Next Article in Journal
Sustainable Catalytic Synthesis of 2,5-Diformylfuran from Various Carbohydrates
Previous Article in Journal
Controlled Synthesis and Photoelectrochemical Performance Enhancement of Cu2−xSe Decorated Porous Au/Bi2Se3 Z-Scheme Plasmonic Photoelectrocatalyst
Previous Article in Special Issue
Synthesis and Characterization of Silver Nanoparticles Prepared with Carrasquilla Fruit Extract (Berberis hallii) and Evaluation of Its Photocatalytic Activity
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

The Effect of Calcination Temperature on Various Sources of ZrO2 Supported Ni Catalyst for Dry Reforming of Methane

by
Ahmed Aidid Ibrahim
1,*,
Anis Hamza Fakeeha
1,*,
Mahmud Sofiu Lanre
1,
Abdulrhman S. Al-Awadi
1,
Salwa Bader Alreshaidan
2,
Yousef Abdulrahman Albaqmaa
1,
Syed Farooq Adil
2,
Ateyah A. Al-Zahrani
1,
Ahmed Elhag Abasaeed
1 and
Ahmed S. Al-Fatesh
1
1
Chemical Engineering Department, College of Engineering, King Saud University, P.O. Box 800, Riyadh 11421, Saudi Arabia
2
Department of Chemistry, College of Science, King Saud University, P.O. Box 800, Riyadh 11451, Saudi Arabia
*
Authors to whom correspondence should be addressed.
Catalysts 2022, 12(4), 361; https://doi.org/10.3390/catal12040361
Submission received: 16 January 2022 / Revised: 9 March 2022 / Accepted: 20 March 2022 / Published: 23 March 2022
(This article belongs to the Special Issue Advanced Nanomaterials - Synthesis and Applications in Catalysis)

Abstract

:
Dry reforming of methane (DRM) over an Ni-based catalyst is an innovative research area due to the growing environmental awareness about mitigating global warming gases (CH4 and CO2) and creating a greener route of synthesis. Herein, 5% Ni supported on ZrO2 obtained from various sources was prepared by the impregnation method. The catalysts were calcined at 600, 700, and 800 °C. Furthermore, Ni-RC stabilized with MgO, SiO2, TiO2, and Y2O3 were tested. Characterization techniques employed comprise the N2 physisorption, infrared spectroscopy, Raman, thermogravimetric analysis, XRD, and TEM. The results of the present study indicated that the ZrO2 support source had a profound effect on the overall performance of the process. The best catalyst Ni-RC gave an average conversion of CH4 and CO2 of 61.5% and 63.6% and the least deactivation of 10.3%. The calcination pretreatment differently influenced the catalyst performance. When the average methane conversion was higher than 40%, increasing the calcination temperature decreased the activity. While for the low activity catalysts with an average methane conversion of less than 40% the impact of the calcination temperature did not constantly decrease with the temperature rise. The stabilization of Ni-RC denoted the preference Y2O3 stabilized catalyst with average values of CH4 and CO2 conversion of about 67% and 72%, respectively. The thorough study and fine correlation will be advantageous for technologically suitable Ni-15Y-RC catalysts for DRM.

1. Introduction

Almost 85% of the total world’s energy comes from fossil fuels. Nevertheless, the combustion of various types of fossil fuels over a period of one and a half centuries, has caused numerous problems, denoted as global warming, caused by the release of greenhouse gases. In other words, the utilization of fossil fuels, covering the energy requirements of the world, is accompanied by intolerable air pollution due to the generation of greenhouse gases [1,2,3]. The world’s population is continuously increasing, and the burning of fossil fuels highly promotes the amount of atmospheric carbon dioxide and other heat-catching gases, resulting in climate change issues. CH4 gas is a greenhouse gas that contributes to global world warming [4,5]. Dry reforming of methane (DRM) is a catalytic conversion method that reduces the emission of greenhouse gases CO2 and CH4 and transforms them into useful products. The product of the DRM CO and H2 gases with an optimum H2/CO ratio near to one is therefore beneficial for liquid hydrocarbon production via the Fischer–Tropsch synthesis [6,7]. The main reaction for DRM reaction is displayed in Equation (1), whereas the accompanying side reactions are given in Equations (2)–(4) [8]. Equation (2) represents the disproportionation, while Equation (3) denotes the methane cracking. Equation (4), commonly termed as reverse water–gas shift reaction, is the key method for the H2O formation. Currently, DRM is the preferable research topic, as it displays outstanding performance in its capacity to reduce greenhouse gases [9]. DRM is an endothermic process that entails the use of high temperatures, which favors rapid catalyst deactivation and thus lowers its industrial applications.
This aspect renders the CH4 to decompose and the coke to form, as shown in Equations (2) and (3). Supported metallic catalysts are usually employed for DRM. Transitional metals such as Ni are practically used [10,11,12]. Though noble metals perform pronounced activity and stability, their utilization is avoided, owing to their scarcity and high prices [13]. The supporting interactions and dispersion of active species determine the efficiency of the supported catalysts. The adjustment of Ni-based catalysts by various supports is an effective technique to diminish carbon formation and promote the lifetime of the catalyst:
CH 4 + CO 2     2 H 2 + 2 CO Δ H 298   K 0 = 247.31   kJ mol ,
2 CO   C O 2 + C Δ   H 298   K 0 = 172.44   kJ mol ,
CH 4 2 H 2 + C Δ H 298   K 0 = 74.87   kJ mol ,
H 2 + CO 2 H 2 O + CO Δ H 298   K 0 = 41.17   kJ mol .
Many investigators synthesized appropriate catalysts by using numerous supports such as Al2O3, SiO2, ZrO2, etc. [14]. Basic supports like ZrO2 are found to accelerate the CO2 adsorption and decomposition, giving a strong metal-support interaction for reduced coke deposition [15]. Another property of catalyst support, which was found to be very useful for reforming methane, is oxygen storage capacity. The supporting merit in DRM is to give a stable, wide surface area to efficiently disperse the active phase and inhibit its sintering, and to contribute to the reaction mechanism through activation of CO2 [16,17]. ZrO2 is among the most universally used zirconium compounds in nature. Its phases of crystallization comprise cubic, monoclinic, and tetragonal [18,19,20] at ambient pressure. The major phase depends rigorously on parameters such as temperature, doping, sample preparation method, oxygen vacancy, and crystallite size [21,22].
The calcination pretreatments modify the structures of the supports [23,24]. Several investigations were performed to display the influences of this parameter during the catalyst formulation [25,26,27,28,29,30,31,32]. Low calcination temperature might cause incomplete decomposition of the metal salt precursor, thereby reducing the active component, whereas high-temperature sinters the active species, reduces the specific surface area, and deforms the carrier structure [29]. The Fischer–Tropsch activities of the catalysts were examined with and without previous calcination. It was found that the calcination in the air increased the liquid product selectivity of all the catalysts towards the diesel fraction. On the other hand, Sun et al. elaborated the catalytic performance of DRM over nitrogen doped activated carbon (Co/AC-N) catalysts, focusing on the calcination temperature effects [33]. Their result displayed that the catalytic activity was influenced by the calcination temperature as it altered the Co2+/Co3+ molar ratio and the content of nitrogen functional groups.
In this investigation, we elaborated on the impact of different sources of zirconia support with amphoteric nature and the influence of different calcination temperatures on the overall catalyst efficiency towards the dry reforming CH4 reaction. Thus, we synthesized 5% Ni supported on ZrO2 via the impregnation technique and applied it to DRM. The ZrO2 was obtained from diverse sources. We sorted the broad effects of calcination, on the performance of 5%Ni/ZrO2 catalyst. The substantial alterations in the catalyst structure and surface effects ascribed to the calcination were fully elaborated. The catalysts were calcined at 600 °C, 700 °C, and 800 °C.

2. Results and Discussion

Figure 1 presents the N2 physisorption profiles of the catalysts. The N2 physisorption isotherms of catalysts indicate the type IV isotherm with an H1 hysteresis loop at higher relative pressure, which is in line with the International Union of Pure and Applied Chemistry (IUPAC) classification, which designates that most of the pores are cylindrical in shape with an ordered mesoporous framework. The figure displays an intense rise at P/Po = 0.75 in the second half of the isotherm, owing to capillary condensation in inter particular mesopores/macropores. This matches the state where the adsorbate critical temperature is higher than the adsorption temperature. Mesoporous supports would also help the formation of active carbon, which has a habit of prompt gasification. The N2-physisorption, of Ni-RC and Ni-ZH catalysts showed similar trends and the extent of N2-physisorption decreased with the calcination temperature in accordance with the surface area measurement. The strong loss of surface area on the increasing calcination temperature is very common with the ZrO2 system due to the sintering of strong ZrO2 crystallite into bigger and denser aggregate [34,35]. Chan et al. carried out different calcination temperatures in preparation of ZrO2 and found very similar results to our ZrO2 supported Ni catalyst [34]. Chan et al. also found that phase transformation had an insignificant role in the decrease of surface area at high temperatures [34]. Figures S2–S4 display the pore size distribution of the catalysts at different calcination temperatures. The result indicates that the pore sizes are mesoporous, and that the higher calcination temperature increases the porosity. The catalysts displayed pore volume rise along with the pore size at lower pore size values.
The N2 physisorption results and crystalline size of 5%Ni/ZrO2 catalysts for different calcination temperatures are displayed in Table 1.
The 5%Ni/ZrO2 calcined catalysts and pristine supports were examined via FT-IR to pinpoint different kinds of bonds. Figure 2 exhibited bands positioned around 510 and 580 cm−1 characteristics for Ni-O and Zr-O widening vibrations [34], whereas ZrO2 was ascribed to the band around 730–750 cm−1.
The performance of the catalysts for various calcination temperatures was carried out in terms of methane and carbon dioxide conversions (Figure 3). The deactivation factor was also calculated. Table 2 provides the results of the performance. The result indicates that the catalysts possess different activities with respect to the source type. Conversions of methane and carbon dioxide declined with time during the reaction as a result of deactivation. The relatively higher conversion of CO2 is attributed to the reverse water gas shift reaction, which further consumed CO2 to generate CO, in accordance with Equation (4).
The calculated deactivation factors of CH4 conversion over Ni-MK, Ni-ZH, Ni-DK, Ni-EN, and Ni-RC are 22.2%, 27.2%, 31.1%, 10.3%, and 10.3%, respectively. The results of Table 2 show the TGA values of the catalysts. It is apparent that the less active catalysts Ni-Mk and Ni-EN provide an average low value of TGA. This could be related to the poor performance, which causes less carbon to accrue [23,35]. The calcination pretreatment affects the catalyst performance differently. For instance, for the relatively high activity catalysts with an average methane conversion higher than 40% such as Ni-RC, Ni-ZH, and Ni-DK, increasing the calcination temperature decreases the activity. While for the low activity catalysts with average temperatures lower than 40%, the calcination impact fluctuated. For example, the Ni-MK catalyst’s activity increases from 600 °C to 700 °C calcination and then drops from 700 °C to 800 °C, while for Ni-EN catalyst, the activity increases when increasing the calcination temperature from 600 °C to 700 °C and then decreases from 700 °C to 800 °C. The variation of the catalyst with calcination temperatures may be related to the changes in ZrO2 morphologies, which could show the various oxygen mobility and diverse capacity of metal-support synergies. The aspects of zirconia influenced coke removal and thermal sintering of Ni. The study of the effects of zirconia source and the calcination temperature resulted in the superiority of the Ni-RC catalyst calcined at 700 °C. The Raman instrument offers further evidence about the graphitization degree and the kind of carbon on the used catalysts. Figure 4 displays the Raman spectra of the used catalysts. The spectra are divisible into zones. Zone I lies between 1250 and 1650 cm−1 and zone II occurs between 2400 and 3000 cm−1. In the range between 1250 and 1650 cm−1, the dominant characteristic peaks, D-band and G-band, are situated at 1341 cm−1 and at 1568 cm−1, respectively. The D-band is attributed to carbon deposits with imperfect structures that are disordered (amorphous) [36] and the G-band represents graphite. Because of the unification and implications of the band of zone I, the 2D band appears at 2666 [36]. The extent of carbon crystallinity formed during the reaction is appropriately assessed from the relative intensity sizes of the G and D bands (ID/IG). Minor ratios specify the supremacy of crystallinity owing to the graphitized carbon. For the examined catalysts, the readings of ID/IG were 0.50, 0.92, 0.92, 0.87, and 0.53 for the Ni−MK, Ni−ZH, Ni−EN, Ni−RC, and Ni-DK, in that order. The Ni−Mk and Ni-ZH catalysts exhibited the highest (ID/IG) ratios, indicating low crystallinity owing to the deficiency of graphitized carbon Therefore, the crystallinity of the Ni−Mk and Ni-ZH catalyst is less pronounced than other catalysts, and its graphitization degree was evidently diminished. The degree was obviously diminished.
The thermogravimetric analysis (TGA) was performed on the catalysts after reaction to evaluate the amount of carbon formed on the catalysts. As depicted in Figure 5, the weight loss that started at about 550 °C arises from the removal of different types of carbon species. The amount of carbon deposited for Ni-RC is as follows: 60.0%, 65.8%, and 68.8% for calcination temperatures of 600, 700, and 800 °C, respectively, denoting the rise of calcination temperature, which increases the carbon formation and hence reduces the stability. The TGA values for all catalysts are recorded in Table 2.
Further studies were performed by stabilizing the Ni-RC catalyst and calcining it at 700 °C. The stabilizers used include MgO, SiO2, TiO2, and Y2O3. The activity and stability results of the Ni supported on stabilized zirconia (RC) catalysts calcined at 700 °C are displayed in Figure 6. The Y2O3 stabilized catalyst outperforms the other catalysts, and it gave average conversions of CH4 and CO2 of 67.2% and 72%, respectively. The stability investigation of the best catalyst obtained (Ni-15Y-RC) is given in Figure S5. The performance of Ni-15Y-RC appears to be promising when compared to literature data obtained by other research teams on ZrO2-based catalysts. Table 3 displays the performance comparison to literature data.
The X-ray diffraction analysis of the fresh catalysts was done to determine their crystallinity and structure. The peaks of fresh catalysts are exhibited in Figure 7. The Ni diffractions are specked at two theta values of 37.6° and 42.8°. The peaks at 18.2°, 24.0°, 27.8°, 63.2°, 69.9°, 75.7°, and 82.6° are assigned to the monoclinic ZrO2 (JCPDS file No. 81–1314). Alternatively, the tetragonal ZrO2 diffraction peaks are situated at 29.9°, 34.3°, 35.1°, 50.6°, and 59.8° [37]. It is found that non-stabilized Ni-RC and that of Ti and Mg stabilized catalysts indicate the presence of small NiO peaks. This means the addition of Ti and Mg stabilizers did not affect the structural distribution of NiO, while on the other hand, the absence of NiO peaks for the Si and Y stabilizers indicates the high dispersion of species in the zirconia matrix.
The TEM analysis of the best stabilized Ni-15Y-RC catalyst was performed for a better understanding of the surface morphology and carbon formation on the surface. Figure 8a,a’ shows the fresh catalyst, illustrating that the catalyst has good scattering and smaller active metal particles size, while Figure 8b,b’ shows the used catalyst where sintering or agglomeration has been noted, and the thin layers of multiwall carbon nanotubes.

3. Experimental

3.1. Materials

Nickel nitrate hexahydrate [Ni(NO3)2.6H2O, 98%, Alfa Aesar]. ZrO2 supports were obtained from different sources: MK (Canada); RC (Japan); ZH (Japan); DK (Japan), and EN (China). All the chemicals were of analytical grade and used without further purification. A fixed Ni loading of 5 wt.% was adopted.
The catalysts were prepared via the wet-impregnation technique. First, the solution with a stoichiometric amount of [Ni (NO3)3 9H2O] was prepared in double-distilled water, then ZrO2/or ZrO2 stabilized support was impregnated with the previously prepared solution of active metal salt. Then, mixing was carried out under constant stirring at 80 °C for at least 3 h. The mixture was then dried overnight at 120 °C and followed by calcination at the desired temperature for 3 h. The 5%Ni/ZrO2 catalysts are referred to as Ni-MK, Ni-RC, Ni-ZH, Ni-DK, and Ni-EN. While the Ni-RC stabilized with 15%M (M = Mg, Ti, Si, and Y) catalysts are referred to as Ni-15Mg-RC, Ni-15Ti-RC, Ni-15Si-RC, and Ni-15Y-RC. Table 4 shows the textural properties of various ZrO2 supports. The mathematical equation below is used for calculating the conversion of reactants (CH4 and CO2).
R e a c t a n t   c o n v e r s i o n   % = m o l   ( R ) i n m o l   ( R ) o u t m o l   ( R ) i n × 100 .

3.2. Catalytic Testing

Dry reforming of CH4 over 5%Ni/ZrO2 catalysts was performed at 700 °C and at atmospheric pressure in a vertical stainless steel fixed-bed tubular (i.d., 9.1 mm; length, 0.3 m) micro-reactor (PID Eng and Tech micro activity). The set-up for the experiment, as shown in Figure S1, consists of three sections: the feed section where the cylinders containing feed gas such as CH4, CO2, N2 are situated; the reaction section assembly, which consists of a microtubular reactor. The reactor temperature is measured with the aid of a K-type-thermocouple touching the catalyst bed. The product gases from the reaction section were sent for analysis to estimate the catalytic activity; the analysis section, where the reaction products together with feed mixtures were analyzed using an online gas GC system. The thermal conductivity detector contained in the GC was used to analyze the gas composition. Activity tests were performed using 0.1 g of catalyst. A thermocouple, placed axially at the center of the catalyst bed measures the temperature during the reaction. Prior to each test, the catalysts were activated with H2 (20 mL/min) for 1 h at 700 °C. Experiments were carried out using a feed gas mixture (CH4, CO2, and N2) at a ratio of 30/30/10 and space velocity: 42,000 mL/(h·gcat). The outlet gas composition was analyzed by online gas chromatography (GC-2014 Shimadzu Corp., Kyoto, Japan) fitted out with a thermal conductivity detector. CH4 and CO2 conversions were computed:
CH4 conversion (%) = (CH4,in − CH4,out )/CH4,in × 100; CO2 conversion (%) = (CO2,in − CO2,out)/CO2,in × 100; The deactivation factor (DF) after 7 h reaction time is calculated: DF (%) = (Starting methane conversion − Last methane conversion)/Starting methane conversion × 100.

3.3. Catalyst Characterization

The catalysts were characterized. The specific surface area of catalysts was done by nitrogen physisorption at −196 °C. A Micromeritics Tristar II 3020 unit was employed to acquire the surface area by standard Brunauer–Emmett–Teller (BET). Thermogravimetric analysis in atmospheric air via an EXSTAR SII TG/DTA 7300 analyzer was used to calculate the amount of carbon formation on the surface of used catalysts. A laser Raman (NMR-4500) spectrometer (JASCO, Tokyo, Japan) was used to provide the graphitization degree and the type of carbon deposited over the used catalysts. An excitation beam with a 532 nm wavelength was used. The structure of the spent samples was seized using a transmission electron microscope (TEM) “120 kV JEOL JEM-2100F”. TEM micrographs were documented at 120 kV. The crystal structure and the phases of fresh catalysts were examined with the X-ray diffractometer. A Miniflex Rigaku diffractometer, which has Cu Ka X-ray radiation, functioned at 40 kV and 40 mA, was used for the test. Data gathering was performed using a 2 h angle span of 10–85° and a step magnitude of 0.01°.

4. Conclusions

The current work comprised the development of 5%Ni-ZrO2 and its stabilized catalysts through the impregnation method for dry reforming of methane. ZrO2 acquired from different sources was employed as support. The stabilizers used were MgO, TiO2, SiO2, and Y2O3. The examination focused on the impact of using supports attained from diverse sources such as Canada, Japan, and China. Then, the performance was tested. The obtained catalysts were calcined at 600, 700, and 800 °C. BET, FT-IR, Raman, XRD, TEM, and TGA characterizations were used to investigate spent and fresh catalysts. The BET results point out average values ranging between 3 and 29 m2/g and showed that generally increasing the calcination temperature decreased the BET values. The Ni-RC catalyst presented the highest surface area and the highest activity performance towards DRM and the lowest deactivation factor of 10.3%. Ni-MK and Ni-EN catalysts presented the lowest average weight loss of 24.3% and 23.4%. This was due to their low activity behavior. The results of the FT-IR displayed the relevant characteristic transmissions bands for the Zr-O and Ni-O stretching vibrations at their respective locations. The Raman bands of the used samples assessed the amount of crystallinity of the carbon formed in the reaction. The crystallinity of the Ni−MK is more noticeable than other catalysts, and its graphitization degree was markedly improved. The Ni stabilized zirconia catalyst revealed the preference of the Y2O3 stabilizer. The XRD of the fresh stabilized catalysts depicted the presence of both monoclinic and tetragonal phases of ZrO2 besides the NiO. The TEM analysis of the best stabilized Ni-15Y-RC catalyst revealed good scattering and small size of Ni in the fresh catalyst and Ni sintering or agglomeration with thin layers of multiwall carbon nanotubes in the used catalyst. The RC-100 supported Ni catalyst calcined at 700 °C gave the best performance in the DRM process.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/catal12040361/s1. Figure S1: Experimental set-up; Figure S2: Pore size distribution for catalysts calcined at 600 °C; Figure S3: Pore size distribution for catalysts calcined at 700 °C; Figure S4: Pore size distribution for catalysts calcined at 800 °C; Figure S5: Stability study of the best catalyst (Ni-15Y-RC).

Author Contributions

A.A.I., A.S.A.-F. and M.S.L.: synthesized the catalysts, performed all the experiments, and characterization tests and wrote the manuscript. Y.A.A., S.F.A., A.A.A.-Z. and S.B.A., did characterization tests and wrote parts of the manuscript. A.S.A.-A., A.E.A. and A.H.F.; writing—review and editing, contributed to the analysis of the data, and proofread the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

The authors would like to appreciate sincerely to Researchers Supporting Project number (RSP-2021/368), King Saud University, Riyadh, Saudi Arabia.

Acknowledgments

The authors would like to extend their sincere appreciation to Researchers Supporting Project number (RSP-2021/368), King Saud University, Riyadh, Saudi Arabia.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Teh, L.P.; Setiabudi, H.D.; Timmiati, S.N.; Aziz, M.A.A.; Annuar, N.H.R.; Ruslan, N.N. Recent progress in ceria-based catalysts for the dry reforming of methane: A review. Chem. Eng. Sci. 2021, 242, 116606. [Google Scholar] [CrossRef]
  2. García-Diéguez, M.; Pieta, I.S.; Herrera, M.C.; Larrubia, M.A.; Alemany, L.J. Nanostructured Pt- and Ni-based catalysts for CO2-reforming of methane. J. Catal. 2010, 270, 136–145. [Google Scholar] [CrossRef]
  3. Karmaker, A.K.; Rahman, M.M.; Hossain, M.A.; Ahmed, M.R. Exploration and corrective measures of greenhouse gas emission from fossil fuel power stations for Bangladesh. J. Clean. Prod. 2020, 244, 118645. [Google Scholar] [CrossRef]
  4. Shahzad, U.; Fareed, Z.; Shahzad, F.; Shahzad, K. Investigating the nexus between economic complexity, energy consumption and ecological footprint for the United States: New insights from quantile methods. J. Clean. Prod. 2021, 279, 123806. [Google Scholar] [CrossRef]
  5. Springmann, M.; Mason-D’Croz, D.; Robinson, S.; Wiebe, K.; Godfray, H.C.J.; Rayner, M.; Scarborough, P. Mitigation potential and global health impacts from emissions pricing of food commodities. Nat. Clim. Chang. 2016, 7, 69–74. [Google Scholar] [CrossRef]
  6. Watson, J.; Zhang, Y.; Si, B.; Chen, W.T.; de Souza, R. Gasification of biowaste: A critical review and outlooks. Renew. Sustain. Energy Rev. 2018, 83, 1–17. [Google Scholar] [CrossRef]
  7. Sikarwar, V.S.; Zhao, M.; Fennell, P.S.; Shah, N.; Anthony, E.J. Progress in biofuel production from gasification. Prog. Energy Combust. Sci. 2017, 61, 189–248. [Google Scholar] [CrossRef]
  8. Al-Fatish, A.S.A.; Ibrahim, A.A.; Fakeeha, A.H.; Soliman, M.A.; Siddiqui, M.R.H.; Abasaeed, A.E. Coke formation during CO2 reforming of CH4 over alumina-supported nickel catalysts. Appl. Catal. A Gen. 2009, 364, 150–155. [Google Scholar] [CrossRef]
  9. 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. [Google Scholar] [CrossRef]
  10. Patel, R.; Al-Fatesh, A.S.; Fakeeha, A.H.; Arafat, Y.; Kasim, S.O.; Ibrahim, A.A.; Al-Zahrani, S.A.; Abasaeed, A.E.; Srivastava, V.K.; Kumar, R. Impact of ceria over WO3–ZrO2 supported Ni catalyst towards hydrogen production through dry reforming of methane. Int. J. Hydrogen Energy 2021, 46, 25015–25028. [Google Scholar] [CrossRef]
  11. Al-Fatesh, A.S.; Fakeeha, A.H.; Ibrahim, A.A.; Abasaeed, A.E. Ni supported on La2O3+ZrO2 for dry reforming of methane: The impact of surface adsorbed oxygen species. Int. J. Hydrogen Energy 2021, 46, 3780–3788. [Google Scholar] [CrossRef]
  12. 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]
  13. Tsyganok, A.I.; Inaba, M.; Tsunoda, T.; Suzuki, K.; Takehira, K.; Hayakawa, T. Combined partial oxidation and dry reforming of methane to synthesis gas over noble metals supported on Mg–Al mixed oxide. Appl. Catal. A Gen. 2004, 275, 149–155. [Google Scholar] [CrossRef]
  14. Das, S.; Thakur, S.; Bag, A.; Gupta, M.S.; Mondal, P.; Bordoloi, A. Support interaction of Ni nanocluster based catalysts applied in CO2 reforming. J. Catal. 2015, 330, 46–60. [Google Scholar] [CrossRef]
  15. Singha, R.K.; Yadav, A.; Shukla, A.; Iqbal, Z.; Pendem, C.; Sivakumar, K.; Bal, R. Promoting Effect of CeO2 and MgO for CO2 Reforming of Methane over Ni-ZnO Catalyst. ChemistrySelect 2016, 1, 3075–3085. [Google Scholar] [CrossRef]
  16. Abdel Karim Aramouni, N.; Zeaiter, J.; Kwapinski, W.; Leahy, J.J.; Ahmad, M.N. Molybdenum and nickel-molybdenum nitride catalysts supported on MgO-Al2O3 for the dry reforming of methane. J. CO2 Util. 2021, 44, 101411. [Google Scholar] [CrossRef]
  17. Papadopoulou, C.; Matralis, H.; Verykios, X. Utilization of Biogas as a Renewable Carbon Source: Dry Reforming of Methane. Catal. Altern. Energy Gener. 2012, 9781461403449, 57–127. [Google Scholar] [CrossRef]
  18. McCullough, J.D.; Trueblood, K.N. The crystal structure of baddeleyite (monoclinic ZrO2). Acta Crystallogr. 1959, 12, 507–511. [Google Scholar] [CrossRef]
  19. Chen, A.; Zhou, Y.; Miao, S.; Li, Y.; Shen, W. Assembly of monoclinic ZrO2 nanorods: Formation mechanism and crystal phase control. CrystEngComm 2016, 18, 580–587. [Google Scholar] [CrossRef]
  20. Smith, D.K.; Cline, C.F. Verification of Existence of Cubic Zirconia at High Temperature. J. Am. Ceram. Soc. 1962, 45, 249–250. [Google Scholar] [CrossRef]
  21. Lin, C.; Zhang, C.; Lin, J. Phase Transformation and Photoluminescence Properties of Nanocrystalline ZrO2 Powders Prepared via the Pechini-type Sol−Gel Process. J. Phys. Chem. C 2007, 111, 3300–3307. [Google Scholar] [CrossRef]
  22. Shukla, S.; Seal, S.; Vij, R.; Bandyopadhyay, S.; Rahman, Z. Effect of Nanocrystallite Morphology on the Metastable Tetragonal Phase Stabilization in Zirconia. Nano Lett. 2002, 2, 989–993. [Google Scholar] [CrossRef]
  23. 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]
  24. Yang, E.H.; Moon, D.J. CO2 Reforming of Methane over Ni0/La2O3 Catalyst without Reduction Step: Effect of Calcination Atmosphere. Top. Catal. 2017, 60, 697–705. [Google Scholar] [CrossRef]
  25. Zhang, M.; Zhang, J.; Zhou, Z.; Zhang, Q.; Tan, Y.; Han, Y. Effects of calcination atmosphere on the performance of the co-precipitated Ni/ZrO2 catalyst in dry reforming of methane. Can. J. Chem. Eng. 2021. [Google Scholar] [CrossRef]
  26. Ji, B.; Li, Q.; Huang, Q. Enhanced leaching recovery of rare earth elements from a phosphatic waste clay through calcination pretreatment. J. Clean. Prod. 2021, 319, 128654. [Google Scholar] [CrossRef]
  27. Aramouni, N.A.K.; Zeaiter, J.; Kwapinski, W.; Leahy, J.J.; Ahmad, M.N. Trimetallic Ni-Co-Ru catalyst for the dry reforming of methane: Effect of the Ni/Co ratio and the calcination temperature. Fuel 2021, 300, 120950. [Google Scholar] [CrossRef]
  28. Fertal, D.R.; Monai, M.; Proaño, L.; Bukhovko, M.P.; Park, J.; Ding, Y.; Weckhuysen, B.M.; Banerjee, A.C. Calcination temperature effects on Pd/alumina catalysts: Particle size, surface species and activity in methane combustion. Catal. Today 2021, 382, 120–129. [Google Scholar] [CrossRef]
  29. Wang, J.; Dong, X.; Wang, Y.; Li, Y. Effect of the calcination temperature on the performance of a CeMoOx catalyst in the selective catalytic reduction of NOx with ammonia. Catal. Today 2015, 245, 10–15. [Google Scholar] [CrossRef]
  30. Liu, Q.; Bian, C.; Jin, Y.; Pang, L.; Chen, Z.; Li, T. Influence of calcination temperature on the evolution of Fe species over Fe-SSZ-13 catalyst for the NH3-SCR of NO. Catal. Today 2020, 388–389, 158–167. [Google Scholar] [CrossRef]
  31. 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. Hydrogen Energy 2021, 46, 31041–31053. [Google Scholar] [CrossRef]
  32. Zhang, T.; Zhang, Y.; Ning, P.; Wang, H.; Ma, Y.; Xu, S.; Liu, M.; Zhang, Q.; Xia, F. The property tuning of NH3-SCR over iron-tungsten catalyst: Role of calcination temperature on surface defect and acidity. Appl. Surf. Sci. 2021, 538, 147999. [Google Scholar] [CrossRef]
  33. Sun, Y.; Zhang, G.; Xu, Y.; Zhang, R. Catalytic performance of dioxide reforming of methane over Co/AC-N catalysts: Effect of nitrogen doping content and calcination temperature. Int. J. Hydrogen Energy 2019, 44, 16424–16435. [Google Scholar] [CrossRef]
  34. Chan, K.S.; Chuah, G.K.; Jaenicke, S. Preparation of stable, high surface area zirconia. J. Mater. Sci. Lett. 1994, 13, 1579–1581. [Google Scholar] [CrossRef]
  35. Chuah, G.K.; Jaenicke, S.; Cheong, S.A.; Chan, K.S. The influence of preparation conditions on the surface area of zirconia. Appl. Catal. A Gen. 1996, 145, 267–284. [Google Scholar] [CrossRef]
  36. 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]
  37. Sun, J.; Yamaguchi, D.; Tang, L.; Periasamy, S.; Ma, H.; Hart, J.N.; Chiang, K. Enhancement of oxygen exchanging capability by loading a small amount of ruthenium over ceria-zirconia on dry reforming of methane. Adv. Powder Technol. 2021, 103407. [Google Scholar] [CrossRef]
  38. Mesrar, F.; Kacimi, M.; Liotta, L.F.; Puleo, F.; Ziyad, M. Syngas production from dry reforming of methane over ni/perlite catalysts: Effect of zirconia and ceria impregnation. Int. J. Hydrogen Energy 2018, 43, 17142–17155. [Google Scholar] [CrossRef]
  39. Quéré, C.; Andrew, R.; Friedlingstein, P.; Sitch, S.; Hauck, J.; Pongratz, J.; Pickers, P.; Ivar Korsbakken, J.; Peters, G.; Canadell, J.; et al. Global Carbon Budget 2018. Earth Syst. Sci. Data 2018, 10, 2141–2194. [Google Scholar] [CrossRef] [Green Version]
  40. Kim, W.Y.; Lee, Y.H.; Park, H.; Choi, Y.H.; Lee, M.H.; Lee, J.S. Coke tolerance of Ni/Al2O3 nanosheet catalyst for dry reforming of methane. Catal. Sci. Technol. 2016, 6, 2060–2064. [Google Scholar] [CrossRef]
  41. Al-Mubaddel, F.S.; Kumar, R.; Sofiu, M.L.; Frusteri, F.; Ibrahim, A.A.; Srivastava, V.K.; Kasim, S.O.; Fakeeha, A.H.; Abasaeed, A.E.; Osman, A.I.; et al. Optimizing acido-basic profile of support in Ni supported La2O3+Al2O3 catalyst for dry reforming of methane. Int. J. Hydrogen Energy 2021, 46, 14225–14235. [Google Scholar] [CrossRef]
  42. Al-Fatesh, A.S.; Kumar, R.; Kasim, S.O.; Aidid, A.; Anis, I.; Fakeeha, H.; Abasaeed, A.E.; Atia, H.; Armbruster, U.; Kreyenschulte, C.; et al. Effect of Cerium Promoters on an MCM-41-Supported Nickel Catalyst in Dry Reforming of Methane. Cite This Ind. Eng. Chem. Res 2021, 2022, 174. [Google Scholar] [CrossRef]
  43. Pan, C.; Guo, Z.; Dai, H.; Ren, R.; Chu, W. Anti-sintering mesoporous Ni–Pd bimetallic catalysts for hydrogen production via dry reforming of methane. Int. J. Hydrogen Energy 2020. [Google Scholar] [CrossRef]
  44. Setiabudi, H.; Lim, K.; Ainirazali, N.; Chin, S.Y.; Kamarudin, N.H.N. CO2 reforming of CH4 over Ni/SBA-15: Influence of Ni loading on the metal-support interaction and catalytic activity. J. Mater. Environ. Sci. 2017, 8, 573–581. [Google Scholar]
Figure 1. N2 physisorption isotherms for prepared catalysts at different calcination temperatures (A) at 600 °C (B) at 700 °C, and (C) at 800 °C.
Figure 1. N2 physisorption isotherms for prepared catalysts at different calcination temperatures (A) at 600 °C (B) at 700 °C, and (C) at 800 °C.
Catalysts 12 00361 g001
Figure 2. FT-IR spectra of the different catalysts and supports.
Figure 2. FT-IR spectra of the different catalysts and supports.
Catalysts 12 00361 g002
Figure 3. The performance evaluation of the prepared catalysts at different calcination temperatures. (A) Methane conversion, (B) carbon dioxide conversion, (C) carbon deposits, and (D) deactivation factor.
Figure 3. The performance evaluation of the prepared catalysts at different calcination temperatures. (A) Methane conversion, (B) carbon dioxide conversion, (C) carbon deposits, and (D) deactivation factor.
Catalysts 12 00361 g003
Figure 4. Raman spectra of the used catalysts operated at 700 °C for 7 h.
Figure 4. Raman spectra of the used catalysts operated at 700 °C for 7 h.
Catalysts 12 00361 g004
Figure 5. TGA patterns for prepared catalysts at different calcination temperatures (A) at 600 °C, (B) at 700 °C, and (C) at 800 °C, after 7 h reaction time in DRM, operated at 700 °C for 1 atmosphere.; GHSV 42,000 (mL/(g·h)).
Figure 5. TGA patterns for prepared catalysts at different calcination temperatures (A) at 600 °C, (B) at 700 °C, and (C) at 800 °C, after 7 h reaction time in DRM, operated at 700 °C for 1 atmosphere.; GHSV 42,000 (mL/(g·h)).
Catalysts 12 00361 g005
Figure 6. The performance evaluation of stabilized catalysts at 700 °C. (A) Methane and carbon dioxide conversions, (B) carbon deposits and the deactivation factor.
Figure 6. The performance evaluation of stabilized catalysts at 700 °C. (A) Methane and carbon dioxide conversions, (B) carbon deposits and the deactivation factor.
Catalysts 12 00361 g006
Figure 7. XRD patterns for fresh and stabilized catalysts.
Figure 7. XRD patterns for fresh and stabilized catalysts.
Catalysts 12 00361 g007
Figure 8. TEM micrographs and matching particle size distribution for fresh Ni-15Y-RC (a,a’) and used Ni15Y-RC (b,b’) catalyst calcined at 700 °C.
Figure 8. TEM micrographs and matching particle size distribution for fresh Ni-15Y-RC (a,a’) and used Ni15Y-RC (b,b’) catalyst calcined at 700 °C.
Catalysts 12 00361 g008
Table 1. N2 physisorption results and crystalline size of 5%Ni/ZrO2 catalyst for different calcination temperatures.
Table 1. N2 physisorption results and crystalline size of 5%Ni/ZrO2 catalyst for different calcination temperatures.
CatalystCalcination Temp. (°C)BET
(m2/g)
VPore
(cm3/g)
DPore
(nm)
Ni-MK6003.50.0229.9
7003.20.0241.4
8002.70.0234.9
Ni-ZH60043.20.1615.5
7009.80.0524.8
80010.00.0524.8
Ni-DK60019.60.1639.7
70014.80.1239.2
80013.20.1249.7
Ni-EN60024.40.2137.5
70015.50.1338.6
80016.00.1337.6
Ni-RC60050.40.2721.5
70017.70.1231.7
80018.70.1231.1
Table 2. The performance evaluation of catalysts.
Table 2. The performance evaluation of catalysts.
CatalystCalcination Temperature (°C)Conversion (%)Carbon Deposits (%)Deact. Factor (DF) (%)
CH4CO2
InitialFinalInitialFinal
Ni-MK60021.920.132.630.112.78.2
70046.036.354.246.446.721.1
80036.522.939.634.014.037.3
Ni-ZH60059.344.961.355.575.024.3
70062.245.369.557.938.127.2
80055.038.555.149.261.230.0
Ni-DK60052.336.157.749.248.531.0
70050.632.749.345.063.435.4
80047.534.757.248.653.326.9
Ni-EN60044.630.353.242.117.330.5
70042.331.952.843. 925.124.6
80045.128.352.639.327.837.25
Ni-RC60067.646.566.861.760.61.63
70067.446.568.458.265.831.0
80064.476.4 *66.859.5 *68.7−1.7
Coke = obtained from TGA after 7 h reaction time; % DF after 7 h reaction time = (Starting methane conversion—Last methane conversion)/Starting methane conversion × 100; Initial = 20 min; Final = 420 min; * final = 280 min.
Table 3. Performance comparison of the present result with that of other studies.
Table 3. Performance comparison of the present result with that of other studies.
CatalystReaction Temperature (°C)GHSV (mL/(g·h))Average * CH4 Conversion (%)Ref.
5%Ni/La+Zr700800066[36]
0.1%Ni/Ce+Zr850300038[37]
15%Ni/perlite+Zr70060,00079[38]
15%Ni+Co/Al+-Zr-I8502472[39]
5%Ni+Co/Al+Zr5502418[40]
5%Ni/PO4+Zr80013,00045[12]
Ni/Al2O370042,00053[41]
Ni/MCM-4180039,00085[42]
Ni/SiO270024,00041[43]
3Ni/SBA-1580015,00072[44]
Ni/La+Zr700800070[36]
10 wt.% Ni/SA-523970068070[8]
5%Ni/Y+Zr70042,00067Present work
* The mean of initial and final conversions.
Table 4. Textural properties of various ZrO2 supports: BET specific surface area (BET), pore volume (Pv), and pore size (DP).
Table 4. Textural properties of various ZrO2 supports: BET specific surface area (BET), pore volume (Pv), and pore size (DP).
Country of OriginProduct
Brand-Name
Abbreviated
Name
BET (m2/g)Pore-Volume
(cm3/g)
Pore-Diameter
(µm)
Tokyo, Japan
(Daiichi Kigenso Kagaku Kogyo Co., Ltd.)
RC-100
(zirconia oxide)
RC1000.387.2
Z-1325
(zirconia hydroxide)
ZH3710.363.6
DK-1
(zirconia oxide)
DK250.2028.4
Hefei, China
(Anhui Elite Industrial Co., Ltd.)
ELTN
(zirconia oxide)
EN280.2132.5
CanadaMKnano
(zirconia oxide)
MK40.1019.2
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Ibrahim, A.A.; Fakeeha, A.H.; Lanre, M.S.; Al-Awadi, A.S.; Alreshaidan, S.B.; Albaqmaa, Y.A.; Adil, S.F.; Al-Zahrani, A.A.; Abasaeed, A.E.; Al-Fatesh, A.S. The Effect of Calcination Temperature on Various Sources of ZrO2 Supported Ni Catalyst for Dry Reforming of Methane. Catalysts 2022, 12, 361. https://doi.org/10.3390/catal12040361

AMA Style

Ibrahim AA, Fakeeha AH, Lanre MS, Al-Awadi AS, Alreshaidan SB, Albaqmaa YA, Adil SF, Al-Zahrani AA, Abasaeed AE, Al-Fatesh AS. The Effect of Calcination Temperature on Various Sources of ZrO2 Supported Ni Catalyst for Dry Reforming of Methane. Catalysts. 2022; 12(4):361. https://doi.org/10.3390/catal12040361

Chicago/Turabian Style

Ibrahim, Ahmed Aidid, Anis Hamza Fakeeha, Mahmud Sofiu Lanre, Abdulrhman S. Al-Awadi, Salwa Bader Alreshaidan, Yousef Abdulrahman Albaqmaa, Syed Farooq Adil, Ateyah A. Al-Zahrani, Ahmed Elhag Abasaeed, and Ahmed S. Al-Fatesh. 2022. "The Effect of Calcination Temperature on Various Sources of ZrO2 Supported Ni Catalyst for Dry Reforming of Methane" Catalysts 12, no. 4: 361. https://doi.org/10.3390/catal12040361

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