Elsevier

Journal of Catalysis

Volume 242, Issue 1, 15 August 2006, Pages 92-102
Journal of Catalysis

Syngas by catalytic partial oxidation of methane on rhodium: Mechanistic conclusions from spatially resolved measurements and numerical simulations

https://doi.org/10.1016/j.jcat.2006.05.008Get rights and content

Abstract

The mechanism for the catalytic partial oxidation of CH4 on Rh-coated α-Al2O3 foam monoliths was investigated by measuring species and temperature profiles along the catalyst axis and comparing them with numerical simulations. A thin quartz capillary connected to a quadrupole mass spectrometer was moved through the catalyst with a spatial resolution of ∼0.3 mm. Profiles were measured under autothermal operation for C/O ratios of 0.7, 1.0 and 1.3. The influence of the flow rate (5 vs. 10 l min−1) was studied for syngas stoichiometry (C/O = 1). Numerical simulations were performed with a 38 step surface mechanism using both a porous 2D-model with mass and heat transfer and a simple plug-flow model. The experimental profiles reveal complete O2 conversion within 2 mm of the catalyst entrance for all C/O ratios and flows. H2 and CO are formed partly in the oxidation zone and partly after O2 is fully converted by steam reforming. CO2 is formed in small amounts in the oxidation zone and remains constant thereafter, except for C/O = 0.7, where some water gas shift is observed. CO2 reforming does not occur under the experimental conditions. Based on the experimental findings, a two-zone picture of the reaction mechanism is proposed. The 2D numerical simulations and the measured profiles agree qualitatively for all experimental conditions. Quantitative agreement is best for syngas stoichiometry (C/O = 1.0) at 5 and 10 l min−1 flow rate. Some quantitative differences are observed for C/O = 0.7 and 1.3. The plug flow model is for all conditions inferior to the 2D model. The importance of spatial profiles for mechanism and reactor model validation is highlighted.

Introduction

The conversion of methane into hydrogen, liquid fuels, or chemicals is of growing economic importance. Natural gas, which consists mainly of methane, is available in similar amounts on earth as oil, but it is far less exploited. According to recent assessments, oil production will peak before the year 2040 [1]. Strong efforts in science and technology are underway to convert natural gas into liquids to open up a large petroleum resource that can compensate potential oil shortfalls [2].

The first step in the transformation of methane to liquids is the production of syngas. The catalytic partial oxidation (CPO) of methane (Eq. (1)) on Rh-coated foam monoliths is an efficient way to achieve this transformation [3]. In contrast to the highly endothermic steam reforming (Eq. (2)), the CPO reaction (1) is slightly exothermic.CH4+(1/2)O2CO+2H2,ΔHr=36kJmol−1,CH4+H2OCO+3H2,ΔHr=+206kJmol−1. Steam reforming on Ni requires contact times in the range of 1 s to achieve sufficient CH4 conversion. For the CPO reaction on Rh, CH4 conversion close to 100% and >90% selectivities to H2 and CO can be achieved in a few ms [4]. CPO reactors can be operated autothermally and with a much higher gas hourly space velocity than a steam reformer, making them attractive for remote gas field applications. For process integration, methane CPO supplies a H2/CO ratio of 2 that is more favorable for downstream chemistry (methanol, Fischer–Tropsch synthesis) than the higher ratio obtained by steam reforming.

There is an ongoing debate about the mechanism of the methane CPO. The present literature review focuses on results on Rh, but similar arguments may hold for other catalysts. A direct mechanism (pyrolysis–oxidation) and an indirect (combustion–reforming) mechanism are discussed in the literature [5]. The pyrolysis–oxidation mechanism assumes that H2 and CO are primary reaction products formed in the oxidation zone at the catalyst entrance. After methane pyrolysis (CH4  Cs + 4Hs), surface carbon reacts with surface oxygen to CO (Cs + Os  CO) and surface hydrogen atoms combine to H2 (Hs + Hs  H2). In contrast, the combustion-reforming mechanism postulates a two-zone model with a CH4 combustion zone at the catalyst entrance (CH4 + 2O2  CO2 + 2H2O), and H2 and CO production in a reforming zone downstream (CH4 + H2O  CO + 3H2, CH4 + CO2  2CO + 2H2, respectively).

Because the mechanisms postulate different reaction zones, spatially resolved measurements would help verify the assumptions; however, only one paper reporting intracatalyst species and temperature profiles (at high pressures, 0.2–0.8 MPa) has been published so far. Lyubovsky et al. [6] used a stack of Microlith® screens and measured species and temperature behind each screen. They argued for mainly direct formation of CO and indirect formation of H2 by steam reforming. This conclusion was drawn by extrapolating CO and H2 selectivities to zero CH4 conversion. But because the spatial resolution was only 2 mm, and the CH4 and O2 conversion at the first point were already 40% and >80%, respectively, extrapolation over such a large range could be imprecise. With the capillary technique used in the present paper, 0.3 mm resolution has been achieved.

Nonspatially resolved investigations into the mechanism are numerous in the literature. Among these are some supporting the direct mechanism [7], [8], [9], [10], [11], [12], some supporting the indirect mechanism [13], [14], [15], and others supporting a mixed mechanism [6], [16]. In some cases, the mechanistic conclusions contradict each other. However, the literature results indicate a dependence of the mechanism on (i) temperature [10], [17], [18], [19], (ii) pressure [13], (iii) Rh oxidation state [8], [11], [12], [13], [14], [16], [18], [19], (iv) catalyst loading [11], [17], and (v) nature of the support [8], [14], [15], [18], [20]. Therefore, contradictions may arise because the mechanistic conclusions cannot be extrapolated beyond the conditions of the particular experiment.

Without discussing individual results, an overall picture of the reaction can be extracted from the studies cited above. It seems that reduced Rh sites (Rh metal) are active for syngas production with CH4 dissociation as the rate-limiting step. Oxidized Rh sites (RhxOy) lead to the formation of total oxidation products (H2O, CO2). The oxidation state of the Rh surface depends on temperature and gas atmosphere. Many studies report CO2 and H2O formation after the reaction was started on an oxidized Rh catalyst, but the selectivities turn to syngas formation after a short operational time, because reduced Rh sites are formed. At high temperatures, the overall gas atmosphere is sufficiently reducing to restore metallic Rh sites quickly even if gas-phase oxygen is present. Chemisorbed oxygen can lead to total oxidation if the surface temperature is low (e.g., 500 °C). In this case H2 and CO are not desorbing quickly enough and become oxidized. The higher the temperature, the higher the selectivity to H2 and CO. Most published studies agree that CO2 reforming is unimportant in the reaction network. It seems that the support is important, especially at low Rh loadings. Some oxides can serve as an oxygen source for the Rh surface (inverse spillover of OH or H2O), even though the intrinsic catalytic activity of the support is low.

As with the experimental studies, the numerical studies published differ in their conclusions about the reaction mechanism. The original papers from Hickman and Schmidt assumed the direct mechanism, because a high-temperature surface model (19 reactions) was developed that described the experimental conversion and selectivity data reasonably well [4], [21]. In this model, CH4 dissociation on metallic Rh sites was lumped into a single step, and H2 and CO were formed according to the pyrolysis–oxidation mechanism. Except for CO2 adsorption, all adsorption–desorption steps in this original surface mechanism were reversible. Therefore, steam reforming was not excluded by definition, but CO2 reforming was excluded (low sticking coefficient of CO2 on Rh). A 2D simulation using that mechanism showed no contribution of steam reforming [22]; it was too slow under the given conditions. A plug-flow study using the 19 step surface model and a gas-phase mechanism with 227 reversible reactions (GRI-Mech 2.11) showed that gas-phase reactions were insignificant for atmospheric pressure but become important at elevated pressures (>5 bar) [22], [23], [24]. Experimental results corroborate this [13]. Gas-phase reactions are usually not considered in atmospheric pressure simulations.

The mechanism was further improved by including steam reforming and water–gas shift (38 reactions) [25]. CO2 readsorption was also included. These refinements are in better agreement with experiments showing that steam reforming on Rh is possible in ms contact times [25], [26], [27]. The 38 step surface mechanism has been validated against integral steady-state [28], [29] and transient experimental data [30], [31]. Coverage-dependent desorption energies for CO and O2 were included [30], because they are important for light-off agreement. The 38 surface reaction step mechanism without coverage dependencies [28] was used in the present publication to compare numerical simulations with experimental spatially resolved profile measurements.

Another, much more extensive C1 mechanism on Rh (104 reactions) was recently reported by Mhadeshwar and Vlachos [32]. All activation energies in this mechanism are coverage-dependent, temperature-dependent, or both. Besides methane CPO, this mechanism also considers methane reformation by water and CO2 and decomposition of oxygenates on Rh. For methane CPO, this mechanism predicts distinct oxidation (CO2 and H2O are products) and reforming (CO and H2 are products) zones. According to Mhadeshwar and Vlachos, there is no H2 formation in the oxidation zone for experiments closely matching the present work.

Our experiments and simulations were motivated by the following conclusions from the literature:

  • (i)

    Mass transport, heat transport, and chemistry are strongly coupled in this reaction. Mechanistic conclusions from low-pressure studies (e.g., temporal analysis of products) and from studies with highly diluted reactants or catalysts have to be verified for higher pressures and higher reactant concentrations ⇒ Experimental data must be measured as close to technically relevant conditions as possible.

  • (ii)

    Integral conversion and selectivity data can be described equally well by different mechanisms ⇒ Numerical simulations must be compared with experiments with higher information content-here, spatially resolved data.

Section snippets

Reactor configuration and capillary sampling

The reactor setup and capillary sampling system are shown schematically in Fig. 1. The reaction was carried out in a quartz tube (i.d. = 19 mm, length = 40 cm). High-purity reactants CH4, O2; and the internal standard Ar were fed through calibrated mass flow controllers through a side port at the bottom of the tube and left the reactor from the top for incineration. An injection needle inserted through a septum at an end port at the bottom of the tube enabled guided movement of the capillary

Species and temperature profiles

According to literature results, the mechanism for methane CPO depends most strongly on the Rh oxidation state and temperature. To study the influence of these parameters, profile measurements were performed for two different flow rates (5 and 10 l min−1) and three different C/O ratios (0.7,1.0,1.3).

Experimental results

Fig. 6 compares the experimental profiles for the reactants CH4 and O2 and the desired products H2 and CO for all C/O ratios studied (0.7,1.0,1.3) and a total flow of 5 l min−1. All species profiles, including those measured for C/O = 1.0 at 10 l min−1, show that H2 and CO are formed partly in the presence of oxygen and partly after total oxygen consumption by steam reforming. CO2 reforming does not contribute to the syngas formation (cf. Fig. 3, Fig. 4). A very small increase in CO2 for C/O = 0.7

Conclusions

The species and temperature profiles presented in this work are the first profiles published for methane CPO on Rh-coated foams. A spatial resolution of (0.3 mm) was reached. The profiles clearly show that under technically relevant conditions (autothermal operation, up to 10 l min−1 flow), syngas is formed by a combination of direct partial oxidation and steam reforming. Numerical simulations using the 38 surface reaction mechanism with a 2D porous reactor model including heat and mass

Acknowledgements

R.H. acknowledges funding through an Alexander von Humboldt Foundation Feodor Lynen Fellowship. The authors acknowledge financial support by the Department of Energy. K.A.W. thanks the Minnesota Supercomputer Institute for the generous computational resources and partial support of this work and A.B. Mhadeshwar and D.G. Vlachos for sharing their C1 mechanism on Rh.

References (44)

  • D.J. Wilhelm et al.

    Fuel Process. Technol.

    (2001)
  • S. Freni et al.

    J. Power Sources

    (2000)
  • E.P.J. Mallens et al.

    J. Catal.

    (1997)
  • W.Z. Weng et al.

    Catal. Today

    (2000)
  • L. Basini et al.

    Catal. Today

    (2001)
  • C.T. Au et al.

    J. Catal.

    (1997)
  • D. Wang et al.

    J. Catal.

    (1996)
  • K. Nakagawa et al.

    J. Catal.

    (1999)
  • S. Rabe et al.

    Appl. Catal. A Gen.

    (2005)
  • E. Ruckenstein et al.

    J. Catal.

    (1999)
  • C.T. Goralski et al.

    Chem. Eng. Sci.

    (2000)
  • G. Veser et al.

    Chem. Eng. Sci.

    (2000)
  • E.J. Klein et al.

    Stud. Surf. Sci. Catal.

    (2001)
  • J. Wei et al.

    J. Catal.

    (2004)
  • O. Deutschmann et al.

    Stud. Surf. Sci. Catal.

    (2001)
  • R. Schwiedernoch et al.

    Chem. Eng. Sci.

    (2003)
  • M.V. Twigg et al.

    Chem. Eng. Res. Des.

    (2002)
  • J.T. Richardson et al.

    Appl. Catal. A Gen.

    (2000)
  • C.Y. Zhao et al.

    Mater. Sci. Eng. A Sruct. A

    (2004)
  • K.L. Hohn et al.

    Appl. Catal. A Gen.

    (2001)
  • M.J. Stutz et al.

    Chem. Eng. Sci.

    (2005)
  • T.S. Ahlbrandt

    Int. Geol. Rev.

    (2002)
  • Cited by (210)

    • Partial oxidation process for syngas production

      2022, Advances in Synthesis Gas: Methods, Technologies and Applications: Syngas Production and Preparation: Volume 1
    View all citing articles on Scopus
    View full text