nm)Pore Volume (cm3/g)
10%NiO/8%PO4 + ZrO2157.50.24.45
10%NiO + 1%Ce/8%PO4 + ZrO2157.50.24.45
10%NiO + 1.5%Ce/8%PO4 + ZrO2145.20.184.46
10%NiO + 2%Ce/8%PO4 + ZrO2166.70.224.67
10%NiO + 2.5%Ce/8%PO4 + ZrO2155.00.204.60
10%NiO + 3%Ce/8%PO4 + ZrO2168.20.224.67
10%NiO + 5%Ce/8%PO4 + ZrO2127.20.164.55
1 BET- Brunauer-Emmett-Teller.
Table 2. TPR data of the studied catalyst samples in terms of amount of H2 consumption and temperature of peak maxima (TM).
Table 2. TPR data of the studied catalyst samples in terms of amount of H2 consumption and temperature of peak maxima (TM).
CatalystsH2 Uptake
(mmole/g)
TM1 (°C)TM2 (°C)TM3 (°C)Reduction Degree (%)
10%Ni/8%PO4 + ZrO23.176-5887501.86
10%Ni + 1%Ce/8%PO4 + ZrO23.023-5937801.77
10%Ni + 2%Ce/8%PO4 + ZrO22.9563005787461.73
10%Ni + 3%Ce/8%PO4 + ZrO22.5173295647322.44
10%Ni + 5%Ce/8%PO4 + ZrO22.7963345777551.64
Table 3. Distribution of basic sites for Ni-supported catalysts.
Table 3. Distribution of basic sites for Ni-supported catalysts.
SampleWeak Basic Sites (µmol/g)Medium Basic Sites (µmol/g)Strong Basic Sites (µmol/g)Total Basicity (µmol/g)
10Ni/8%PO4 + ZrO218.02719.957-37.984
10Ni + 2%Ce/8%PO4+ZrO218.01940.213-58.232
10Ni + 3%Ce/8%PO4 + ZrO230.88147.051-77.932
10Ni + 5%Ce/8%PO4 + ZrO224.45624.353-48.809
Table 4. Comparison between the CH4 conversion acquired and those in the literature using different catalysts for the dry reforming of CH4.
Table 4. Comparison between the CH4 conversion acquired and those in the literature using different catalysts for the dry reforming of CH4.
CatalystCH4:CO2GHSV1
(mL/g/h)
T
(°C)
%
CH4
References
10%Ni/8%PO4+ZrO21:11300080080[4]
Ni-Zr/SBA-151:11500080088[30]
Ni-HTNT1:11200070074[45]
0.5% Co/ Ni/γ-Al2O31:11200065070[46]
1Co-1Ce-1Ca/AC-N1:172080068[47]
Ni-Ce/Mg-Al-O1:17200080086[48]
10Ni15La/illite-clay5:46000080080[49]
Ni/50% CeO2–50% TiO21:11440075090[50]
10%Ni+3%Ce/8%PO4+ZrO21:12800080095Present work
1 GHSV- Gas Hourly Space Velocity.

Share and Cite

MDPI and ACS Style

Ibrahim, A.A.; Al-Fatesh, A.S.; Kumar, N.S.; Abasaeed, A.E.; Kasim, S.O.; Fakeeha, A.H. Dry Reforming of Methane Using Ce-modified Ni Supported on 8%PO4 + ZrO2 Catalysts. Catalysts 2020, 10, 242. https://doi.org/10.3390/catal10020242

AMA Style

Ibrahim AA, Al-Fatesh AS, Kumar NS, Abasaeed AE, Kasim SO, Fakeeha AH. Dry Reforming of Methane Using Ce-modified Ni Supported on 8%PO4 + ZrO2 Catalysts. Catalysts. 2020; 10(2):242. https://doi.org/10.3390/catal10020242

Chicago/Turabian Style

Ibrahim, Ahmed A., Ahmed S. Al-Fatesh, Nadavala Siva Kumar, Ahmed E. Abasaeed, Samsudeen O. Kasim, and Anis H. Fakeeha. 2020. "Dry Reforming of Methane Using Ce-modified Ni Supported on 8%PO4 + ZrO2 Catalysts" Catalysts 10, no. 2: 242. https://doi.org/10.3390/catal10020242

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