Next Article in Journal
Semiconductor Nanomaterial Photocatalysts for Water-Splitting Hydrogen Production: The Holy Grail of Converting Solar Energy to Fuel
Previous Article in Journal
Highly Dispersed Ni on Nitrogen-Doped Carbon for Stable and Selective Hydrogen Generation from Gaseous Formic Acid
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Stability and Activity of Rhodium Promoted Nickel-Based Catalysts in Dry Reforming of Methane

by
Jehad Saleh
1,
Ahmed Sadeq Al-Fatesh
1,*,
Ahmed Aidid Ibrahim
1,*,
Francesco Frusteri
2,
Ahmed Elhag Abasaeed
1,
Anis Hamza Fakeeha
1,
Fahad Albaqi
3,
Khalid Anojaidi
3,
Salwa B. Alreshaidan
4,
Ibrahim Albinali
3,
Abdulrahman A. Al-Rabiah
1 and
Abdulaziz Bagabas
3,*
1
Chemical Engineering Department, College of Engineering, King Saud University, P.O. Box 800, Riyadh 11421, Saudi Arabia
2
CNR-ITAE, Istituto di Tecnologie Avanzate per Energia “Nicola Giordano”, Via S. Lucia Sopra Contesse 5, 98126 Messina, Italy
3
President Office, King Abdulaziz City for Science and Technology (KACST), P.O. Box 6086, Riyadh 11442, Saudi Arabia
4
Department of Chemistry, Faculty of Science, King Saud University, P.O. Box 800, Riyadh 11451, Saudi Arabia
*
Authors to whom correspondence should be addressed.
Nanomaterials 2023, 13(3), 547; https://doi.org/10.3390/nano13030547
Submission received: 15 January 2023 / Revised: 26 January 2023 / Accepted: 27 January 2023 / Published: 29 January 2023

Abstract

:
The rhodium oxide (Rh2O3) doping effect on the activity and stability of nickel catalysts supported over yttria-stabilized zirconia was examined in dry reforming of methane (DRM) by using a tubular reactor, operated at 800 °C. The catalysts were characterized by using several techniques including nitrogen physisorption, X-ray diffraction, transmission electron microscopy, H2-temperature programmed reduction, CO2-temperature programmed Desorption, and temperature gravimetric analysis (TGA). The morphology of Ni-YZr was not affected by the addition of Rh2O3. However, it facilitated the activation of the catalysts and reduced the catalyst’s surface basicity. The addition of 4.0 wt.% Rh2O3 gave the optimum conversions of CH4 and CO2 of ~89% and ~92%, respectively. Furthermore, the incorporation of Rh2O3, in the range of 0.0–4.0 wt.% loading, enhanced DRM and decreased the impact of reverse water gas shift, as inferred by the thermodynamics analysis. TGA revealed that the addition of Rh2O3 diminished the carbon formation on the spent catalysts, and hence, boosted the stability, owing to the potential of rhodium for carbon oxidation through gasification reactions. The 4.0 wt.% Rh2O3 loading gave a 12.5% weight loss of carbon. The TEM images displayed filamentous carbon, confirming the TGA results.

Graphical Abstract

1. Introduction

It is anticipated that energy demand in 2030 will be 55% more than that in the last three decades [1]. However, it is essential to have clean energy in the light of the tighten environmental legislations. In this content, the abundant natural gas reserves could provide a solution because of its cleaner combustion and easier reforming than the crude oil [2]. Different reforming techniques are employed for CH4 valorisation. For instance, steam reforming of methane (SRM) [3], dry reforming of methane (DRM) [4], and partial oxidation of methane (POM) [5]. The reforming of CH4, the major constituent of natural gas, can be performed to produce synthesis gas (syngas, CO and H2 mixture) via one of the following three processes:
CH 4 + CO 2 2 H 2 + 2 CO Δ H 298 ° = + 247 kJ mol   dry   reforming
CH 4 + H 2 O 3 H 2 + CO Δ H 298 ° = + 206 kJ mol   steam   reforming
CH 4 + 1 2 O 2 2 H 2 + CO Δ H 298 ° = 36 kJ mol   partial   oxidation
These processes have presently revived research interests [6,7,8].
Indeed, dry reforming of methane (DRM), which is an environmentally friendly and energetically striking method utilizes the greenhouse gases CH4 and CO2 to generate syngas with an equal H2/CO mole ratio, which is appropriate for Fischer–Tropsch syngas to produce liquid fuel [9]. Compared with other methane reforming processes, the overall operation cost is estimated to be 20% lower to produce high purity syngas with low CO2 content [10]. However, the DRM process has serious drawbacks; the core ones are related to energy costs due to high endothermicity and catalyst deactivation (carbon deposition and sintering), limiting its commercial application [11]. The carbon formation can be resolved from the catalyst by construction modification. To avoid homogeneous cracking and coking of hydrocarbons, highly active metals should be chosen [12,13]. Noble metals such as platinum (Pt), rhodium (Rh), and palladium (Pd) have been widely investigated due to their superior performance, but their high prices restrain their application [14,15]. Cheap transition metals are the potential switch. For instance, nickel (Ni), as an active site, has been studied and shown good performance in DRM with a high syngas selectivity [16]. A catalyst with 5 wt.% nickel oxide loading was employed in long-term stability testing and showed stable catalytic performance up to 50 h time-on-stream without any decrease in methane conversion in the dry reforming process [17]. However, the eminent propensity of Ni to sintering and coking under DRM conditions is the major drawback in the development of stable Ni-based catalysts [18]. The design of cost-effective, efficient DRM catalysts is a grand challenge in this topic. Bimetallic catalysts, providing synergistic effects via metal-to-metal interactions, appear to be an effective strategy for achieving these goals. The addition of a noble metal to Ni-based catalysts characteristically results in a higher dispersion of metal particles, owing to the dilution effect of the noble metal, smaller alloyed particles, enhanced reducibility, and better resistance to oxidation, sintering, and carbon formation, favouring DRM performance and time-on-stream stability [19,20]. For the selection of the active metal, researchers have investigated Rh, Ru, Pt, and Ni. Rh showed the highest activity and best anti-coking performance [21]. The Ni-Rh bimetallic system may be satisfactory for engineering applications when considering the performance and cost [22]. Neuberg et al. highlighted the promoting consequence of Rh on the activity and stability of Pt-based methane combustion catalyst (Pt-Rh/γ-Al2O3) [23]. The Rh-promoted catalysts showed better stability, as a result of developing smaller crystal sizes. Alternatively, Arbag et al. reported better performance and stability of Ni/MCM-41ctalaytst in DRM due to the reduction of reverse water gas shift reaction [24]. Khalighi et al. examined the catalytic performance of Rh- and Ni-containing catalysts, supported by CoAl2O4 in DRM. They found that the 3 wt.% Rh-containing catalyst gave the maximum methane conversion with no deactivation or carbon formation [25]. Tarifa et al. studied steam reforming of clean biogas over Rh or Ru catalysts supported on NiCrAl open-cell metallic foam [26]. The Rh catalysts showed great activity, stability, and low carbon deposition due to the presence of large Ni particles. Ocsachoque et al. investigated the activity, stability, and carbon deposition of a Rh-promoted Ni-based catalyst (Ni/Al2O3) during DRM [1]. Their outcome displayed that the Rh addition favoured metal-support interaction and provided a high activity. Furthermore, the elimination of carbon deposition can be enhanced by selecting a support with good oxygen storage/release abilities and less acidic sites [27]. ZrO2 has good thermal stability and three crystal structures (monoclinic, tetragonal, and cubic) [28]. The various ZrO2 forms also have different porous structures, which might influence the properties of the generated coke. ZrO2, as a support, exhibited better stability during calcination and reforming due to a high Hüttig and Tamman temperature [11]. Zhang et al. [29] used ZrO2 for steam reforming and found that the modified catalysts, like Y2Zr2O7, had a pyrochlore structure with more mobile oxygen species to enhance coking resistance. Support optimization is a good way to prevent carbon deposition. Supports with higher surface basicity showed higher DRM efficiency. Modification of catalyst support helped to improve the catalyst stability with less carbon deposition [30,31]. Pedrero et al. investigated the partial oxidation of methane and ethane over Rh and Ni supported on zirconia-modified alumina catalysts. Their results stipulated that the modified support improved significantly the performances of the catalysts [32]. Similarly, Lv et al. reported the dry reforming of low-carbon alkane over Ni/La2O3-ZrO2. Their results showed high performance and stability, which were due to the modifications of support and the formation of a stable pyrochlore La2Zr2O7 [33]. When Y2O3 is doped with transition metals such as Zr, the number of oxygen vacancies and thermal stability are further improved. Impregnation is a common catalyst preparation method, which is predominant in the production of industrial catalysts [34]. The addition of yttria into zirconia supports prevents phase transition in zirconia at high reaction temperatures [35]. Moreover, it compensates for the vacancy in zirconia, offers oxygen ion conductivity in the lattice, obstructs the reaction of carbon deposit with ZrO2 support, confines the crystalline size of NiO, and improves the surface parameter as well as develops strong basic sites [35,36,37]. Therefore, it is convincing and essential to develop an innovative catalyst with satisfactory physicochemical properties to employ in the DRM process. On this basis, we were motivated to combine Rh and Ni, and supported them on yttria-stabilized zirconia catalysts for DRM. The effect of different loadings of Rh2O3 on the activity and stability of 5 wt.% NiO at 800 °C, were tested for determining the optimum loading. The current work aims to get a further comprehensive vision of the role of Rh2O3 on the DRM reaction and establish the coke-resistant effect of this promoter. The effect of Rh2O3 loading was explored by suitable characterization techniques.

2. Materials and Methods

2.1. Material

The chemicals used were obtained commercially and were used without further purification. Nickel nitrate hexahydrate [Ni (NO3)2.6H2O] was obtained from Riedel-De Haen AG, Seelze, Germany, rhodium chloride (RhCl3) was purchased from Sigma-Aldrich (Ward Hill, MA, USA), and the 8.0 wt.% yttria-stabilized zirconia was purchased from Alfa Aesar (St. louis, Mo, USA).

2.2. Catalyst Preparation

The wet impregnation method was employed to prepare the rhodium-promoted, yttria-stabilized zirconia-supported nickel catalysts. The appropriate amounts of nickel nitrate hexahydrate to obtain 5.0 wt.% NiO, rhodium chloride to obtain 1.0, 2.0, 3.0, 4.0, or 5.0 wt.% Rh2O3, and the support were mixed and ground together. Drops of ultrapure water were added to the solid mixture to make a paste. Subsequently, this paste was stirred mechanically until it was dried. Wetting and drying were performed three times for ensuring the homogeneous distribution of the components within the support matrix. Calcination at 600 °C for three hours was then performed to obtain the rhodium-promoted, yttria-stabilized zirconia-supported nickel catalysts, which were abbreviated as Rh-x (x = 0, 1, 2, 3, 4, 5).

2.3. Catalyst Characterization

The catalysts were characterized by N2 adsorption–desorption at −196 °C by using a Micromeritics Tristar II 3020 for porosity and surface area analyses. Hydrogen temperature-programmed reduction was performed using 0.070 g of catalyst, placed inside the holder of a Micromeritics AutoChem II apparatus. The readings were taken at 150 °C under argon gas flow for half an hour, followed by cooling to ambient temperature. The next step involved heating by the furnace up to 800 °C, ramping at 10 °C min−1 in an atmosphere of a H2/Ar mixture (1:9 vol. %), flowing at 40 mL/min. The thermal conductivity unit recorded the H2 consumption. The X-ray diffraction patterns of the catalysts were recorded on a Miniflex Rigaku diffractometer, equipped with Cu Kα X-ray radiation. The device was run at 40 kV and 40 mA. The basicity was determined by using the CO2 temperature-programmed desorption, where a sample of 0.1 g was washed in helium flow and was then saturated for 1.0 h at 200 °C in an atmosphere of 20 vol.% CO2/He. A carrier flow of helium at 25 mL/min was used in the temperature range of 25–720 °C (heating rate, 12 °C/min) for CO2 desorption. The morphology of the catalyst samples was examined via a high-resolution transmission electron microscope (HRTEM model: JEM-2100 F, JEOL, Akishima, Tokyo, Japan). The quantity of carbon deposit on the spent catalysts was determined by the thermo-gravimetric analysis. A platinum pan was filled with 10–15 mg of the spent catalyst. Heating was performed from room temperature to 900 °C at a rate of 20 °C min−1. The change in mass was constantly monitored as the heating progressed.

2.4. Catalyst Activity Test

The dry reforming of methane catalytic activity test started with a 0.1 g catalyst, put in stainless steel vertical fixed-bed tubular reactor (0.91 cm i.d. and 0.30 m long) (PID Eng. and Tech Micro Activity) by using a ball of glass wool. The reaction was carried out under 1.0 atm pressure. A K-type stainless sheathed thermocouple was used to maintain the temperature of the reaction. The catalyst was activated with reductive treatment under the flow of hydrogen (20 mL/min) for 60 min at 600 °C. Then, after eliminating the physiosorbed H2, the N2 treatment was performed for 15 min. The mixture of feed gas was CH4/CO2/N2 at 30, 30, 10 mL/min with a space velocity of 42,000 mL/(h.gcat), passed through the reactor. TCD-equipped gas chromatography (GC-2014 SHIMADZU, Kyoto, Japan) was used to analyse the product. The following expressions were used for the conversions:
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. Results

The surface properties of the fresh catalysts were examined using nitrogen adsorption–desorption isotherms. Figure 1 shows the results of the isotherms. According to the IUPAC sorting system, the observed isotherms belonged to type IV classification with [38,39] an H3-type hysteresis loop, implying the existence of slit-like pores in the catalysts. In the range of 0.8–1.0, the relative pressure (P/P0) grew. In addition, the catalysts showed low specific volume adsorption of nitrogen gas, in the range of 4.5–8.0 cm3/g. Table 1 outlines the textural aspects of the samples. The catalysts showed specific surface areas in the range of 27–31 m2/g. Nevertheless, the pore diameter was in the range of 22–25 nm.
Figure 2 exhibits the XRD patterns of the catalysts at 2θ° = 5–80°. The XRD pattern of yttria-stabilized zirconia support showed the cubic phase of zirconia (JCPDS No. 49-1642), which appeared at 2θ = 30°, 35°, 50°, 60°, 63°, and 75° in reference to the (111), (200), (220), (311), (222), and (400), respectively, in all catalysts. Meanwhile, the peaks for NiO were at 37.2° and 43.1°. The XRD patterns for the Rh2O3-promoted samples were identical to those of the unpromoted one, denoting that the introduction of Rh2O3 did not bring additional compounds. The absence of Rh2O3 in the XRD patterns indicated its fine dispersion on the surface.
To perform H2-temperature programming reduction (H2-TPR) analysis, a sample of ca. 100 mg of catalyst was placed in a r micro-reactor (i.d., 4 mm), fed with a 5 vol. % H2/Ar at a flow rate of 3.6 L/h at STP, in the range of 0–700 °C, with a heating rate of 12 °C/min. A thermal conductivity detector (TCD) monitored the hydrogen consumption. Figure 3 displays the TPR profiles of the fresh Rh-x (x = 0, 1, 2, 3, 4, and 5 wt.%) catalysts. The reduction peaks for the Rh-promoted Ni catalysts appeared in the low temperature range (80–250 °C). These were attributed to Ni catalysts loosely attached to the support, while the broad peaks with overlapped shoulders in the medium temperature range (300–500 °C) was for the unpromoted Ni catalysts, whereby the Ni was moderately attached to the support. The incorporation of the Rh-promoter caused the reduction peaks for nickel oxide to shift to lower temperatures in the range of 250–350 °C; the higher the loading of Rh was, the lower the reduction temperature was, viz. the weaker the interaction between the nickel oxide and the support was. Moreover, all the catalysts did not show any appreciable peak beyond 550 °C, stipulating that all the existing oxides were reduced and the absence of spinel phases like NiRh2O3. The incorporation of Rh2O3 (1–3 wt.%) slightly boosted the rate of H2 consumption, as shown in Table 2. Further increase in Rh2O3 loading (3–5 wt.%) did not affect the amount of hydrogen consumption, indicating that Rh2O3 was impeded under the surface of the catalysts.
The CO2-TPD characterization was conducted to justify the basicity effect that is observed by adding Rh2O3. Because CO2 adsorption capacity plays a fundamental role in promoting CO2 conversion in DMR, the effect of Rh2O3 on CO2 adsorption capacity was the first parameter to be considered [40]. Figure 4 demonstrates that all the catalysts did not effectively promote CO2 adsorption at temperatures above 400 °C. Thus, the low desorption temperatures of CO2 in the range of 50–400 °C reflected a weak basicity due to the availability of few basic sites on the surface of the catalysts [41,42]. The peak at 110 °C was ascribed to the adsorption of CO2 on OH groups which is weak, while peaks at 250 °C and 340 °C were attributed to moderate adsorption of CO2 on metal–oxygen pairs [43]. The increase in Rh2O3 amount produced peaks with lower intensities, and hence, further diminished the basicity of the catalysts. The obtained results underscored that the basic properties of the catalysts could be altered by doping the Rh2O3 promoter. Table 3 displays the amount of CO2 consumption in mol/gcat during the CO2-TPD process. The Rh-promoted samples assumed values between 0.8 and 1.0.
Figure 5 displays the methane conversion profiles. The activity of the Ni-Rh-x (x = 0, 1, 2, 3, 4, 5 wt.%) catalysts was determined at 700 °C. The acquired activity clearly indicated that the addition of a Rh2O3 promoter had a profound impact on the performance. The various Rh2O3 loadings of the promoter increased the conversion values from 7 to 20% in comparison to the unpromoted catalyst. Indeed, the 4 wt.% Rh2O3 loadings gave the optimum activity value of 88% and the best stability since the activity drop was less than 1% during the 420 min of reaction. The improvement of conversion with Rh2O3 loadings up to 4 wt.% was credited to additional activity sites offered by the Rh metal. However, the higher loading of Rh2O3 such as 5 wt.% reduced the activity due to Rh covering the Ni active sites. Figure 6 shows the carbon dioxide conversion profiles. It also displayed a similar improvement in conversion with respect to the promotional effects of Rh2O3. The conversion values were higher than the corresponding CH4 conversion in Figure 5. This incident can be related to the occurrence of the reverse water gas shift reaction Equation (6), where the CO2 reacts with formed H2.
CO 2 + H 2   H 2 O + CO
Moreover, it can be also attributed to the gasification of carbon deposition by carbon dioxide over metallic rhodium [41]. The catalytic efficiency of this research was compared to the results cited in the past in Table 4. The results show the relevance and the excellence of the adopted process for the formation of a competitive catalyst.
Thermodynamic Analysis
The Predictive Soave–Redlich–Kwong (PSRK) equation of state was used, as a thermodynamic model, to determine the changes in enthalpy (ΔH), entropy (ΔS), and Gibbs free energy (ΔG) at 800 °C for the main reaction of dry reforming of methane (DRM) and the side reaction of reverse water gas shift (RWGS) over each catalyst. The equilibrium constants (Keq) for these two reactions were calculated at 800 °C by using the “REquil” model. All calculations were performed using Aspen Plus, V13. Table 5 shows the thermodynamic parameters and the equilibrium constants of the two reactions.
The equilibrium conversion of methane was found to be 73.6% at 800 °C. Because the produced hydrogen from DRM was simultaneously consumed by the RWGS, the equilibrium was shifted to higher values, as stated by Le Chatelier’s principle.
As shown in Table 5, the values of ΔH, ΔS, and ΔG increased with increasing Rh2O3 content (the promoter) from 0.0 to 4.0 wt.%. Such observations indicated that less carbon dioxide and hydrogen were involved in RWGS with increased loading of the promoter due to the much higher endothermic nature of DRM (ΔH°298 = 247,021 kJ/kmol) than that of RWGS (ΔH°298 = 41,058 kJ/kmol) and the much higher Keq value of DRM than that of RWGS. However, increasing the promoter content to 5.0 wt.% led to a reduction in the values of these thermodynamic parameters, probably owing to the coverage of some nickel metal active sites by some rhodium metal particles, and hence, the loss of some of the synergetic effect between the nickel and rhodium metals. Therefore, this thermodynamics analysis provided a support for the highest observed catalytic performance for the Rh-4 catalyst.
Figure 7 displays the TEM images of the Ni-Rh-x (x = 0, 4 wt.%) catalysts at 50 or 200 nm scales. Figure 7A,B show fresh Rh-0 and Rh-4, respectively, where the agglomerated catalysts particles showed the same morphology irrespective of Rh2O3 promoter loading. Figure 7C,D show spent Rh-0 and Rh-4, respectively. The spent catalyst particles showed the formation of filamentous carbon.
For the quantitative determination of carbon formations after the reaction, TGAs of the Ni-Rh-x (x = 0, 1, 2, 3, 4, 5 wt.%) catalysts were performed under an air atmosphere (Figure 8). The un-promoted catalyst Rh-0 gave the highest amount of weight loss (18.5%). The gradual addition of Rh2O3 clearly influenced the coke formation, where 7–15 wt.% of coke was deposited on catalysts containing 1–5 wt.% Rh2O3. The optimally promoted catalyst (Rh-4) provided the least weight loss, with the exclusion of Rh-5, which was marked by a reaction drop. The observed trend in the reduction of the amount of carbon deposition with increasing Rh2O3 loading could be ascribed to the ability of Rh to catalyse carbon oxidation via carbon gasification by carbon dioxide and steam [41], as shown in the following chemical equations:
CO 2 + C   2 CO
H 2 O + C   CO + H 2
On this basis, the carbon formation over our catalysts in the process of DRM might be attributed to the decomposition of methane [52]:
CH 4 2 H 2 + C

4. Conclusions

In this article, the impact of a rhodium oxide promoter on a nickel catalyst supported over yttria-stabilized zirconia for CH4 reforming using a CO2 oxidant was investigated at 800 °C. The dry impregnation method was used for the synthesis of all catalysts. The characterization results substantiated that the addition of Rh2O3 did not markedly affect the morphology. However, it lowered the basicity, reduced the coke formation, and brought down the reduction temperature. These combined qualities demonstrated that 4 wt.% loading of Rh2O3 gave the optimum catalyst performance in activity and stability during the dry reforming of methane with minimization of reverse water gas shift, as inferred from the thermodynamics analysis.

Author Contributions

Conceptualization and funding acquisition, J.S.; formal analysis and writing—original draft preparation, A.S.A.-F., A.A.I. and A.B.; investigation and resources, F.F. and S.B.A.; methodology and writing—review and editing, F.A., I.A. and K.A.; supervision and project administration, A.E.A. and A.H.F.; investigation and resources, A.A.A.-R. All authors have read and agreed to the published version of the manuscript.

Funding

The funding is from Researchers Supporting Project (number RSP2023R368), King Saud University, Riyadh, Saudi Arabia.

Acknowledgments

The authors would like to extend their sincere appreciation to the Researchers Supporting Project (number RSP2023R368), King Saud University, Riyadh, Saudi Arabia.

Conflicts of Interest

The authors declare that they have no known competing financial interests or personal relationships.

References

  1. Ocsachoque, M.; Quincoces, C.E.; González, M.G. Effect of Rh Addition on Activity and Stability over NI/γ-AL2O3 Catalysts during Methane Reforming with CO2. Stud. Surf. Sci. Catal. 2007, 167, 397–402. [Google Scholar] [CrossRef]
  2. Liu, Q.; Wu, X.; Wang, X.; Jin, Z.; Zhu, D.; Meng, Q.; Huang, S.; Liu, J.; Fu, Q. Carbon and Hydrogen Isotopes of Methane, Ethane, and Propane: A Review of Genetic Identification of Natural Gas. Earth Sci. Rev. 2019, 190, 247–272. [Google Scholar] [CrossRef]
  3. Zhou, Y.; Haynes, D.; Baltrus, J.; Roy, A.; Shekhawat, D.; Spivey, J.J. Methane Steam Reforming at Low Steam-to-Carbon Ratio: The Effect of Y Doping in Rh Substituted Lanthanum Zirconates. Appl. Catal. A Gen. 2020, 606, 117802–117814. [Google Scholar] [CrossRef]
  4. Mao, Y.; Zhang, L.; Zheng, X.; Liu, W.; Cao, Z.; Peng, H. Coke-Resistance over Rh–Ni Bimetallic Catalyst for Low Temperature Dry Reforming of Methane. Int. J. Hydrogen Energy 2023, in press. [Google Scholar] [CrossRef]
  5. Gao, Y.; Wei, Y.; Sun, W.; Zhao, G.; Liu, Y.; Lu, Y. Insight into Deactivation of the Carbon-/Sintering-Resistant Ni@Silicalite-1 for Catalytic Partial Oxidation of Methane to Syngas. Fuel 2022, 320, 123892. [Google Scholar] [CrossRef]
  6. Abdulrasheed, A.; Jalil, A.A.; Gambo, Y.; Ibrahim, M.; Hambali, H.U.; Hamid, M.Y.S. 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]
  7. Aramouni, N.A.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, 82, 2570–2585. [Google Scholar] [CrossRef]
  8. Akri, M.; Zhao, S.; 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]
  9. Shah, M.; Bordoloi, A.; Nayak, A.K.; Mondal, P. Effect of Ti/Al Ratio on the Performance of Ni/TiO2-Al2O3 Catalyst for Methane Reforming with CO2. Fuel Process. Technol. 2019, 192, 21–35. [Google Scholar] [CrossRef]
  10. Ross, J.R.H. Natural Gas Reforming and CO2 Mitigation. Catal. Today 2005, 100, 151–158. [Google Scholar] [CrossRef]
  11. Usman, M.; Wan Daud, W.M.A.; Abbas, H.F. Dry Reforming of Methane: Influence of Process Parameters—A Review. Renew. Sustain. Energy Rev. 2015, 45, 710–744. [Google Scholar] [CrossRef] [Green Version]
  12. Vita, A.; Italiano, C.; Fabiano, C.; Pino, L.; Laganà, M.; Recupero, V. Hydrogen-Rich Gas Production by Steam Reforming of n-Dodecane: Part I: Catalytic Activity of Pt/CeO2 Catalysts in Optimized Bed Configuration. Appl. Catal. B Environ. 2016, 199, 350–360. [Google Scholar] [CrossRef]
  13. Shoynkhorova, T.B.; Rogozhnikov, V.N.; Simonov, P.A.; Snytnikov, P.V.; Salanov, A.N.; Kulikov, A.V.; Gerasimov, E.Y.; Belyaev, V.D.; Potemkin, D.I.; Sobyanin, V.A. Highly Dispersed Rh/Ce0.75Zr0.25O2-δ-ƞ-Al2O3/FeCrAl Wire Mesh Catalyst for Autothermal n-Hexadecane Reforming. Mater. Lett. 2018, 214, 290–292. [Google Scholar] [CrossRef]
  14. Amjad, U.E.S.; Quintero, C.W.M.; Ercolino, G.; Italiano, C.; Vita, A.; Specchia, S. Methane Steam Reforming on the Pt/CeO2 Catalyst: Effect of Daily Start-Up and Shut-Down on Long-Term Stability of the Catalyst. Ind. Eng. Chem. Res. 2019, 58, 16395–16406. [Google Scholar] [CrossRef]
  15. Chung, W.C.; Chang, M.B. Review of Catalysis and Plasma Performance on Dry Reforming of CH4 and Possible Synergistic Effects. Renew. Sustain. Energy Rev. 2016, 62, 13–31. [Google Scholar] [CrossRef]
  16. Myint, M.N.Z.; Yan, B.; Wan, J.; Zhao, S.; Chen, J.G. Reforming and Oxidative Dehydrogenation of Ethane with CO2 as a Soft Oxidant over Bimetallic Catalysts. J. Catal. 2016, 343, 168–177. [Google Scholar] [CrossRef] [Green Version]
  17. Meshkani, F.; Rezaei, M.; Andache, M. Investigation of the Catalytic Performance of Ni/MgO Catalysts in Partial Oxidation, Dry Reforming and Combined Reforming of Methane. J. Ind. Eng. Chem. 2014, 20, 1251–1260. [Google Scholar] [CrossRef]
  18. Tang, Y.; He, C.; Imler, G.H.; Parrish, D.A.; Shreeve, J.M. Aminonitro Groups Surrounding a Fused Pyrazolotriazine Ring: A Superior Thermally Stable and Insensitive Energetic Material. ACS Appl. Energy Mater. 2019, 2, 2263–2267. [Google Scholar] [CrossRef]
  19. Shoynkhorova, T.B.; Simonov, P.A.; Potemkin, D.I.; Snytnikov, P.V.; Belyaev, V.D.; Ishchenko, A.V.; Svintsitskiy, D.A.; Sobyanin, V.A. Highly Dispersed Rh-, Pt-, Ru/Ce0.75Zr0.25O2–δ Catalysts Prepared by Sorption-Hydrolytic Deposition for Diesel Fuel Reforming to Syngas. Appl. Catal. B Environ. 2018, 237, 237–244. [Google Scholar] [CrossRef]
  20. Lee, S.; Bae, M.; Bae, J.; Katikaneni, S.P. Ni–Me/Ce0.9Gd0.1O2−x (Me: Rh, Pt and Ru) Catalysts for Diesel Pre-Reforming. Int. J. Hydrogen Energy 2015, 40, 3207–3216. [Google Scholar] [CrossRef]
  21. Xiao, Z.; Zhang, X.; Hou, F.; Wu, C.; Wang, L.; Li, G. Tuning Metal-Support Interaction and Oxygen Vacancies of Ceria Supported Nickel Catalysts by Tb Doping for n-Dodecane Steam Reforming. Appl. Surf. Sci. 2020, 503, 144319. [Google Scholar] [CrossRef]
  22. Kim, D.H.; Kang, J.S.; Lee, Y.J.; Park, N.K.; Kim, Y.C.; Hong, S.I.; Moon, D.J. Steam Reforming of N-Hexadecane over Noble Metal-Modified Ni-Based Catalysts. Catal. Today 2008, 136, 228–234. [Google Scholar] [CrossRef]
  23. Neuberg, S.; Pennemann, H.; Shanmugam, V.; Zapf, R.; Kolb, G. Promoting Effect of Rh on the Activity and Stability of Pt-Based Methane Combustion Catalyst in Microreactors. Catal. Commun. 2021, 149, 106202. [Google Scholar] [CrossRef]
  24. Arbag, H.; Yasyerli, S.; Yasyerli, N.; Dogu, G. Activity and Stability Enhancement of Ni-MCM-41 Catalysts by Rh Incorporation for Hydrogen from Dry Reforming of Methane. Int. J. Hydrogen Energy 2010, 35, 2296–2304. [Google Scholar] [CrossRef]
  25. Khalighi, R.; Bahadoran, F.; Panjeshahi, M.H.; Zamaniyan, A.; Tahouni, N. High Catalytic Activity and Stability of X/CoAl2O4 (X = Ni, Co, Rh, Ru) Catalysts with No Observable Coke Formation Applied in the Autothermal Dry Reforming of Methane Lined on Cordierite Monolith Reactors. Microporous Mesoporous Mater. 2020, 305, 110371. [Google Scholar] [CrossRef]
  26. Tarifa, P.; Schiaroli, N.; Ho, P.H.; Cañaza, F.; Ospitali, F.; de Luna, G.S.; Lucarelli, C.; Fornasari, G.; Vaccari, A.; Monzon, A.; et al. Steam Reforming of Clean Biogas over Rh and Ru Open-Cell Metallic Foam Structured Catalysts. Catal. Today 2022, 383, 74–83. [Google Scholar] [CrossRef]
  27. Xiao, Z.; Zhang, X.; Wang, L.; Li, G. Optimizing the Preparation of Ni-Ce-Pr Catalysts for Efficient Hydrogen Production by n-Dodecane Steam Reforming. Int. J. Energy Res. 2020, 44, 1828–1842. [Google Scholar] [CrossRef]
  28. Stawarczyk, B.; Keul, C.; Eichberger, M.; Figge, D.; Edelhoff, D.; Lümkemann, N. Three Generations of Zirconia: From Veneered to Monolithic. Part I. Quintessence Int. 2017, 48, 369–380. [Google Scholar]
  29. Zhang, X.; Peng, L.; Fang, X.; Cheng, Q.; Liu, W.; Peng, H.; Gao, Z.; Zhou, W.; Wang, X. Ni/Y2B2O7 (B[Dbnd]Ti, Sn, Zr and Ce) Catalysts for Methane Steam Reforming: On the Effects of B Site Replacement. Int. J. Hydrogen Energy 2018, 43, 8298–8312. [Google Scholar] [CrossRef]
  30. Liu, W.; Li, L.; Zhang, X.; Wang, Z.; Wang, X.; Peng, H. Design of Ni-ZrO2@SiO2 Catalyst with Ultra-High Sintering and Coking Resistance for Dry Reforming of Methane to Prepare Syngas. J. CO2 Util. 2018, 27, 297–307. [Google Scholar] [CrossRef]
  31. Margossian, T.; Larmier, K.; Kim, S.M.; Krumeich, F.; Fedorov, A.; Chen, P.; Müller, C.R.; Copéret, C. Molecularly Tailored Nickel Precursor and Support Yield a Stable Methane Dry Reforming Catalyst with Superior Metal Utilization. J. Am. Chem. Soc. 2017, 139, 6919–6927. [Google Scholar] [CrossRef]
  32. Pedrero, C.M.; Carrazán, S.G.; Ruiz, P. Preliminary Results on the Role of the Deposition of Small Amounts of ZrO2 on Al2O3 Support on the Partial Oxidation of Methane and Ethane over Rh and Ni Supported Catalysts. Catal. Today 2021, 363, 111–121. [Google Scholar] [CrossRef]
  33. Lv, J.; Wang, D.; Wei, B.; Jin, L.; Li, Y.; Hu, H. Integrated Process of Coal Pyrolysis with Dry Reforming of Low Carbon Alkane over Ni/La2O3-ZrO2 with Different La/Zr Ratio. Fuel 2021, 292, 120412. [Google Scholar] [CrossRef]
  34. Achouri, I.E.; Abatzoglou, N.; Fauteux-Lefebvre, C.; Braidy, N. Diesel Steam Reforming: Comparison of Two Nickel Aluminate Catalysts Prepared by Wet-Impregnation and Co-Precipitation. Catal. Today 2013, 207, 13–20. [Google Scholar] [CrossRef]
  35. INGEL, R.P.; III, D.L. Lattice Parameters and Density for Y2O3-Stabilized ZrO2. J. Am. Ceram. Soc. 1986, 69, 325–332. [Google Scholar] [CrossRef]
  36. Darby, R.J.; Kumar, R.V. Activation Enthalpies for Oxygen Ion Motion in Cubic Yttria-Stabilized Zirconia. J. Mater. Sci. 2008, 43, 6567–6570. [Google Scholar] [CrossRef]
  37. Patel, R.; Fakeeha, A.H.; Kasim, S.O.; Sofiu, M.L.; Ibrahim, A.A.; Abasaeed, A.E.; Kumar, R.; Al-Fatesh, A.S. Optimizing Yttria-Zirconia Proportions in Ni Supported Catalyst System for H2 Production through Dry Reforming of Methane. Mol. Catal. 2021, 510, 111676. [Google Scholar] [CrossRef]
  38. Xiao, Y.; Zheng, X.; Chen, X.; Jiang, L.; Zheng, Y. Synthesis of Mg-Doped Ordered Mesoporous Pd-Al2O3 with Different Basicity for CO, NO, and HC Elimination. Ind. Eng. Chem. Res. 2017, 56, 1687–1695. [Google Scholar] [CrossRef]
  39. Oliveira, D.; Andrada, A.S. Synthesis of Ordered Mesoporous Silica MCM-41 with Controlled Morphology for Potential Application in Controlled Drug Delivery Systems. Ceramica 2019, 65, 170–179. [Google Scholar] [CrossRef]
  40. Alipour, Z.; Rezaei, M.; Meshkani, F. Effect of Alkaline Earth Promoters (MgO, CaO, and BaO) on the Activity and Coke Formation of Ni Catalysts Supported on Nanocrystalline Al2O2 in Dry Reforming of Methane. J. Ind. Eng. Chem. 2014, 20, 2858–2863. [Google Scholar] [CrossRef]
  41. Zhang, L.; Zhang, Q.; Liu, Y.; Zhang, Y. Dry Reforming of Methane over Ni/MgO-Al2O2 Catalysts Prepared by Two-Step Hydrothermal Method. Appl. Surf. Sci. 2016, 389, 25–33. [Google Scholar] [CrossRef]
  42. He, L.; Liang, B.; Li, L.; Yang, X.; Huang, Y.; Wang, A.; Wang, X.; Zhang, T. Cerium-Oxide-Modified Nickel as a Non-Noble Metal Catalyst for Selective Decomposition of Hydrous Hydrazine to Hydrogen. ACS Catal. 2015, 5, 1623–1628. [Google Scholar] [CrossRef]
  43. Zhan, Y.; Han, J.; Bao, Z.; Cao, B.; Li, Y.; Street, J.; Yu, F. Biogas Reforming of Carbon Dioxide to Syngas Production over Ni-Mg-Al Catalysts. Mol. Catal. 2017, 436, 248–258. [Google Scholar] [CrossRef]
  44. Al-Fatesh, A.S.; Arafat, Y.; Kasim, S.O.; Ibrahim, A.A.; Abasaeed, A.E.; Fakeeha, A.H. In Situ Auto-Gasification of Coke Deposits over a Novel Ni-Ce/W-Zr Catalyst by Sequential Generation of Oxygen Vacancies for Remarkably Stable Syngas Production via CO2-Reforming of Methane. Appl. Catal. B Environ. 2021, 280, 119445. [Google Scholar] [CrossRef]
  45. Lin, S.; Wang, J.; Mi, Y.; Yang, S.; Wang, Z.; Liu, W.; Wu, D.; Peng, H. Trifunctional Strategy for the Design and Synthesis of a Ni-CeO2@SiO2 Catalyst with Remarkable Low-Temperature Sintering and Coking Resistance for Methane Dry Reforming. Chin. J. Catal. 2021, 42, 1808–1820. [Google Scholar] [CrossRef]
  46. Al-Fatesh, A.S.; Arafat, Y.; Atia, H.; Ibrahim, A.A.; Ha, Q.L.M.; Schneider, M.; M-Pohl, M.; Fakeeha, A.H. CO2-Reforming of Methane to Produce Syngas over Co-Ni/SBA-15 Catalyst: Effect of Support Modifiers (Mg, La and Sc) on Catalytic Stability. J. CO2 Util. 2017, 21, 395–404. [Google Scholar] [CrossRef]
  47. Wang, C.; Jie, X.; Qiu, Y.; Zhao, Y.; Al-Megren, H.A.; Alshihri, S.; Edwards, P.P.; Xiao, T. The Importance of Inner Cavity Space within Ni@SiO2 Nanocapsule Catalysts for Excellent Coking Resistance in the High-Space-Velocity Dry Reforming of Methane. Appl. Catal. B Environ. 2019, 259, 118019. [Google Scholar] [CrossRef]
  48. Li, R.; Xu, W.; Deng, J.; Zhou, J. Coke-Resistant Ni-Co/ZrO2-CaO-Based Microwave Catalyst for Highly Effective Dry Reforming of Methane by Microwave Catalysis. Ind. Eng. Chem. Res. 2021, 60, 17458–17468. [Google Scholar] [CrossRef]
  49. Fidalgo, B.; Arenillas, A.; Menéndez, J.A. Mixtures of Carbon and Ni/Al2O2 as Catalysts for the Microwave-Assisted CO2 Reforming of CH4. Fuel Process. Technol. 2011, 92, 1531–1536. [Google Scholar] [CrossRef]
  50. Kustov, L.M.; Tarasov, A.L.; Tkachenko, O.P.; Kapustin, G.I. Nickel-Alumina Catalysts in the Reaction of Carbon Dioxide Re-Forming of Methane under Thermal and Microwave Heating. Ind. Eng. Chem. Res. 2017, 56, 13034–13039. [Google Scholar] [CrossRef]
  51. Li, L.; Jiang, X.; Wang, H.; Wang, J.; Song, Z.; Zhao, X.; Ma, C. Methane Dry and Mixed Reforming on the Mixture of Bio-Char and Nickel-Based Catalyst with Microwave Assistance. J. Anal. Appl. Pyrolysis 2017, 125, 318–327. [Google Scholar] [CrossRef]
  52. Hasnan, N.S.N.; Timmiati, S.N.; Lim, K.L.; Yaakob, Z.; Kamaruddin, N.H.N.; Teh, L.P. Recent Developments in Methane Decomposition over Heterogeneous Catalysts: An Overview. Mater. Renew. Sustain. Energy 2020, 9, 8. [Google Scholar] [CrossRef] [Green Version]
Figure 1. N2 physisorption of Rh-x (x = 0, 1, 2, 3, 4, and 5 wt.%).
Figure 1. N2 physisorption of Rh-x (x = 0, 1, 2, 3, 4, and 5 wt.%).
Nanomaterials 13 00547 g001
Figure 2. XRD pattern of fresh Rh-x (x = 0, 1, 2, 3, 4, and 5 wt.%) catalyst samples. (♠) NiO and (♣) cubic phases of yttria-stabilized zirconia samples.
Figure 2. XRD pattern of fresh Rh-x (x = 0, 1, 2, 3, 4, and 5 wt.%) catalyst samples. (♠) NiO and (♣) cubic phases of yttria-stabilized zirconia samples.
Nanomaterials 13 00547 g002
Figure 3. H2-TPR profiles of fresh Rh-x (x = 0, 1, 2, 3, 4, and 5 wt.%) catalysts.
Figure 3. H2-TPR profiles of fresh Rh-x (x = 0, 1, 2, 3, 4, and 5 wt.%) catalysts.
Nanomaterials 13 00547 g003
Figure 4. CO2-TPD profiles for Ni- Rh-x (x = 0, 1, 3, 4, 5 wt.%) catalysts.
Figure 4. CO2-TPD profiles for Ni- Rh-x (x = 0, 1, 3, 4, 5 wt.%) catalysts.
Nanomaterials 13 00547 g004
Figure 5. CH4 conversion versus TOS for the Ni-Rh-x (x = 0, 1, 2, 3, 4, 5 wt.%) catalysts operated at 800 °C.
Figure 5. CH4 conversion versus TOS for the Ni-Rh-x (x = 0, 1, 2, 3, 4, 5 wt.%) catalysts operated at 800 °C.
Nanomaterials 13 00547 g005
Figure 6. CO2 conversion versus TOS for the Ni-Rh-x (x = 0, 1, 2, 3, 4, 5 wt.%) catalysts operated at 800 °C.
Figure 6. CO2 conversion versus TOS for the Ni-Rh-x (x = 0, 1, 2, 3, 4, 5 wt.%) catalysts operated at 800 °C.
Nanomaterials 13 00547 g006
Figure 7. TEM images of (A) fresh Rh-0; (B) fresh Rh-4; (C) spent Rh-0; (D) spent Rh-4.
Figure 7. TEM images of (A) fresh Rh-0; (B) fresh Rh-4; (C) spent Rh-0; (D) spent Rh-4.
Nanomaterials 13 00547 g007
Figure 8. TGA profiles for the Rh-x (x = 0, 1, 2, 3, 4, 5 wt.%) catalysts after TOS of 420 min in DRM, at 800 °C and 1 atm.
Figure 8. TGA profiles for the Rh-x (x = 0, 1, 2, 3, 4, 5 wt.%) catalysts after TOS of 420 min in DRM, at 800 °C and 1 atm.
Nanomaterials 13 00547 g008
Table 1. Textural aspects of the catalysts.
Table 1. Textural aspects of the catalysts.
Sample-TypeBET Area
(m2/g)
Pore Volume
(cm3/g)
Pore Diameter
(nm)
Rh-0310.1925.00
Rh-1270.1523.00
Rh-2270.1623.00
Rh-3270.1524.00
Rh-4280.1522.00
Rh-5270.1522.00
Table 2. H2 consumption for promoted and unpromoted Rh2O3 catalysts.
Table 2. H2 consumption for promoted and unpromoted Rh2O3 catalysts.
SampleDescriptionH2 Consumption (mol/gcat)
Rh-05% Ni /8%Y2O3 + 92% ZrO20.68
Rh-15% Ni+ 1% Rh/8% Y2O3 + 92% ZrO20.80
Rh-25% Ni+ 2% Rh/8% Y2O3 + 92% ZrO20.90
Rh-35% Ni+ 3% Rh/8% Y2O3 + 92% ZrO21.00
Rh-45% Ni+ 4% Rh/8% Y2O3 + 92% ZrO21.00
Rh-55% Ni+ 5% Rh/8% Y2O3 + 92% ZrO21.00
Table 3. H2 consumption for promoted Rh2O3 catalysts.
Table 3. H2 consumption for promoted Rh2O3 catalysts.
SampleDescriptionCO2 Consumption (μmol/gcat)
Rh-05% Ni /8%Y2O3 + 92% ZrO20.72
Rh-15% Ni+ 1% Rh/8% Y2O3 + 92% ZrO20.8
Rh-25% Ni+ 2% Rh/8% Y2O3 + 92% ZrO20.9
Rh-35% Ni+ 3% Rh/8% Y2O3 + 92% ZrO21.0
Rh-45% Ni+ 4% Rh/8% Y2O3 + 92% ZrO21.0
Rh-55% Ni+ 5% Rh/8% Y2O3 + 92% ZrO21.0
Table 4. Catalytic performance of new method compared to previous works for the dry reforming of methane.
Table 4. Catalytic performance of new method compared to previous works for the dry reforming of methane.
CatalystT
(°C)
CH4/CO2 CH4
Conversion
(%)
Ref.
Ni-2.5%Ce/W-Zr7001:178[44]
Ni-CeO2@SiO28001:180[45]
Co-Ni/Sc-SBA-157001:166[46]
Ni@SiO27001:171[47]
Ni-Co/ZrO2-CaO + SiC8001:197[48]
5.6 wt.% Ni/Al2O3 + FY58001:190[49]
10%Ni/ Al2O3 − F7001:172[50]
20%Ni/ Al2O3 (mixed with biochar)8001:179[51]
5% Ni+ 4% Rh/8% Y2O3 + 92% ZrO28001:189The present work
Table 5. Calculated thermodynamic parameter over the various catalysts at 800 °C.
Table 5. Calculated thermodynamic parameter over the various catalysts at 800 °C.
CatalystThermodynamics Parameter
ΔH800, kJ/kmolΔS800, kJ/(kmol.K)ΔG800, kJ/kmol
Rh-0+136,845.2+54.4208+78,444
Rh-1+137,654.4+54.5777+79,085
Rh-2+145,049.0+56.2075+84,730
Rh-3+150,385.0+57.2514+88,946
Rh-4+151,860.1+57.5033+90,151
Rh-5+148,103.0+56.8009+87,147
ReactionKeq (800 °C)
DRM: CH4(g) + CO2(g) = 2H2(g) + 2CO(g)151.348
RWGS: CO2(g) + H2(g) = CO(g) + H2O(g)0.912605
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Saleh, J.; Al-Fatesh, A.S.; Ibrahim, A.A.; Frusteri, F.; Abasaeed, A.E.; Fakeeha, A.H.; Albaqi, F.; Anojaidi, K.; Alreshaidan, S.B.; Albinali, I.; et al. Stability and Activity of Rhodium Promoted Nickel-Based Catalysts in Dry Reforming of Methane. Nanomaterials 2023, 13, 547. https://doi.org/10.3390/nano13030547

AMA Style

Saleh J, Al-Fatesh AS, Ibrahim AA, Frusteri F, Abasaeed AE, Fakeeha AH, Albaqi F, Anojaidi K, Alreshaidan SB, Albinali I, et al. Stability and Activity of Rhodium Promoted Nickel-Based Catalysts in Dry Reforming of Methane. Nanomaterials. 2023; 13(3):547. https://doi.org/10.3390/nano13030547

Chicago/Turabian Style

Saleh, Jehad, Ahmed Sadeq Al-Fatesh, Ahmed Aidid Ibrahim, Francesco Frusteri, Ahmed Elhag Abasaeed, Anis Hamza Fakeeha, Fahad Albaqi, Khalid Anojaidi, Salwa B. Alreshaidan, Ibrahim Albinali, and et al. 2023. "Stability and Activity of Rhodium Promoted Nickel-Based Catalysts in Dry Reforming of Methane" Nanomaterials 13, no. 3: 547. https://doi.org/10.3390/nano13030547

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