Next Article in Journal
Economy Assessment for the Chiral Amine Production with Comparison of Reductive Amination and Transamination Routes by Multi-Enzyme System
Next Article in Special Issue
Effect of the Ni/Al Ratio on the Performance of NiAl2O4 Spinel-Based Catalysts for Supercritical Methylcyclohexane Catalytic Cracking
Previous Article in Journal
Modeling and Experimental Studies on Adsorption and Photocatalytic Performance of Nitrogen-Doped TiO2 Prepared via the Sol–Gel Method
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

The Effects of CeO2 and Co Doping on the Properties and the Performance of the Ni/Al2O3-MgO Catalyst for the Combined Steam and CO2 Reforming of Methane Using Ultra-Low Steam to Carbon Ratio

by
Nichthima Dharmasaroja
1,2,
Tanakorn Ratana
1,2,
Sabaithip Tungkamani
1,2,
Thana Sornchamni
3,
David S. A. Simakov
4 and
Monrudee Phongaksorn
1,2,*
1
Department of Industrial Chemistry, King Mongkut’s University of Technology North Bangkok, Bangkok 10800, Thailand
2
Research and Development Center for Chemical Engineering Unit Operation and Catalyst Design (RCC), King Mongkut’s University of Technology North Bangkok, Bangkok 10800, Thailand
3
PTT Research and Technology Institute, PTT Public Company Limited, Wangnoi, Ayutthaya 13170, Thailand
4
Department of Chemical Engineering, University of Waterloo, Waterloo, ON N2L 3G1, Canada
*
Author to whom correspondence should be addressed.
Catalysts 2020, 10(12), 1450; https://doi.org/10.3390/catal10121450
Submission received: 18 November 2020 / Revised: 8 December 2020 / Accepted: 9 December 2020 / Published: 11 December 2020
(This article belongs to the Special Issue Nickel-Based Catalysts for Hydrocarbon Fuel Reforming)

Abstract

:
In this paper, the 10 wt% Ni/Al2O3-MgO (10Ni/MA), 5 wt% Ni-5 wt% Ce/Al2O3-MgO (5Ni5Ce/MA), and 5 wt% Ni-5 wt% Co/Al2O3-MgO (5Ni5Co/MA) catalysts were prepared by an impregnation method. The effects of CeO2 and Co doping on the physicochemical properties of the Ni/Al2O3-MgO catalyst were comprehensively studied by N2 adsorption-desorption, X-ray diffraction (XRD), transmission electron microscopy (TEM), H2 temperature programmed reduction (H2-TPR), CO2 temperature programmed reduction (CO2-TPD), and thermogravimetric analysis (TGA). The effects on catalytic performance for the combined steam and CO2 reforming of methane with the low steam-to-carbon ratio (S/C ratio) were evaluated at 620 °C under atmospheric pressure. The appearance of CeO2 and Co enhanced the oxygen species at the surface that decreased the coke deposits from 17% for the Ni/MA catalyst to 11–12% for the 5Ni5Ce/MA and 5Ni5Co/MA catalysts. The oxygen vacancies in the 5Ni5Ce/MA catalyst promoted water activation and dissociation, producing surface oxygen with a relatively high H2/CO ratio (1.6). With the relatively low H2/CO ratio (1.3), the oxygen species at the surface was enhanced by CO2 activation-dissociation via the redox potential in the 5Ni5Co/MA catalyst. The improvement of H2O and CO2 dissociative adsorption allowed the 5Ni5Ce/MA and 5Ni5Co/MA catalysts to resist the carbon formation, requiring only a low amount of steam to be added.

Graphical Abstract

1. Introduction

The combined steam and CO2 reforming of methane (CSCRM) (Equation (1)) is a process that combines steam reforming of methane (Equation (2)) and CO2 reforming of methane (Equation (3)) in one process. The CSCRM has received noticeable attention as it consumes two main greenhouse gases (CH4 and CO2) with water vapor to produce the synthesis gas (a mixture of H2 and CO) [1,2,3,4]. Although CSCRM can control an H2/CO ratio of 2 in the syngas product by the inlet feed composition [2,5,6], the obstacles for a commercial CSCRM consist of the catalyst deactivation and the energy consumption. Main causes of the catalyst deactivation are the coke deposition, which is a by-product from side reactions (the CH4 decomposition (Equation (4)) and the Boudouard reaction (Equation (5)) [7,8,9]. To prevent the formation of carbon, steam in the feed must be sufficient. However, the quality of energy consumption can be over expected due to the evaporation of water. Therefore, the development of high-performance catalysts for CSCRM operating at a low steam-to-carbon ratio could be a critical challenge of syngas production technologies.
The non-noble metal catalysts, Ni-based catalysts, have focused the high catalytic performance and low cost as compared to the noble metal [1,10,11,12,13,14]. According to numerous works on the catalyst development, Ni-bimetallic catalysts have been proposed as carbon tolerance catalysts for methane reforming. Noble metals and non-noble metals are commonly used with Ni for this type of catalysts [15,16]. Among several metals, the oxygen vacancy sites can be created when Ce4+ in the oxide form of cerium (CeO2) transforms to Ce3+. The released oxygen then removes the carbonaceous species on the Ni surface, suppressing the coke formation [7,17,18,19,20]. Furthermore, the strong interaction between Ni and Ce prevents the agglomeration of Ni nanoparticles in the Ni-Ce/montmorillonite catalyst [21]. Shan et al. [22] suggested that Ni2+ ions can be incorporated into the lattice of CeO2 and oxygen vacancies that are simultaneously generated. The presence of CeO2 decreases the formation of NiAl2O4 due to the formation of CeAlO3 and Ce1−XNiXO2 [23,24,25]. Moreover, CO2 can transform into carboxylate species and react with surface hydroxyls to produce formate (HCOO) species on Ce3+ that promoted the water gas shift reaction [25].
Cobalt (Co) is also an alternative metal component for bimetallic Ni catalysts because of the cobalt oxide properties. Co as an oxide form possess a weak metal-oxygen bond strength with high turnover frequency for a redox reaction via the reduction pathway of Co3O4 → CoO → Co° [26]. Thus, cobalt oxides (CoOx) play a major role in the oxidation of carbon species via redox reactions that decrease the carbon deposition [27]. During reduction, the combination of Ni-Co metal creates the formation of the Ni-Co alloy structure [11]. The Ni-Co alloy enhances the prevention of crystal growth, the adsorption of oxygen species, and the distribution of active sites, resulting in higher reactant conversions [28]. Moreover, the partial Co metal diffusing from the bulk structure to the spinel-like phases (CoAl2O4) deducts the sintering of active metal during the reaction [29].
The reactant ratio adjustment in the feed plays a vital role in the carbon deposition control. Although CH4 is the source of hydrogen for H2 production, CH4 dissociation is the main reaction of coke formation, especially under the relatively low CO2 content conditions. Certain research articles concluded that the satisfaction ratio of CO2 to CH4 (CO2/CH4) is greater than unity [26]. Steam in the feed promotes the steam reforming of methane as well as the water gas shift reaction, which increases the H2/CO ratio in the syngas product. The addition of H2O molecules also provides more of an oxygen source for the carbon removal mechanism (Equation (6)) [11,27,28,30]. The extra energy to evaporate a large volume of steam at the inlet and to separate water from the gaseous outlet is concerning [11,29].
3CH4 + CO2 + 2H2O→8H2 + 4CO  △H°298 K = +659 kJ/mole
CH4 + H2O → 3H2 + CO      △H°298 K = +206 kJ/mole
CH4 + CO2 → 3H2 + CO      △H°298 K = +247 kJ/mole
CH4 → C + 2H2          △H°298 K = +74.8 kJ/mole
2CO → C + CO2          △H°298 K = −173.3 kJ/mole
C + H2O → H2 + CO        △H°298 K = +131.3 kJ/mole
This work evaluated the effect of CeO2 and Co promoters over the CSCRM (using a low steam- to-carbon ratio (S/C ratio) that accompanies CO2 and H2O oxidants) catalytic performance of the Ni/MgO-Al2O3. For this propose, 10 wt% Ni/MgO-Al2O3 (10Ni/MA), 5 wt% Ni–5 wt% Ce/MgO-Al2O3 (5Ni5Ce/MA), and 5 wt% Ni–5 wt% Co/MgO-Al2O3 (5Ni5Co/MA) were prepared by impregnation methods. The CSCRM performances of all catalyst samples were investigated at 620 °C under atmospheric pressure. The carbon accumulation on the surface of the spent catalysts was investigated using thermogravimetric analysis (TGA). The correlation between catalytic performance and properties, characterized by N2 adsorption-desorption, X-ray diffraction (XRD), transmission electron microscope (TEM), H2 temperature programmed reduction (H2-TPR), and CO2 temperature programmed desorption (CO2-TPD), were revealed.

2. Results and Discussion

The crystalline phases of the calcined catalysts were analyzed by XRD (Figure 1). In all catalysts, the diffraction peaks of meixnerite Mg6Al2(OH)18·4H2O were indicated at 2θ of 11.5°, 22.5°, and 35.0° [31]. The characteristic diffraction peaks at 2 theta of 36.8°, 44.8°, and 65.3° can be assigned to the spinel phase of support (MgAl2O4), existing as an overlap with the NiAl2O4 (or CoAl2O4) spinel. The presence of NiAl2O4 and CoAl2O4 spinel reflects the strong metal-support interaction in the catalysts. The board peaks of CeO2 diffractogram at 2θ of 28.5°, 47.5°, and 56.3° were observed on the 5Ni5Ce/MA catalyst. In the 5Ni5Co/MA catalyst, the peaks of the Co3O4 phase at 2θ of 19.3° and 31.5° were detected. The discrete peaks located at 2θ = 37.0°, 43.0°, 62.4°, 74.8°, and 78.7° corresponding to the NiO phase overlapped with MgO peaks [32].
The N2 adsorption-desorption isotherms and the pore size distribution of 10Ni/MA, 5Ni5Ce/MA, and 5Ni5Co/MA catalysts are shown in Figure 2. The isotherms of all catalysts are categorized as type IV according to the International Union of Pure and Applied Chemistry (IUPAC) classification and the pore sizes of samples are mainly in the range of 4.6–6.1 nm, implying the characteristic of mesoporous materials. Hysteresis loops of types H1 and H3 corresponding to the combination of cylindrical pores and parallel plate-shaped pores are found in all catalysts [33]. The surface area and total pore volume of samples are listed in Table 1. Compared to the monometallic Ni supported catalyst (10Ni/MA), bimetallic catalysts (5Ni5Ce/MA and 5Ni5Co/MA) provided a larger catalyst surface and pore volume. TEM pictures of all samples presented in Figure 3 suggests a better dispersion of the metal species with a smaller average metal size on 5Ni5Ce/MA and 5Ni5Co/MA catalysts as compared to the 10Ni/MA catalyst, which agree with N2 adsorption-desorption results. It can be explained that the presence of CeO2 and Co3O4 in Ni/MA reduces the NiO agglomeration, resulting in the increase in Ni dispersion, surface area, and the pore volume.
The TPR profile of the 10Ni/MA catalyst (Figure 4) showed three peaks associated with the reduction of NiO located on the surface of MA support (a sharp reduction peak at 350 °C), the reduction of NiO nanoparticles confined in the MA support (a peak shoulder at 450 °C), and the reduction of Mg(Ni,Al)O (the main peak at 790 °C), which is a form of the strong metal-support interaction [34,35]. The 5Ni5Ce/MA catalyst revealed four reduction board peaks at 295 °C, 410 °C, 530 °C, and 845 °C. The lowest temperature peak is preliminarily ascribed to the oxygen mobility on the surface catalyst that substitutes the Ni ion incorporated with CeO2 at the surface and the reduction of Ce4+ located on the surface [24,36,37]. The second and third peaks can be tentatively assigned to the reduction of Ni-CeOx solid solution. The highest temperature peak could be the reduction peak of Ce species simultaneously with the reduction peak of Ni species strongly interacting with the support. Compared to the TPR profile of the 10N/MA, the reduction peak at high temperature in the TPR profile of 5Ni5Ce/MA shifted to higher temperature because the smaller Ni particle size in the 5Ni5Ce/MA catalyst led to the stronger metal-support interaction [38]. As seen in the TPR profile of 5Ni5Co/MA, the low-temperature signal at about 270 °C corresponded to the simultaneous reduction of NiO to Ni and Co3O4 to CoO. The shoulder at 370 °C related to the reduction of CoO to Co species [39]. The peak at 790 °C starting at about 500 °C is attributed to the reduction of NiAl2O4 or the CoAl2O4 spinel phase. The reduction peak at low temperature in the TPR profile of 5Ni5Co/MA catalyst shifted to a lower temperature compared to 10Ni/MA. It implied to the weak interaction between metal and support, which can be attributed to the formation of the Ni–Co alloy [40,41,42,43].
The strength and quantity of basic sites on the surface of catalysts were evaluated from the CO2 desorption temperature and the desorption areas of peaks in the CO2-TPD profiles (Figure 5), respectively. Figure 5 reflected three types of basic sites. In each CO2-TPD profile of the reduced catalysts, the peak at the lowest desorption temperature represents the weakly chemisorbed CO2 on basic sites connected to the Brönsted hydroxyl groups. The peak at the middle temperature refers to medium base sites comprised to the Lewis acid-base pairing, and the highest temperature peak is assigned to strong basic sites related to the low-coordination surface oxygen (O2−) anions [44,45]. The total number of basic sites of 10Ni/MA, 5Ni5Ce/MA, and 5Ni5Co/MA catalysts were 0.19, 0.22, and 0.16 mmol/gcat (Table 1), respectively. The 5Ni5Ce/MA catalyst showed the highest number of basic sites because of the oxygen mobility from the redox property of Ce metal. Among all catalysts, 5Ni5Ce/MA and 5Ni5Co/MA provided a greater number of medium basic sites (correlating to the oxophilicity of the surface metal in the catalyst). The oxygen mobility and the stronger oxophilicity suggest more coverage of O* species that can improve the removal of coke deposition [45].
CSCRM tests operated with the low steam to carbon ratio (S/C = 0.28) at isothermal condition of 620 °C under the ambient pressure for 6 h were demonstrated. The catalytic performance of the catalyst samples was evaluated in term of the CH4 and CO2 conversions as well as the H2/CO ratio (Figure 6a–c). The results of the 10Ni/MA catalyst can be separated into two periods of time-on-steam described as part I and II. The part I (0–125 min) represented the different conversions for CH4 and CO2 with a low H2/CO ratio (<1). Compared to part I, part II (125–360 min) showed the higher CH4 conversion, the lower CO2 conversion, and the higher H2/CO ratio (>1). It referred that the CO2 reforming of methane dominated the overall process at the initial time. An increase of CH4 conversion with a decrease of CO2 conversion (part II) indicated that the consumed CH4 reacted with steam and CO2 because the steam and CO2 acted as a co-oxidant [46,47], resulting in a higher H2/CO ratio of 1 < × < 1.6. This evidence reflected the sufficient time for H2O dissociative adsorption for the low S/C condition. Considering the performance of the 5Ni5Ce/MA catalyst, CH4 conversion was less by half compared to the 10Ni/MA because of the decrease in active metal content. The 5Ni5Ce/MA catalyst showed the lowest CO2 conversion with the highest H2/CO ratio, suggesting the most occurrence of steam reforming of methane during the CSCRM process on 5Ni5Ce/MA among these catalysts. It can be explained that the oxygen mobility in 5Ni5Ce/MA enhances the water association-dissociation (Ce2O3 + H2O → CeO2 + H2) on the surface of the catalyst [21,48,49]. Moreover, CO2 conversions at 90–120 min of the 5Ni5Ce/MA catalyst (relatively fast H2O activation) were increased when compared to the initial time as H2O should be insufficient at a moment (for this low S/C case), resulting in the H2/CO ratio (<1). The resulting trends (conversions and H2/CO ratio) of the 5Ni5Co/MA catalyst were similar to the 10Ni/MA catalyst and reactant conversions were lower than the 10Ni/MA. For CH4 conversions, the cobalt metal was less active for CH4 dissociation than Ni metal because of the higher activation energy of CH4 dissociation [50,51]. In part I, the difference between CO2 and CH4 conversions of 5Ni5Co/MA and 10Ni/MA were similar as well as the H2/CO ratios, indicating the same magnitude of CO2 reforming of methane domination in the process. In part II, the CO2 and CH4 conversions of 5Ni5Co/MA were closer than those of 10Ni/MA and the 5Ni5Co/MA catalyst illustrated the lowest H2/CO ratios. These results expressed the highest magnitude of CO2 reforming of methane domination on the 5Ni5Co/MA catalyst after 125 min time-on-stream. Li et al. [50] reported that the Co metal surface promotes more dissociative adsorption of CO2 than the Ni metal surface due to the oxophilic property of Co, which is in good agreement with CO2-TPD results. It also reveals that the H2O association-dissociation on cobalt metal was complicated, resulting in the lowest H2/CO ratio.
The quantity and the types of carbon deposition on the spent catalysts were elucidated using the TGA results (Figure 7). The percentage of weight loss directly relates to the amount of carbon deposition. TGA profiles of the spent catalysts represented the physically adsorbed water and amorphous carbon (≤250 °C), the graphitic carbon (250 °C–450 °C), and the carbon filament (≥450 °C) on the surface [18,52,53,54]. The last two types, which are not easily oxidizable, have been considered as the major reason for the catalyst deactivation. The total weight loss of spent 10Ni/MA, 5Ni5Ce/MA, and 5Ni5Co/MA catalysts were 17%, 11%, and 12%, respectively. Although the main type of coke in all catalysts was the graphitic carbon, the highest temperature of carbon removal was found from the spent 10Ni/MA catalyst. It implied that carbon deposition was formed more easily via the CH4 dissociation/decomposition on the Ni/MA catalyst. The coke deposited can be deduced by the oxygen intermediate in H2O activation-dissociation for 5Ni5Ce/MA and oxygen intermediates in CO2 activation-dissociation for 5Ni5Co/MA. Consequently, with Ce and Co promoters, the carbon tolerance of Ni/MA catalyst was improved, requiring only a small amount of steam to be added in the feed of the methane, reforming in order to prevent carbon deposition [55,56,57].
A simplified reaction mechanism of CSCRM to produce synthesis gas on a Ni-based catalyst consists of three element reactions: CH4 dissociation/decomposition (Equation (7)), H2O activation-dissociation (Equation (8)), and CO2 activation-dissociation (Equation (9)). CH4 dissociation/decomposition produces carbon surface (C*) and H2. CO2 and H2O activation-dissociation generate an oxygen surface (O*) with CO and H2, respectively. The C* on the surface can be removed by reacting with the O* surface to form CO (Equation (10)). Therefore, the catalytic behavior of the catalyst for the reaction steps can be tentatively evaluated by conversions of reactants, H2/CO ratio, and the quantity of coke formation. The activity toward the CH4 dissociation/decomposition step directly relates to CH4 conversion. The H2O and CO2 activation-dissociation can be preliminary predicted using the H2/CO ratio with the amount of coke. If the high H2/CO ratio with low carbon accumulation occur, the catalyst is selective for H2O activation-dissociation. Likewise, the catalyst is selective for CO2 activation-dissociation when the low H2/CO ratio and low carbon accumulation are founded. According to the results under the low S/C ratio condition of CSCRM process, it suggested that 10Ni/MA is the most active catalyst for CH4 dissociation/decomposition. CeO2 in 5Ni5Ce/MA and Co in 5Ni5Co/MA boosts the H2O and CO2 activation-dissociation, respectively.
CH4 + * → 2H2 + C*
CO2 + * → CO + O*
H2O + * → H2 + O*
C* + O* → CO

3. Materials and Methods

3.1. Catalyst Preparation

The Al2O3-MgO (MA) was synthesized by the sol-gel method. The mixture of aluminium isopropoxide (Acros OrganicsTM, Morris County, NJ, USA 98%), nitric acid (CARLO ERBA Reagents, Val-de-Reuil, France 65%), magnesium ethoxide (Sigma-Aldrich, St. Louis, MO, USA 98%), and distilled water were refluxed at 80 °C for 20 h. The resulting gel was dried at 60 °C for 24 h and calcined at 800 °C for 6 h. The MA support was crushed and sieved to 355–710 µm. The 10 wt% Ni/Al2O3-MgO (10Ni/MA), 5 wt% Ni-5 wt% Ce/Al2O3-MgO (5Ni5Ce/MA), and 5 wt% Ni-5 wt% Co/Al2O3-MgO (5Ni5Co/MA) catalyst were prepared by the incipient wetness impregnation of the MA support with a solution of nitrates. Precursors included Ni(NO3)2.6H2O (Merck 99%), Ce(NO3)3.6H2O (Honeywell FlukaTM, Charlotte, NC, USA 99%), and Co(NO3)3.6H2O (Sigma-Aldish 99%). Then, the wet solid cake was dried at 60 °C for 24 h and calcined at 650 °C for 5 h.

3.2. Catalyst Characterization

The textural properties were characterized by N2 adsorption-desorption isotherms at −196 °C employing a BEL-Japan BELSORP-mini II (BEL JAPAN, INC., Osaka, Japan). Samples were previously in the nitrogen flow at 350 °C for 4 h. The specific surface area and total pore volume were calculated by the Brunauer–Emmett–Teller (BET) method, whereas the pore size distribution was evaluated by the Barrett-Joyner-Halenda (BJH) method.
The crystalline phase compositions were determined by the X-ray powder diffraction (XRD) patterns using a Model D8 Discover (Bruker AXS, Billerica, MA, USA) with Cu-Kα radiation operating at 40 kV and 40 mA using a speed of 0.02°/min.
The metal particle size and metal distribution on the fresh catalysts were investigated by transmission electron microscope (TEM) on a JEOL JEM-2010 (JEOL Ltd., Welwyn Garden City, England) operating at an acceleration voltage of 200 keV. The catalyst powders were suspended in absolute ethanol, sonicated, and deposited on a carbon copper grid (300 Mesh). Grids were dried before TEM characterization.
The reducibility and metal-support interaction were analyzed by the H2 temperature programmed reduction (H2-TPR) on a BELCAT-basic system (BEL JAPAN, INC., Osaka, Japan) equipped with a thermal conductivity detector (TCD). A 50-mg sample was pretreated in an Ar flow at 220 °C for 40 min and cooled to room temperature. Then, 5% H2/Ar gas mixture was passed through the catalyst from room temperature to 900 °C at the rate of 10 °C/min.
The basic site distribution was evaluated using the CO2 temperature programmed desorption (CO2-TPD) data obtained on the BELCAT-basic system. The calcined catalyst (50 mg) was pre-reduced in an H2 pure at 620 °C for 2 h and, subsequently, cooled to room temperature in a He flow. The isothermal adsorption of CO2 gas was performed for 30 min before the sample was flushed in a He flow. The CO2 desorption was then monitored by TCD in a He flow from room temperature to 800 °C with a ramping rate of 10 °C/min.
The carbon deposition on the spent catalysts was analyzed using thermogravimetric analysis (TGA) on a Model TGA/DSC1 (METTLER TOLEDO, Columbus, OH, USA). A quantity of 15 mg of the spent catalysts was combusted under air flow of 30 mL/min from room temperature to 800 °C at a heating rate of 10 °C/min.

3.3. Catalytic Activity Test

CSCRM tests were performed in a fixed-bed reactor at 620 °C under atmospheric pressure for 6 h. A 200-mg catalyst was in situ reduced at 620 °C in a pure H2 flow of 30 mL/min for 6 h. The volumetric ratio of CH4:CO2:H2O:N2 = 3:5:2.26:4 (ultra-low S/C of 0.28) was fed with the flow rate of 60 mL/min. It should be noted that N2 was used as a carrier gas of the steam. After passing through a glass cold trap, the composition of outlet gases was examined using an on-line gas chromatograph (Agilent GC7890A Agilent, Santa Clara, CA, USA) equipped with TCD. The reactant conversions and the H2/CO ratio were calculated as:
%   CH 4   conversion = Flow   rate   CH 4 , in Flow   rate   CH 4 , out Flow   rate   CH 4 , in × 100
%   CO 2   conversion = Flow   rate   CO 2 , in Flow   rate   CO 2 , out Flow   rate   CO 2 , in × 100
H 2 CO   ratio = Flow   rate   of   H 2 , out Flow   rate   of   CO , out

4. Conclusions

The 5 wt% Ni–5 wt% Ce/Al2O3-MgO (5Ni5Ce/MA), 5 wt% Ni–5 wt% Co/Al2O3-MgO (5Ni5Ce/MA), and 10 wt% Ni/Al2O3-MgO (10Ni/MA) catalysts have been synthesized, characterized by a number of analytical techniques, and tested for their catalytic performance in combined steam and CO2 reformation of methane (CSCRM) with the low S/C ratio. The characterization results revealed that the appearance of CeO2 and Co in the Ni/MA catalyst improved the metal dispersion, resulting in the smaller Ni nanoparticle size. The CeO2 promoter raises the number of all types of basic sites, whereas the Co promoter slightly increases the number of medium basic sites because of its oxophilic property.
Under the low S/C ratio condition of CSCRM process at 620 °C, although the highest activity toward CH4 conversion was obtained from the 10Ni/MA catalyst, the 10Ni/MA catalyst exhibited the highest carbon accumulation. The O* species at the surface were insufficient to remove carbon deposits. The addition of each promoter (CeO2 and Co) enhanced the O* species at the surface by improving the activation-dissociation of the different co-oxidant (H2O and CO2), suppressing the carbon accumulation and resulting in the opposite trend of the H2/CO ratio. CeO2 promoted the H2O activation-dissociation, increasing the H2/CO ratio. Co promoted the CO2 activation-dissociation, decreasing the H2/CO ratio.

Author Contributions

Conceptualization, N.D., T.R., S.T., and M.P. Data curation, N.D. and M.P. Investigation, N.D., T.R., S.T., and M.P. Methodology, N.D. and M.P. Formal analysis, N.D. and M.P. Funding support, T.S. Writing—original draft, N.D. and M.P. Writing—review and editing, N.D., T.S., D.S.A.S., and M.P. Supervision, D.S.A.S., S.T., and M.P. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Thailand Science Research and Innovation (TSRI) via Research and Researchers for Industries (RRI) with the PTT Public Company Limited (grant number PHD59I0027).

Conflicts of Interest

The authors declare no conflict of interest.

Nomenclature

BETBrunauer–Emmett–Teller
BJHBarrett–Joyner–Halenda
CSCRMCombined steam and CO2 reforming of methane
CO2-TPDCO2temperature programmed desorption
H2-TPRH2 temperature programmed reduction
IUPACInternational Union of Pure and Applied Chemistry
RWGSReverse water gas shift
S/C ratioSteam-to-carbon (H2O/(CH4 + CO2)) ratio
TCDThermal conductivity detector
TEMTransmission electron microscope
TGAThermogravimetric analysis
XRDX-ray diffraction

References

  1. Jabbour, K. Tuning combined steam and dry reforming of methane for “metgas” production: A thermodynamic approach and state-of-the-art catalysts. J. Energy Chem. 2020, 48, 54–91. [Google Scholar] [CrossRef]
  2. Lee, S.M.; Hwang, I.H.; Kim, S.S. Enhancement of catalytic performance of porous membrane reactor with Ni catalyst for combined steam and carbon dioxide reforming of methane reaction. Fuel Process. Technol. 2019, 188, 197–202. [Google Scholar] [CrossRef]
  3. Jang, W.J.; Jeong, D.W.; Shim, J.O.; Kim, H.M.; Roh, H.S.; Son, I.H.; Lee, S.J. Combined steam and carbon dioxide reforming of methane and side reactions: Thermodynamic equilibrium analysis and experimental application. Appl. Energy 2016, 173, 80–91. [Google Scholar] [CrossRef]
  4. Ryi, S.K.; Lee, S.W.; Park, J.W.; Oh, D.K.; Park, J.S.; Kim, S.S. Combined steam and CO2 reforming of methane using catalytic nickel membrane for gas to liquid (GTL) process. Catal. Today 2014, 236, 49–56. [Google Scholar] [CrossRef]
  5. Jabbour, K.; Massiani, P.; Davidson, A.; Casale, S.; El Hassan, N. Ordered mesoporous “one-pot” synthesized Ni-Mg(Ca)-Al2O3 as effective and remarkably stable catalysts for combined steam and dry reforming of methane (CSDRM). Appl. Catal. B Environ. 2017, 201, 527–542. [Google Scholar] [CrossRef] [Green Version]
  6. Li, L.; Yang, Z.; Chen, J.; Qin, X.; Jiang, X.; Wang, F.; Song, Z.; Ma, C. Performance of bio-char and energy analysis on CH4 combined reforming by CO2 and H2O into syngas production with assistance of microwave. Fuel 2018, 215, 655–664. [Google Scholar] [CrossRef]
  7. Koo, K.Y.; Lee, S.; Jung, U.H.; Roh, H.-S.; Yoon, W.L. Syngas production via combined steam and carbon dioxide reforming of methane over Ni–Ce/MgAl2O4 catalysts with enhanced coke resistance. Fuel Process. Technol. 2014, 119, 151–157. [Google Scholar] [CrossRef]
  8. 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]
  9. Banisharifdehkordi, F.; Baghalha, M. Catalyst deactivation in industrial combined steam and dry reforming of natural gas. Fuel Process. Technol. 2014, 120, 96–105. [Google Scholar] [CrossRef]
  10. Holladay, J.D.; Hu, J.; King, D.L.; Wang, Y. An overview of hydrogen production technologies. Catal. Today 2009, 139, 244–260. [Google Scholar] [CrossRef]
  11. Gangadharan, P.; Kanchi, K.C.; Lou, H.H. Evaluation of the economic and environmental impact of combining dry reforming with steam reforming of methane. Chem. Eng. Res. Des. 2012, 90, 1956–1968. [Google Scholar] [CrossRef]
  12. Álvarez, M.A.; Centeno, M.Á.; Odriozola, J.A. Ru-Ni Catalyst in the Combined Dry-Steam Reforming of Methane: The Importance in the Metal Order Addition. Top. Catal. 2016, 59, 303–313. [Google Scholar] [CrossRef]
  13. Chen, C.; Wang, X.; Chen, X.; Liang, X.; Zou, X.; Lu, X. Combined steam and CO2 reforming of methane over one-pot prepared Ni/La-Si catalysts. Int. J. Hydrog. Energy 2019, 44, 4780–4793. [Google Scholar] [CrossRef]
  14. Al-Fatesh, A.; Singh, S.K.; Kanade, G.S.; 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]
  15. 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]
  16. Abdel, N.; Aramouni, K.; Touma, J.G.; Tarboush, B.A.; Zeaiter, J.; Ahmad, M.N. Catalyst design for dry reforming of methane: Analysis review. Renew. Sustain. Energy Rev. 2018. [Google Scholar] [CrossRef]
  17. Ay, H.; Üner, D. Dry reforming of methane over CeO2 supported Ni, Co and Ni–Co catalysts. Appl. Catal. B Environ. 2015, 179, 128–138. [Google Scholar] [CrossRef]
  18. Damyanova, S.; Pawelec, B.; Palcheva, R.; Karakirova, Y.; Sanchez, M.C.C.; Tyuliev, G.; Gaigneaux, E.; Fierro, J.L.G. Structure and surface properties of ceria-modified Ni-based catalysts for hydrogen production. Appl. Catal. B Environ. 2018, 225, 340–353. [Google Scholar] [CrossRef]
  19. Wang, L.; Meng, F.; Li, K.; Lu, F. Characterization and optical properties of pole-like nano-CeO2 synthesized by a facile hydrothermal method. Appl. Surf. Sci. 2013, 286, 269–274. [Google Scholar] [CrossRef]
  20. Setiabudi, H.D.; Chong, C.C.; Abed, S.M.; Teh, L.P.; Chin, S.Y. Comparative study of Ni-Ce loading method: Beneficial effect of ultrasonic-assisted impregnation method in CO2 reforming of CH4 over Ni-Ce/SBA-15. J. Environ. Chem. Eng. 2018, 6, 745–753. [Google Scholar] [CrossRef]
  21. Li, L.; Tang, D.; Song, Y.; Jiang, B.; Zhang, Q. Hydrogen production from ethanol steam reforming on Ni-Ce/MMT catalysts. Energy 2018, 149, 937–943. [Google Scholar] [CrossRef]
  22. Shan, W.; Luo, M.; Ying, P.; Shen, W.; Li, C. Reduction property and catalytic activity of Ce1−XNiXO2 mixed oxide catalysts for CH4 oxidation. Appl. Catal. A Gen. 2003, 246, 1–9. [Google Scholar] [CrossRef]
  23. Charisiou, N.D.; Siakavelas, G.; Papageridis, K.N.; Baklavaridis, A.; Tzounis, L.; Avraam, D.G.; Goula, M.A. Syngas production via the biogas dry reforming reaction over nickel supported on modified with CeO2 and/or La2O3 alumina catalysts. J. Nat. Gas Sci. Eng. 2016, 31, 164–183. [Google Scholar] [CrossRef]
  24. Han, J.; Zhan, Y.; Street, J.; To, F.; Yu, F. Natural gas reforming of carbon dioxide for syngas over Ni–Ce–Al catalysts. Int. J. Hydrogen Energy 2017, 42, 18364–18374. [Google Scholar] [CrossRef]
  25. Boaro, M.; Colussi, S.; Trovarelli, A. Ceria-Based Materials in Hydrogenation and Reforming Reactions for CO2 Valorization. Front. Chem. 2019, 7, 28. [Google Scholar] [CrossRef] [PubMed]
  26. Nikoo, M.K.; Amin, N.A.S. Thermodynamic analysis of carbon dioxide reforming of methane in view of solid carbon formation. Fuel Process. Technol. 2011, 92, 678–691. [Google Scholar] [CrossRef] [Green Version]
  27. Siang, T.J.; Pham, T.L.M.; Van Cuong, N.; Phuong, P.T.T.; Phuc, N.H.H.; Truong, Q.D.; Vo, D.-V.N. Combined steam and CO2 reforming of methane for syngas production over carbon-resistant boron-promoted Ni/SBA-15 catalysts. Microporous Mesoporous Mater. 2018, 262, 122–132. [Google Scholar] [CrossRef]
  28. Cao, P.; Adegbite, S.; Wu, T. Thermodynamic Equilibrium Analysis of CO2 Reforming of Methane: Elimination of Carbon Deposition and Adjustment of H2/CO Ratio. Energy Procedia 2017, 105, 1864–1869. [Google Scholar] [CrossRef]
  29. Lim, Y.; Lee, C.-J.; Jeong, Y.S.; Song, I.H.; Lee, C.J.; Han, C. Optimal Design and Decision for Combined Steam Reforming Process with Dry Methane Reforming to Reuse CO2 as a Raw Material. Ind. Eng. Chem. Res. 2012. [Google Scholar] [CrossRef]
  30. Oyama, S.T.; Hacarlioglu, P.; Gu, Y.; Lee, D. Dry reforming of methane has no future for hydrogen production: Comparison with steam reforming at high pressure in standard and membrane reactors. Int. J. Hydrog. Energy 2012, 37, 10444–10450. [Google Scholar] [CrossRef]
  31. Tongamp, W.; Zhang, Q.; Saito, F. Preparation of meixnerite (Mg–Al–OH) type layered double hydroxide by a mechanochemical route. J. Mater. Sci. 2007, 42, 9210–9215. [Google Scholar] [CrossRef]
  32. Hassani Rad, S.J.; Haghighi, M.; Alizadeh Eslami, A.; Rahmani, F.; Rahemi, N. Sol–gel vs. impregnation preparation of MgO and CeO2 doped Ni/Al2O3 nanocatalysts used in dry reforming of methane: Effect of process conditions, synthesis method and support composition. Int. J. Hydrog. Energy 2016, 41, 5335–5350. [Google Scholar] [CrossRef]
  33. Wang, Z.; Jiang, X.; Pan, M.; Shi, Y. Nano-scale pore structure and its multi-fractal characteristics of tight sandstone by n2 adsorption/desorption analyses: A case study of shihezi formation from the sulige gas filed, ordos basin, china. Minerals 2020, 10, 377. [Google Scholar] [CrossRef]
  34. Akbari, E.; Alavi, S.M.; Rezaei, M. Synthesis gas production over highly active and stable nanostructured NiMgOAl2O3 catalysts in dry reforming of methane: Effects of Ni contents. Fuel 2017, 194, 171–179. [Google Scholar] [CrossRef]
  35. Cao, Y.; Li, H.; Zhang, J.; Shi, L.; Zhang, D. Promotional effects of rare earth elements (Sc, Y, Ce, and Pr) on NiMgAl catalysts for dry reforming of methane. RSC Adv. 2016, 6, 112215–112225. [Google Scholar] [CrossRef]
  36. Liu, Z.; Yao, S.; Johnston-Peck, A.; Xu, W.; Rodriguez, J.A.; Senanayake, S.D. Methanol steam reforming over Ni-CeO2 model and powder catalysts: Pathways to high stability and selectivity for H2/CO2 production. Catal. Today 2018, 311, 74–80. [Google Scholar] [CrossRef]
  37. Sangsong, S.; Ratana, T.; Tungkamani, S.; Sornchamni, T.; Phongaksorn, M. Effect of CeO2 loading of the Ce-Al mixed oxide on ultrahigh temperature water-gas shift performance over Ce-Al mixed oxide supported Ni catalysts. Fuel 2019, 252, 488–495. [Google Scholar] [CrossRef]
  38. Luisetto, I.; Tuti, S.; Battocchio, C.; Lo Mastro, S.; Sodo, A. Ni/CeO2–Al2O3 catalysts for the dry reforming of methane: The effect of CeAlO3 content and nickel crystallite size on catalytic activity and coke resistance. Appl. Catal. A Gen. 2015, 500, 12–22. [Google Scholar] [CrossRef]
  39. Lim, T.H.; Cho, S.J.; Yang, H.S.; Engelhard, M.; Kim, D.H. Effect of Co/Ni ratios in cobalt nickel mixed oxide catalysts on methane combustion. Appl. Catal. A Gen. 2015, 505, 62–69. [Google Scholar] [CrossRef] [Green Version]
  40. Lu, M.; Lv, P.; Yuan, Z.; Li, H. The study of bimetallic Ni-Co/cordierite catalyst for cracking of tar from biomass pyrolysis. Renew. Energy 2013, 60, 522–528. [Google Scholar] [CrossRef]
  41. Zhao, L.; Mu, X.; Liu, T.; Fang, K. Bimetallic Ni-Co catalysts supported on Mn-Al oxide for selective catalytic CO hydrogenation to higher alcohols. Catal. Sci. Technol. 2018, 8, 2066–2076. [Google Scholar] [CrossRef]
  42. You, X.; Wang, X.; Ma, Y.; Liu, J.; Liu, W.; Xu, X.; Peng, H.; Li, C.; Zhou, W.; Yuan, P.; et al. Ni-Co/Al2O3 Bimetallic Catalysts for CH4 Steam Reforming: Elucidating the Role of Co for Improving Coke Resistance. ChemCatChem 2014, 6, 3377–3386. [Google Scholar] [CrossRef]
  43. Li, X.; Ai, J.; Li, W.; Li, D. Ni-Co bimetallic catalyst for CH4 reforming with CO2. Front. Chem. Eng. China 2010, 4, 476–480. [Google Scholar] [CrossRef]
  44. Sikander, U.; Sufian, S.; Salam, M.A. A review of hydrotalcite based catalysts for hydrogen production systems. Int. J. Hydrog. Energy 2017, 42, 19851–19868. [Google Scholar] [CrossRef]
  45. Dębek, R.; Motak, M.; Galvez, M.E.; Grzybek, T.; Da Costa, P. Promotion effect of zirconia on Mg(Ni,Al)O mixed oxides derived from hydrotalcites in CO2 methane reforming. Appl. Catal. B Environ. 2018, 223, 36–46. [Google Scholar] [CrossRef]
  46. Challiwala, M.S.; Ghouri, M.M.; Linke, P.; El-Halwagi, M.M.; Elbashir, N.O. A combined thermo-kinetic analysis of various methane reforming technologies: Comparison with dry reforming. J. CO2 Util. 2017, 17, 99–111. [Google Scholar] [CrossRef]
  47. Özkara-Aydınoğlu, Ş. Thermodynamic equilibrium analysis of combined carbon dioxide reforming with steam reforming of methane to synthesis gas. Int. J. Hydrog. Energy 2010, 35, 12821–12828. [Google Scholar] [CrossRef]
  48. Carrasco, J.; López-Durán, D.; Liu, Z.; Duchoň, T.; Evans, J.; Senanayake, S.D.; Crumlin, E.J.; Matolín, V.; Rodríguez, J.A.; Ganduglia-Pirovano, M.V. In Situ and Theoretical Studies for the Dissociation of Water on an Active Ni/CeO2 Catalyst: Importance of Strong Metal-Support Interactions for the Cleavage of O-H Bonds. Angew. Chem. Int. Ed. 2015, 54, 3917–3921. [Google Scholar] [CrossRef]
  49. Sepehri, S.; Rezaei, M. Ce promoting effect on the activity and coke formation of Ni catalysts supported on mesoporous nanocrystalline γ-Al2O3 in autothermal reforming of methane. Int. J. Hydrog. Energy 2017, 42, 11130–11138. [Google Scholar] [CrossRef]
  50. Li, D.; Xu, S.; Song, K.; Chen, C.; Zhan, Y.; Jiang, L. Hydrotalcite-derived Co/Mg(Al)O as a stable and coke-resistant catalyst for low-temperature carbon dioxide reforming of methane. Appl. Catal. A Gen. 2018, 552, 21–29. [Google Scholar] [CrossRef]
  51. Liu, Z.; Lustemberg, P.; Gutiérrez, R.A.; Carey, J.J.; Palomino, R.M.; Vorokhta, M.; Grinter, D.C.; Ramírez, P.J.; Matolín, V.; Nolan, M.; et al. In Situ Investigation of Methane Dry Reforming on Metal/Ceria(111) Surfaces: Metal–Support Interactions and C−H Bond Activation at Low Temperature. Angew. Chem. Int. Ed. 2017, 56, 13041–13046. [Google Scholar] [CrossRef] [PubMed]
  52. Abdulrasheed, A.; Jalil, A.A.; Gambo, Y.; Ibrahim, M.; Hambali, H.U.; Shahul Hamid, M.Y. A review on catalyst development for dry reforming of methane to syngas: Recent advances. Renew. Sustain. Energy Rev. 2019, 108, 175–193. [Google Scholar] [CrossRef]
  53. Li, H.; Meng, F.; Gong, J.; Fan, Z.; Qin, R. Structural, morphological and optical properties of shuttle-like CeO2 synthesized by a facile hydrothermal method. J. Alloys Compd. 2017, 722, 489–498. [Google Scholar] [CrossRef]
  54. Jafarbegloo, M.; Tarlani, A.; Mesbah, A.W.; Muzart, J.; Sahebdelfar, S. NiO-MgO Solid Solution Prepared by Sol-Gel Method as Precursor for Ni/MgO Methane Dry Reforming Catalyst: Effect of Calcination Temperature on Catalytic Performance. Catal. Lett. 2016, 146, 238–248. [Google Scholar] [CrossRef]
  55. Chen, L.; Zhu, Q.; Hao, Z.; Zhang, T.; Xie, Z. Development of a Co-Ni bimetallic aerogel catalyst for hydrogen production via methane oxidative CO2 reforming in a magnetic assisted fluidized bed. Int. J. Hydrog. Energy 2010, 35, 8494–8502. [Google Scholar] [CrossRef]
  56. Dębek, R.; Motak, M.; Duraczyska, D.; Launay, F.; Galvez, M.E.; Grzybek, T.; Da Costa, P. Methane dry reforming over hydrotalcite-derived Ni–Mg–Al mixed oxides: The influence of Ni content on catalytic activity, selectivity and stability. Catal. Sci. Technol. 2016, 6, 6705–6715. [Google Scholar] [CrossRef]
  57. Ni, J.; Chen, L.; Lin, J.; Kawi, S. Carbon deposition on borated alumina supported nano-sized Ni catalysts for dry reforming of CH4. Nano Energy 2012, 1, 674–686. [Google Scholar] [CrossRef]
Figure 1. Powder XRD patterns of fresh 10Ni/MA (a), 5Ni5Ce/MA (b), and 5Ni5Co/MA (c) catalysts.
Figure 1. Powder XRD patterns of fresh 10Ni/MA (a), 5Ni5Ce/MA (b), and 5Ni5Co/MA (c) catalysts.
Catalysts 10 01450 g001
Figure 2. Nitrogen adsorption–desorption isotherms including the corresponding pore size distribution curves (inset) of 10Ni/MA, 5Ni5Ce/MA, and 5Ni5Co/MA catalysts.
Figure 2. Nitrogen adsorption–desorption isotherms including the corresponding pore size distribution curves (inset) of 10Ni/MA, 5Ni5Ce/MA, and 5Ni5Co/MA catalysts.
Catalysts 10 01450 g002
Figure 3. TEM images of 10Ni/MA, 5Ni5Ce/MA, and 5Ni5Co/MA catalysts.
Figure 3. TEM images of 10Ni/MA, 5Ni5Ce/MA, and 5Ni5Co/MA catalysts.
Catalysts 10 01450 g003
Figure 4. H2-TPR profile of 10Ni/MA (a), 5Ni5Ce/MA (b), and 5Ni5Co/MA (c) catalysts.
Figure 4. H2-TPR profile of 10Ni/MA (a), 5Ni5Ce/MA (b), and 5Ni5Co/MA (c) catalysts.
Catalysts 10 01450 g004
Figure 5. CO2-TPD profile of catalysts.
Figure 5. CO2-TPD profile of catalysts.
Catalysts 10 01450 g005
Figure 6. The CH4 (a) and CO2 conversions (b) for CSCRM and H2/CO ratios (c) of 10Ni/MA, 5Ni5Ce(T)/MA, and 5Ni5Ce/MA catalysts at 620 °C under ambient pressure (the relative error is about 3%).
Figure 6. The CH4 (a) and CO2 conversions (b) for CSCRM and H2/CO ratios (c) of 10Ni/MA, 5Ni5Ce(T)/MA, and 5Ni5Ce/MA catalysts at 620 °C under ambient pressure (the relative error is about 3%).
Catalysts 10 01450 g006
Figure 7. Thermogravimetric analysis (TGA) profile of spent catalysts.
Figure 7. Thermogravimetric analysis (TGA) profile of spent catalysts.
Catalysts 10 01450 g007
Table 1. The properties of 10Ni/MA, 5Ni5Ce/MA, and 5Ni5Co/MA catalysts.
Table 1. The properties of 10Ni/MA, 5Ni5Ce/MA, and 5Ni5Co/MA catalysts.
N2 Adsorption-Desorption aH2-TPR bDeconvolution of the CO2-TPD b (mmol/gcat)
CatalystSurface Area (m2/g)Total Pore Volume (cm3/g)H2 Uptake (mmol/gcat)WeakMediumStrongTotal Basicity
10Ni/MA970.2011.90.040.020.130.19
5Ni5Ce/MA1110.244.30.020.030.170.22
5Ni5Co/MA1080.2210.80.030.030.100.16
a The systematic error of N2 adsorption-desorption is ±5%. b The systematic error is ±1% for temperature and ±8% for the quantity of gaseous substances.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Dharmasaroja, N.; Ratana, T.; Tungkamani, S.; Sornchamni, T.; Simakov, D.S.A.; Phongaksorn, M. The Effects of CeO2 and Co Doping on the Properties and the Performance of the Ni/Al2O3-MgO Catalyst for the Combined Steam and CO2 Reforming of Methane Using Ultra-Low Steam to Carbon Ratio. Catalysts 2020, 10, 1450. https://doi.org/10.3390/catal10121450

AMA Style

Dharmasaroja N, Ratana T, Tungkamani S, Sornchamni T, Simakov DSA, Phongaksorn M. The Effects of CeO2 and Co Doping on the Properties and the Performance of the Ni/Al2O3-MgO Catalyst for the Combined Steam and CO2 Reforming of Methane Using Ultra-Low Steam to Carbon Ratio. Catalysts. 2020; 10(12):1450. https://doi.org/10.3390/catal10121450

Chicago/Turabian Style

Dharmasaroja, Nichthima, Tanakorn Ratana, Sabaithip Tungkamani, Thana Sornchamni, David S. A. Simakov, and Monrudee Phongaksorn. 2020. "The Effects of CeO2 and Co Doping on the Properties and the Performance of the Ni/Al2O3-MgO Catalyst for the Combined Steam and CO2 Reforming of Methane Using Ultra-Low Steam to Carbon Ratio" Catalysts 10, no. 12: 1450. https://doi.org/10.3390/catal10121450

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