Você está na página 1de 25

5066

Ind. Eng. Chem. Res. 2003, 42, 5066-5090

Detailed Kinetics of Fischer-Tropsch Synthesis on an Industrial Fe-Mn Catalyst


Jun Yang,, Ying Liu,, Jie Chang, Yi-Ning Wang, Liang Bai, Yuan-Yuan Xu, Hong-Wei Xiang, Yong-Wang Li,*, and Bing Zhong
Group of Catalytic Kinetics & Theoretical Modeling, State Key Laboratory of Coal Conversion, Institute of Coal Chemistry, Chinese Academy of Sciences, Taiyuan 030001, Peoples Republic of China, and Department of Chemistry, Jinan University, Guangzhou 510632, Peoples Republic of China

The detailed kinetics of the Fischer-Tropsch synthesis over an industrial Fe-Mn catalyst was studied in a continuous integral fixed-bed reactor under the conditions relevant to industrial operations [temperature, 540-600 K; pressure, 1.0-3.0 MPa; H2/CO feed ratio, 1.0-3.0; space velocity, (1.6-4.2) 10-3 Nm3 kg of catalyst-1 s-1]. Reaction rate equations were derived on the basis of the Langmuir-Hinshelwood-Hougen-Watson type models for the Fischer-Tropsch reactions and the water-gas-shift reaction. Kinetic model candidates were evaluated by the global optimization of kinetic parameters, which were realized by first minimization of multiresponse objective functions with a genetic algorithm approach and second optimization with the conventional Levenberg-Marquardt method. It was found that an alkylidene mechanism based model could produce a good fit of the experimental data. This model shows that the desorption of the products and the insertion of methylene into the metal-alkylidene bond are the ratedetermining steps. The activation energy for olefins formation is 97.37 kJ mol-1 and smaller than that for the paraffin formation (111.48 kJ mol-1). In this model, the readsorption and secondary reactions of olefins are taken into account, and deviations of hydrocarbon distribution from the conventional ASF distribution can therefore be quantitatively described. However, the deeper information for the olefin-to-paraffin ratio has not intrinsically been described in the present stage, leaving for the further improvements in models to consider the transportationenhanced readsorption and secondary reaction of olefins more practically in the reactor modeling stage.
1. Introduction Fischer-Tropsch synthesis (FTS) is an industrially important process for the conversion of syngas (H2/CO) derived from carbon sources such as coal, peat, biomass, and natural gas into hydrocarbons and oxygenates. Today, it continuously attracts interest as an option for the production of clean transportation fuels and chemical feedstocks.1-3 The FTS product is composed of a complex multicomponent mixture of linear and branched hydrocarbons and oxygenated products, the majority of which are linear hydrocarbons.3,4 The FTS kinetics has extensively been studied, and many attempts have been made for the rate equations describing the FTS reactions.4-13 In most cases, the hydrocarbon products were lumped according to the carbon number of hydrocarbon molecules with an ideal Anderson-Schulz-Flory (ASF) distribution. Although a few of the available kinetic models were developed based on the detailed mechanism of the LangmuirHinshelwood-Hougen-Watson (LHHW) type,4,10-13 only two of them simultaneously considered both the syngas conversion rates and the hydrocarbon formation rates.11-13 We have pointed out the deficiencies existing in the conventional lumped models and those tailing syngas conversion rates with a carbon number distribution formula when a comprehensive simulation is required.13

Chinese Academy of Sciences. Jinan University.

The quality of a detailed kinetic model is closely related to the understanding of the mechanism in the FTS catalytic reaction system, in which a polymerization process has been recognized to be dominant; however, sufficient details on the FTS mechanistic aspect are not yet fully understood.10,14,15 In 1946, Herington first treated the molar distribution of hydrocarbons from FTS in terms of a polymerization mechanism.15,16 The same formulation was rediscovered by Anderson et al. in 1951 and named the ASF distribution.7,16 In the ASF model, the formation of hydrocarbon chains was assumed as a stepwise polymerization procedure and the chain growth probability was assumed to be independent of the carbon number. However, significant deviations from the ideal ASF distribution have been observed in many studies.17-20 Pichler et al.21 for the first time reported the deviations of experimental results from the ASF distribution. The usual deviations of the distribution of the linear hydrocarbons are a relatively higher selectivity to methane, a relatively lower selectivity to ethane, and an increase in the chain growth probability with increasing molecular size in comparison to the ideal ASF distribution. Several different explanations about the cause of these deviations have been proposed.22-24 Some authors25,26 interpreted the deviations from the standard ASF distribution by the superposition of two ASF distributions. They suspected the existence of two sorts of sites for the chain growth on the catalyst surface and, therefore, proposed that each site might individually yield the ideal ASF

10.1021/ie030135o CCC: $25.00 2003 American Chemical Society Published on Web 09/11/2003

Ind. Eng. Chem. Res., Vol. 42, No. 21, 2003 5067

distribution with different chain growth probabilities. However, this explanation cannot interpret the increase of the paraffin/olefin ratio with the chain length. On the basis of the experiments with cofed olefins, it was noted that the readsorption and secondary reaction of olefins had a great influence on the products distribution of FTS.27 Some researchers28-30 proposed a more plausible explanation for these deviations and suggested that the occurrence of secondary reactions of the olefins caused the deviations from the ASF distribution. Iglesia et al.30 developed a model describing the olefin readsorption effect enhanced by intraparticle and interparticle transport processes. They suggested that the diffusion limitation within liquid-filled pores slowed the removal of 1-olefins, which caused an increase of their residence time within the catalyst pores. In their model, the CO hydrogenation model and the olefin readsorption model were treated separately. In 1992, Zimmerman and Bukur31 developed a model for both the formation of linear hydrocarbons and the water-gas-shift (WGS) reaction over iron catalysts. In their model, the readsorption and secondary reaction of olefins were taken into account and the increase of the paraffin/olefin ratio with the carbon number can be predicted. However, the results displayed significant deviations between model predicted and experimental mole fractions, especially for methane and ethene. Kuipers et al.32 proposed a chain-length-dependent olefin readsorption mechanism to explain the fact that the paraffin/olefin ratio increases with the carbon number. On the basis of the olefin cofeeding kinetic experiments over cobalt catalyst in a continuously operated and well-mixed slurry reactor, Schulz and Claeys33 developed a kinetic model for the Fischer-Tropsch reaction system. It is well accepted that the FTS reaction is a surface polymerization reaction, but its detailed mechanism is still not fully understood. Recent mechanistic studies show new evidence in favor of a mechanism start with CO dissociation. It is generally agreed that the FTS proceeds via the dissociation of CO, further forming carbide on the surface in sequence of the hydrogenation of this carbide.10,34,35 Despite considerable research efforts, uncertainties still remain about the mechanism of chain growth in the FTS reaction. Brady and Pettit36 proposed an alkyl mechanism and suggested that the chain growth is initiated by the insertion of methylene into the adsorbed alkyl. Martinez et al.37 proposed a chain growth mechanism, in which the chain growth is initiated by an R-vinyl and the polymerization process proceeds through the reaction between methylene and an alkenyl. Joyner38 developed an alkylidene mechanism and assumed the chain growth process propagated via the successive insertion of methylene into the metal-alkylidene bond. With this model, one can easily explain the formation of R-olefins as the primary products in FTS. These proposed models have provided a basis for a workable kinetic model derived from experimental data. However, models are still far from meeting the demands for both better mechanism understanding and applications in engineering scale-up, leaving a great deal of work under the expectation for a better treatment in considering both the fundamental mechanism and self-consistency in rate expressions.10,13 Lox and Froment11,12 studied the FTS kinetics over an iron catalyst. On the basis of the LHHW scheme, they developed several sets of elementary steps, from which detailed kinetic models were derived by assuming

the carbide mechanism for the FTS and the formate intermediate scheme for the WGS reaction. The rates of CO or H2 consumption and product formation were unified in these models. Experimental data were regressed with large-scale nonlinear optimization approaches for determining the best kinetic parameters, and statistical analysis was applied to validate the feasibility of both models and parameters in them. The final model could describe the distribution of linear paraffins and olefins of FTS obeying the ideal ASF distribution at the level of surface reaction. However, the deviations, observed in many experiments, from the ideal ASF distribution were totally neglected in their model. Very recently, we have proposed a systematic approach for considering more complicated olefin readsorption phenomena in kinetic modeling on the basis of ideas similar to those of Froment.13 Although primary, namely, a limited number of mechanism sets, from which only about 10 kinetic models were scanned, were considered, our optimal model for an industrial iron catalyst showed a better description for the nonideal ASF distribution than that from Lox and Froment. However, the models based on the detailed mechanism scan for FTS processes need to first describe the most intrinsic factors involved, namely, those defined at the surface mechanism level, and the solubility/diffusivityenhanced paraffin/olefin ratio dependence on the carbon number should be described through modification of the models with intrinsic significances. This modification needs to be carefully arranged at both the surface mechanism and the reactor simulation levels because the transportation-enhanced phenomena may not fundamentally be reflected solely either in an intrinsic kinetic model or in reactor models using the intrinsic kinetics correctly. This is because reactor models can fundamentally account for transportation phenomena, and at the same time long-chain olefins staying in the catalyst pores filled with wax are difficult to remove instantly as defined even in the case of kinetic experimental conditions. In fact, this is a dilemma for the kinetics modeling of FTS. The difficulties arise, making kinetic modeling impossible, if one wants to fundamentally take the transportation phenomena into account for the modeling of a kinetic reactor. On the basis of this understanding, we first make an attempt to grasp the most significant intrinsic information at the mechanism kinetic level, and later efforts will be devoted to solving the modification problem for considering the transportation-enhanced phenomena with the combination of kinetics and reactor simulations. The goal of this paper is, therefore, to establish and test detailed mechanistic kinetic models for the Fischer-Tropsch system over an industrial Mn-Fe catalyst while considering many more possibilities in mechanism combinations than before. The new mechanistic model, in which the olefin readsorption and secondary reactions are taken into account, combines all of the FTS reactions with WGS reaction in a self-consistent way, leaving the transportation-enhanced olefin readsorption factor for further work in the combination of kinetics and reactor simulations. 2. Experimental Section The experimental setup used in this study is shown in Figure 1. A micro-fixed-bed reactor (a 1.0 m long stainless steel tube of 0.012 m in internal diameter with

5068 Ind. Eng. Chem. Res., Vol. 42, No. 21, 2003

Figure 1. Experimental setup: (1) gas cylinders; (2) pressure meters; (3) pressure regulators; (4) pressure regulators before the reactor; (5) mass flow controllers; (6) the reactor; (7) wet flowmeter; (8) needle valves; (9) backpressure regulator; (10) purification columns; (11) wax condenser; (12) oil/water condenser; (13) wax trap (hot trap); (14) cold trap; (15) high-temperature ball valves; (16) ball valves; (17) salt bath.

a temperature-controlled molten salt bath to achieve a uniform temperature profile in the catalyst bed) was used for all kinetic experiments. The catalyst used in the kinetic study is an industrial manganese-promoted iron catalyst prepared by Institute of Coal Chemistry (ICC), Chinese Academy of Sciences. The preparation method has been patented, and some open information can be found in the literature.39,41 The fresh catalyst was crushed and sieved to particles with diameters of 0.25-0.36 mm (40-60 ASTM mesh), which has been proved to be a compromising particle size safe for neglectable intraparticle transfer limitations and promisingly easy operations to the reactor during the experiment.10-13 A 3.58 g (3 cm3) catalyst was used for the kinetic experiments. The catalyst was diluted by 24 cm3 inert silica sand [(catalyst:sand ) 1:8 (v/v)] with the same mesh size range. Before the reaction, the catalyst was reduced in situ with syngas (H2/CO ) 2.0) in the reactor. In a typical reduction procedure, operation conditions were adjusted to 0.25-0.30 MPa and 1000 h-1 at 513 K, and then the temperature was raised to 548 K at the rate of 1 K h-1 and retained at that temperature for the whole reduction stage. After reduction for 36 h, the temperature was lowered to 513 K and then the operation conditions were adjusted to the desired value for the FTS. For intrinsic kinetic studies, a high space velocity is needed to exclude the external diffusion limitation. In our work, the space velocity of (1.6-4.2) 10-3 Nm3 kg of catalyst-1 s-1 is much higher than those reported by others,5,13 ensuring the kinetic experiments with the neglected external diffusion effect. The syngas was prepared by blending of pure CO and H2 (>99.99% purity). The CO and H2 passed through a series of an oxygen-removal trap, a heated silica gel molecular sieve trap, and an activated charcoal trap to remove tiny amounts of oxygen, water, and other impurities. The flow rates of CO and H2 were controlled by two Brooks 5850E mass flow controllers. The outlet of the reactor is connected with a hot trap (420 K) and then an ice trap (273 K) at the system pressure. After these product collectors, the pressure was released through a backpressure regulator. The flow rate of the tail gas was monitored by a wet test gas meter. Gas

samples for the gas chromatograph were collected after removal of the water. For kinetic studies, a stabilization period of more than 1000 h was used to ensure that the stable catalytic phases were established. This could avoid the sharp change in crystalline structures observed in previous experimental studies by other researchers.40,41-45 Kinetic samples were cumulatively collected during a typical period of 10-16 h. For each operation condition, it took at least 10 h to ensure the steady-state behavior of the catalyst after a change of the reaction conditions. The products of the synthesis were separated into three portions: gas phase, liquid phase (from the ice trap), and wax phase (from the hot trap). Both the purified syngas and the tail gas were analyzed on a gas chromatograph. H2, O2, N2, CH4, and CO were separated on a 13 molecular sieve packed column (1.5 m 3 mm i.d., Ar carrier flow) and detected with a thermal conductivity detector (TCD). C1-C5 hydrocarbons in the gas phase were analyzed on a C220/C-22 (170-250 m) packed column (7.2 m 3.2 mm i.d.) flame ionization detector (FID) with N2 as the carrier. CO2 was measured on a silica gel packed column (1 m 3 mm i.d, H2 carrier) with TCD and quantified by an external standard method. The oil product from the cold trap was separated on a 60 m 0.25 mm (i.d.) OV-101 capillary column (N2 carrier, FID) with the temperature programmed from 333 K (maintained for 20 min) to 563 K at the rate of 3 K min-1. The wax product from the hot trap was first dissolved in CS2 (0.5-1.0 mass %) and then analyzed on a OV-101 capillary column, FID, and N2 carrier with the temperature programmed (1 K min-1) from 343 to 563 K. The water phase was separated by using a BD-wax (GW, US) capillary column (N2 carrier, FID). All of the data in steady-state reactions showed promising material balances with carbon, hydrogen, and oxygen material balances between 95 and 104%. The experiences from our intensive experimental investigations of the catalytic performances of various FTS catalysts showed that better material balances are difficult to reach in practice. This is partially due to the fact that small fluctuations during a sampling time brought about certain experimental errors for the final results and partially due to the fact that the complexity of the FTS products caused difficulties in analyses. In addition, detailed kinetic modeling requires the information of the compositions of the reactants and almost all FTS products (hydrocarbons relevant) in each stream. It is estimated that about 5% error in the total material balances could produce very large errors (sometimes several times) for the contents of components with very small mole fractions in a stream. This error amplification phenomenon in FTS makes it even more difficult to achieve high modeling accuracy in detailed kinetics. 3. Kinetic Models 3.1. Active Site Assumption. The overall FTS reactions can be simplified as the combination of FTS reactions and the WGS reaction.11,12

paraffin formation: nCO + (2n + 1)H2 ) CnH2n+2 + nH2O (1) olefin formation: nCO + 2nH2 ) CnH2n + nH2O (2) WGS reaction: CO + H2O ) CO2 + H2 (3)

Ind. Eng. Chem. Res., Vol. 42, No. 21, 2003 5069

Active sites for the above reactions in FTS are still not clear especially in the cases of iron catalysts. This is because of the fact that the iron-based catalysts starting from their oxide precursors have experimentally proved to have complex phase transfer during the reduction as well as synthesis operation.40,41-45 The controversy in the active sites (or phases) in the FTS system appeals the need for much attention to be paid to the understanding of the FTS mechanism at a deep level; for example, the work for Co or Ru catalysts, for which metal phases are believed to be active phases for FTS, has been considered at even a molecular level with many interesting results achieved, casting an encouraging light on the detailed mechanism understanding.46 For iron catalysts, however, we expect many more challenges in this direction.41-49 Here, we will not go into these controversies and instead will use experimental data to fit many possible mechanism-derived kinetic models, which are anyhow expected to reflect the most important points of FTS catalysis and chemistry. Beyond the complexities in the FTS catalysis with iron catalysts, it is generally accepted that reactions in eqs 1-3 can be assumed to take place on two kinds of active sites. The active sites for the hydrocarbon formation and the WGS reaction are iron carbides and magnetite (Fe3O4), respectively.4,11-13,47-49 3.2. Kinetic Models for the Formation of Linear Hydrocarbons. 3.2.1. Elementary Reactions of the FTS. The FTS is a complex network of parallel and series reactions involving different extents and determining altogether the overall catalyst performance. The whole synthesis reaction can be simplified as the combination of FTS reactions and the WGS reaction. The FTS reactions considered here consist of surface steps in five categories:50 (1) adsorption of reactants (H2 and CO); (2) chain initiation; (3) chain propagation; (4) chain termination and desorption of products; (5) readsorption and secondary reaction of olefins. The CO adsorbed on the catalyst surface either in the molecule state or in the dissociated state18,51-53

Fischer-Tropsch catalysis cycling. The precise definition of the catalysis cycle during FTS on iron catalysts is, therefore, impossible with the current understanding of the FTS catalysis. Nevertheless, some conventional ideas/treatments may serve for the development of kinetic models. We here take advantage of the conventional idea that the building block CH2 is formed by the reaction of a surface carbon with dissociated hydrogen4,10-13,50,54,56 to build the kinetic models. It is thus assumed that surface carbon species undergoes a reaction with surface hydrogen:

Cs1 + Hs1 ) CHs1 CHs1 + Hs1 ) CH2s1

(7) (8)

or with molecular hydrogen according to an Eley-Rideal (ER) mechanism:

Cs1 + H2 ) CH2s1

(9)

Another possible pathway of the formation of CH2 starts with molecularly adsorbed carbon monoxide and successive hydrogen-assisted dissociation,4,18,55

COs1 + Hs1 ) HCOs1 + s1 HCOs1 + Hs1 ) Cs1 + H2Os1 Cs1 + Hs1 ) CHs1 CHs1 + Hs1 ) CH2s1

(10) (11) (12) (13)

or assisted by molecular hydrogen via the ER mechanism:

COs1 + H2 ) H2COs1 H2COs1 + H2 ) CH2s1 + H2O

(14) (15)

CO + s1 ) COs1 COs1 + s1 ) Cs1 + Os1

(4) (5)
There is still a controversy about the mechanism of chain growth in the FTS. The alkyl mechanism proposes that the reaction is initiated by the formation of a methyl species and that chain growth takes place by the successive insertion of methylene into the metalalkyl bond:36

where s1 is an empty catalytic site, on which hydrocarbon can be formed. The dissociative adsorption of hydrogen takes place on two adjacent free active sites.51-53

H2 + 2s1 ) 2Hs1

(6)

CH2s1 + CnH2n+1s1 ) Cn+1H2n+3s1 + s1

(16)

For the further formation of hydrogenated carbon species, mechanism studies for FTS often assumed the formation of CH2 and CH3 species,4,10-13,55 while recent characterization and theoretical studies in energetics debate for the increased possible formation of CH species on several transition-metal surfaces.46,57 For iron catalysts, the situation considered in theoretical models (or the model catalysts and operation conditions in characterization) describing the catalyst surfaces is definitely too far from a real working catalyst, which continuously changes in both the bulk phase and surfaces due to the carbonization of the iron and the hydrogenation of carbon species on the surface during reduction and reactions.40-45,47,52 The carbides on the catalyst surface may exchange carbon and hydrogen sources with reactants and surface intermediates during

The alkylidene mechanism proposes that the formation of the adsorbed ethylidene initiates the chain formation and that chain growth is facilitated by methylene insertion into the metal-alkylidene bond:38

CH2s1 + CnH2ns1 ) Cn+1H2n+2s1 + s1

(17)

Another mechanism assumes that chain growth is initiated by the adsorption of CO on the active sites already containing a hydrocarbon intermediate and then by a sequence of hydrogenation:56,58

CO + CnH2n+1s1 ) CnH2n+1s1CO CnH2n+1s1CO + H2 ) CnH2n+1s1C + H2O CnH2n+1s1C + H2 ) CnH2n+1s1CH2

(18) (19) (20)

5070 Ind. Eng. Chem. Res., Vol. 42, No. 21, 2003
Table 1. Elementary Reaction Sets for FTS FT I 1 2 3 4 5(n) 6(n) 7(n) 1 2 3 4 5(n) 6 (n) 7 (n) 1 2 3 4 5(n) 6(n) 7(n) FT IV 8(n) 0 1 2 3 4 5 6 7(n) 8(n) 9(n) 0 1(n) 2(n) 3(n) 4(n) 5(n) 6(n) CO + s1 ) COs1 COs1 + H2 ) H2COs1 H2COs1 + H2 ) CH2s1 + H2O H2 + 2s1 ) 2Hs1 CH2s1 + Hs1 ) CH3s1 + s1 CH2s1 + CH3s1 ) CH3CH2s1 + s1 CH2s1 + CnH2n+1s1 ) CnH2n+1CH2s1 + s1 CnH2n+1s1 + Hs1 ) CnH2n+2 + 2s1 CnH2n+1s1 ) CnH2n + Hs1 CO + s1 ) COs1 COs1 + H2 ) H2COs1 H2COs1 + H2 ) CH2s1 + H2O H2 + 2s1 ) 2Hs1 CH2s1 + Hs1 ) CH3s1 + s1 CH2s1 + CH3s1 ) CH3CH2s1 + s1 CH2s1 + CnH2n+1s1 ) CnH2n+1CH2s1 + s1 CnH2n+1s1 + H2 ) CnH2n+2 + Hs1 CnH2n+1s1 ) CnH2n + Hs1 CO + s1 ) COs1 COs1 + H2 ) H2COs1 H2COs1 + H2 ) CH2s1 + H2O H2 + 2s1 ) 2Hs1 CH2s1 + CH2s1 ) CH2CH2s1 + s1 CH2s1 + CnH2ns1 ) CnH2nCH2s1 + s1 CnH2ns1 + Hs1 ) CnH2n+1s1 + s1 CH3s1 + Hs1 ) CH4 + 2s1 CnH2n+1s1 + Hs1 ) CnH2n+2 + 2s1 CnH2ns1 ) CnH2n + s1 H2 + 2s1 ) 2Hs1 CO + s1 ) COs1 COs1 + Hs1 ) HCO s1 + s1 HCOs1 + Hs1 ) Cs1 + H2Os1 H2Os1 ) H2O + s1 Cs1 + Hs1 ) CHs1 + s1 CHs1 + Hs1 ) CH2s1 + s1 CH2s1 + Hs1 ) CH3s1 + s1 CH2s1 + CH3s1 ) CH3CH2s1 + s1 CH2s1 + CnH2n+1s1 ) CnH2n+1CH2s1 + s1 CnH2n+1s1 + Hs1 ) CnH2n+2 + 2s1 CnH2n+1s1 ) CnH2n + Hs1 H2 + 2s1 ) 2Hs1 CO + Hs1 ) Hs1CO CO + CH3s1 ) CH3s1CO CO + CnH2n+1s1 ) CnH2n+1s1CO Hs1CO + H2 ) Hs1C + H2O CH3s1CO + H2 ) CH3s1C + H2O CnH2n+1s1CO + H2 ) CnH2n+1s1C + H2O Hs1C + H2 ) Hs1CH2 CH3s1C + H2 ) CH3s1CH2 CnH2n+1s1C + H2 ) CnH2n+1s1CH2 CnH2n+1s1CH2 ) CnH2n+1CH2s1 CnH2n+1s1 + H2 ) CnH2n+2 + Hs1 CnH2n+1s1 ) CnH2n + Hs1 FT VI 1 2 3 4 5 6(n) 7(n) 8(n) 0 1(n) 2(n) 3(n) 4(n) 5(n) 6 7(n) 8(n) 1 2 3 4 5(n) 6 7 8 9(n) 10(n) 1 2 3 4 5(n) 6 7 8 9(n) 10(n) 11(n) 1 2 3 4(n) 5 6 7(n) 8(n) 9(n) CO + s1 ) COs1 COs1 + s1 ) Cs1 + Os1 Cs1 + H2 ) CH2s1 Os1 + H2 ) H2O + s1 H2 + 2s1 ) 2Hs1 CH2s1 + Hs1 ) CH3s1 + s1 CH2s1 + CH3s1 ) CH3 CH2s1 + s1 CH2s1 + CnH2n+1s1 ) CnH2n+1CH2s1 + s1 CnH2n+1s1 + H2 ) CnH2n+2 + Hs1 CnH2n+1s1 ) CnH2n + Hs1 H2 + 2s1 ) 2Hs1 CO + Hs1 ) Hs1CO CO + CH3s1 ) CH3s1CO CO + CnH2n+1s1 ) CnH2n+1s1CO Hs1CO + Hs1 ) Hs1C + HOs1 CH3s1CO + Hs1 ) CH3s1C + HOs1 CnH2n+1s1CO + Hs1 ) CnH2n+1s1C + HOs1 CnH2n+1s1C + Hs1 ) CnH2n+1s1CH + s1 CnH2n+1s1CH + Hs1 ) CnH2n+1s1CH2 + s1 CnH2n+1s1CH2 ) CnH2n+1CH2s1 HOs1 + Hs1 ) H2O + 2s1 CnH2n+1s1 + Hs1 ) CnH2n+2 + 2s1 CnH2n+1s1 ) CnH2n + Hs1 CO + s1 ) COs1 COs1 + s1 ) Cs1 + Os1 Cs1 + Hs1 ) CHs1 + s1 CHs1 + Hs1 ) CH2s1 + s1 CH2s1 + Hs1 ) CH3s1 + s1 CH2s1 + CH3s1 ) CH3CH2s1 + s1 CH2s1 + CnH2n+1s1 ) CnH2n+1CH2s1 + s1 H2 + 2s1 ) 2Hs1 Os1 + Hs1 ) HOs1 + s1 HOs1 + Hs1 ) H2O + 2s1 CnH2n+1s1 + Hs1 ) CnH2n+2 + 2s1 CnH2n+1s1 ) CnH2n + Hs1 CO + s1 ) COs1 COs1 + s1 ) Cs1 + Os1 Cs1 + Hs1 ) CHs1 + s1 CHs1 + Hs1 ) CH2s1 + s1 CH2s1 + CH2s1 ) C2H4s1 + s1 CH2s1 + CnH2ns1 ) CnH2nCH2s1 + s1 H2 + 2s1 ) 2Hs1 Os1 + Hs1 ) HOs1 + s1 HOs1 + Hs1 ) H2O + 2s1 CnH2ns1 + Hs1 ) CnH2n+1s1 + s1 CnH2n+1s1 + Hs1 ) CnH2n+2 + 2s1 CnH2ns1 ) CnH2n + s1 CO + s1 ) COs1 COs1 + s1 ) Cs1 + Os1 Cs1 + H2 ) CH2s1 CH2s1 + CH2s1 ) CH2CH2s1 + s1 CH2s1 + CnH2ns1 ) CnH2nCH2s1 + s1 H2 + 2s1 ) 2Hs1 Os1 + H2 ) H2O + s1 CnH2ns1 + Hs1 ) CnH2n+1s1 + s1 CnH2n+1s1 + H2 ) CnH2n+2 + Hs1 CnH2ns1 ) CnH2n + s1

FT II

FT VII

FT III

FT VIII

FT IX

FT V

FT X

Termination of the chain growth can occur via several routes. For the desorption of an olefin, a single-site mechanism was assumed that R-olefins (1alkenes) may be formed by a -hydride elimination reaction

and a dual-site reaction with an adsorbed hydrogen atom

CnH2n+1s1 + Hs1 ) CnH2n+2 + 2s1

(24)

CnH2n+1s1 ) CnH2n + Hs1


or by the desorption of adsorbed alkylidene

(21)

CnH2ns1 ) CnH2n + s1

(22)

For the desorption of a paraffin, both a single-site reaction with molecular hydrogen

CnH2n+1s1 + H2 ) CnH2n+2 + Hs1

(23)

were considered. It is generally agreed that the R-olefins desorbed from the catalyst surface can readsorb on the active sites and take place as secondary reactions, which have significant influences on the product distribution of the FTS.29,59 On the basis of the above five categories of elementary reactions, we can define 10 possible mechanisms for FTS, as shown in Table 1. 3.2.2. Rate Expressions of the Hydrocarbon Formation. On the basis of the detailed elementary reaction steps of FTS listed in Table 1, the formation

Ind. Eng. Chem. Res., Vol. 42, No. 21, 2003 5071

rates of the linear paraffins and olefins were derived. For each model, the possible rate-determining steps (RDSs) were identified, while all other steps were assumed to be at quasi-equilibrium. Before we derive the rate expression of hydrocarbon formation, first, it is assumed that the steady-state conditions are reached for both the surface composition of catalyst and the concentration of all of the intermediates involved. Second, it is assumed that the rate constant of the elementary steps for the formation of hydrocarbons is independent of the carbon number of the intermediates involved in the elementary reactions except for methane. Third, there are two types of uniformly distributed active sites respectively for FTS and WGS reactions on the catalyst surface.4 For the derivation of the rate expressions, FT III in Table 1 will be demonstrated. It is assumed that the RDSs are steps 5, 7, and 8 (model FT III). The remaining steps can be considered to be rapid and at equilibrium. The rates of formation of paraffins and olefins with n carbon atoms can thus be written as

atoms. Here we introduce a readsorption factor n, which is defined as follows:

n )

k8- PCnH2n[s1] k8+ [CnH2ns1]

(n g 2)

(30)

The chain growth probability for the carbon chain with n carbon atoms is

Rn )

k5[CH2s1] k5[CH2s1] + k7K6[Hs1]2/[s1] + k8+(1 - n) (n g 2) (31)

The concentration of surface intermediates can be expressed as a function of the partial pressure of CO, H2, and H2O by applying the pseudo-equilibrium relation

PH22PCO PH22PCO [s1] ) K [s1] (32) [CH2s1] ) K1K2K3 3 PH2O PH2O [Hs1] ) K4PH2[s1] (33)

RCH4 ) k7M[CH3s1][Hs1] ) k7MK6[CH2s1][Hs1]2/[s1] (25) RCnH2n+2 ) k7[CnH2n+1s1][Hs1] ) k7K6[CnH2ns1][Hs1]2/[s1] RCnH2n ) k8+[CnH2ns1] - k8-PCnH2n[s1] (n g 2) (26) (n g 2) (27) [CnH2ns1]

Substitution of eqs 32 and 33 in eq 31 yields

Rn )

[Cn-1H2n-2s1]

) PH22PCO [s1] k5K 3 PH2O

The pseudo-steady-state conditions are applied to the concentration of surface intermediates [CnH2ns1]:

PH22PCO k5K [s1] + k7K6K4PH2[s1] + k8+(1 - n) 3 PH2O (n g 2) (34)


Equation 28 can be rearranged as

d[CnH2nS1] ) k5[CnH2ns1][CH2s1] dt k5[Cn-1H2n-2s1][CH2s1] + k7[CnH2n+1s1][Hs1] +

[CnH2ns1] )

k5[CH2s1][Cn-1H2n-2s1] k5[CH2s1] + k7K6[Hs1]2/[s1] + k8+ k8-PCnH2n[s1]

k8+[CnH2nS1] - k8-PCnH2n[s1] ) k5[CnH2ns1][CH2s1] k5[Cn-1H2n-2s1][CH2s1] + k7K6[CnH2ns1][Hs1]2/[s1] + k8+[CnH2ns1] - k8-PCnH2n[s1] ) 0


After rearrangement, eq 28 yields

(n g 2) (28)

k5[CH2s1] + k7K6[Hs1]2/[s1] + k8+

(n g 2) (35)

Equation 35 can be rewritten in the following form:13

Xn ) RAXn-1 + BYn
[CnH2ns1] [Cn-1H2n-2s1] ) k5[CH2s1]
where Xn, RA, Yn, and B are defined as follows:

(36)

k5[CH2s1] + k7K6[Hs1]2/[s1] + k8+ 1 k5[CH2s1]

k8-PCnH2n[s1] k8+[CnH2ns1]

Xn ) [CnH2ns1] RA ) k5[CH2s1]

(n g 2) )

(37)

k5[CH2s1] + k7K6[Hs1]2/[s1] + k8+(1 - n)

(n g 2) (29)

k5[CH2s1] + k7K6K4PH2[s1] + k8+ PH22PCO [s1] k5K 3 PH2O

where k5[CH2s1] is related to the chain growth of hydrocarbon and k8+(1 - k8-PCnH2n[s1]/k8+[CnH2ns1]) is the rate of olefin formation, which is the net effect of desorption and readsorption of olefins with n carbon

PH22PCO k5K [s1] + k7K6K4PH2[s1] + k8+ 3 PH2O

(38)

5072 Ind. Eng. Chem. Res., Vol. 42, No. 21, 2003

B)

k8k5[CH2s1] + k7K6K4PH2[s1] + k8 k8-

) + (39)

Thus, the expression of the concentration of [CnH2n+1s1] is

[CnH2n+1s1] ) K6[CnH2ns1][Hs1]/[s1] ) K6K4


0.5

PH22PCO k5K [s1] + k7K6K4PH2[s1] + k8+ 3 PH2O Yn ) PCnH2n[s1]


From eq 36, we get

PH22.5PCO K [s1] 3 PH2O

Ri i)2

(n g 2) (50)

(n g 2)

(40)

The concentrations of the surface intermediates [CH3s1], [COs1], and [H2COs1] can be derived on the basis of the elementary steps 1-4 and 7 in FT III.

X2 ) RAX1 + BY2 X3 ) RA2X1 + RABY2 + BY3 X4 ) RA3X1 + RA2BY2 + RABY3 + BY4 ............ Xn ) RAn-1X1 + B

(41) (42) (43)

[CH3s1] ) K6K1K2K3K40.5

PH22.5PCO [s1] ) PH2O K6K K4 3


0.5

PH22.5PCO [s1] (51) PH2O (52)

[COs1] ) K1PCO[s1]

RAi-2Y(n-i+2) i)2

(n g 2) (44)

[H2COs1] ) K2PH2[COs1] ) K1K2PCOPH2[s1] ) K PCOPH2[s1] (53) 2


Normalization of the concentration of the sites on the catalyst surface leads to

where X1 and Yn-i+2 are defined as follows:

X1 ) [CH2s1] ) K3PH22PCO/PH2O[s1] Yn-i+2 ) PC(n-i+2)H2(n-i+2)[s1] (n g 2)

(45) (46)

1 ) [s1] + [Hs1] + [COs1] + [CH2s1] + [H2COs1] + [CH3s1] +

Thus, the concentration of [CnH2ns1] can be rewritten as

[CnH2ns1] + [CnH2n+1s1] i)2 i)2

(54)

Combining eqs 32, 33, and 49-53 with eq 54 gives

[CnH2ns1] ) RA
n-1

[CH2s1] + B[s1]

RA i)2

1 ) [s1] +
i-2

K4PH2 + K1PCO[s1] +

PC(n-i+2)H2(n-i+2) )
n
(n-i+2)H2(n-i+2)

RAn-1K PH22PCO/PH2O[s1] + B[s1] 3

RAi-2PC i)2

PH22PCO K [s1] + K1K2PCOPH2[s1] + 3 PH2O (47) PH22.5PCO PH22PCO n [s1] + K K6K40.5K 3 3 PH2O PH2O i)2 K6K4 K 3
0.5

Substituting eqs 32 and 47 into eq 30 yields

(Rj)[s1] + j)2
i

n ) (k8-/k8+) PCnH2n/ RAn-1K PCOPH22/PH2O + 3 k8

{ [
n

PH22.5PCO PH2O

(Rj)[s1] i)2 j)2

(55)

The concentration of free active site [s1] can thus be expressed as follows:

k5K PCOPH22/PH2O[s1] + k7K6K4PH2[s1] + k8+ 3

i)2

(RAi-2PC(n-i+2)H2(n-i+2))

According to the definition of chain growth probability Rn in eq 34, we have

]}

[s1] ) 1/ 1 + (n g 2) (48)

K4PH2 + K1PCO + PH22PCO PH2O + K1K2PCOPH2 + + K 3 PH22PCO PH2O

K 3

[CnH2ns1] ) [CH2s1]

Ri ) i)2
(n g 2) (49)

K6K4 K 3

0.5

PH22.5PCO PH2O

(Rj) + i)2 j)2 (Rj) i)2 j)2


n i

PH22PCO n K [s1] Ri 3 PH2O i)2

K6K4 K 3

0.5

PH22.5PCO PH2O

(56)

Ind. Eng. Chem. Res., Vol. 42, No. 21, 2003 5073

Substituting eq 56 into eqs 49 and 50 yields

[CnH2ns1] ) K 3 PH2 PCO PH2O


2 n

RCnH2n ) k8+(1 - n)K PH22PCO/PH2O 3

Ri/ 1 + i)2

K4PH2 + K1PCO + PH22.5PCO PH2O

Rj / j)2
+ PH2O +
i

1+

K4PH2 + K1PCO + K 3

PH22PCO

K 3

PH22PCO PH2O
n

K1K2PCOPH2 + K6K40.5K 3 +

PH22.5PCO PH2O

+ K1K2PCOPH2 + K6K4 K 3
i

0.5

K 3

PH22PCO PH2O

(Rj) + K6K40.5K3 i)2 j)2

PH22.5PCO PH2O

(Rj) i)2 j)2

PH22PCO n K (1 + K6 K4PH2) 3 P H 2O i)2

(Rj) j)2

(61)

(n g 2) (57) [CnH2n+1s1] ) K6K4 K 3


0.5

PH22.5PCO PH2O

Ri/ 1 + i)2

K4PH2 + K1PCO + PH22.5PCO PH2O


n i

K 3

PH22PCO PH2O
n

+ K1K2PCOPH2 + K6K4 K 3
i

0.5

K 3

PH22PCO PH2O

i)2 j)2

(Rj) + K6K4 K 3

0.5

PH22.5PCO P2

(Rj) i)2 j)2

3.3. Kinetic Models for the WGS Reaction. The kinetics of the WGS reaction has been intensively studied by many researchers, and several mechanisms were proposed.60-64 It is generally accepted that the WGS reaction over supported iron catalysts proceeds via a mechanism of formate species due to a limited change of oxidation states of the iron cations.4,60-64 Rethwisch and Dumesic61 suggested that the WGS reaction proceeds on active sites different from those for FTS, and for supported iron catalysts, the magnetite is the most active phase for the WGS. On the basis of the formate intermediate mechanism, the elementary steps of the WGS, and corresponding expressions of rates, equilibrium constants are summarized in Table 2. If we assumed that the slowest step in the WGS is step IV (WGS 3), the rate of CO2 formation can be written as follows:

(n g 2) (58)
Substituting eqs 57 and 58 into eqs 25-27, we have

RCO2 ) r4 ) kWGS4[COOH-s2] k-WGS4PCO2[H-s2] (62)


From the elementary step listed in Table 3, we have

RCH4 )
3

k7MK4K6K 3 PH2 PCO PH2O


2

PH2 PCO / 1+ PH2O

[CO-s2] ) KWGS1PCO[s2] K4PH2 + K1PCO + PH2 PCO PH2O


i 2.5

(63) (64)

KWGS2PH2O[s2]2 ) [OH-s2][H-s2]

KWGS3[OH-s2][CO-s2] ) [COOH-s2][s2] (65) + [H-s2] ) KWGS5PH2[s2] (66)

K 3

+ K1K2PCOPH2 + K6K40.5K 3

PH22PCO n K (1 + K6 K4PH2) 3 PH2O i)2 RCnH2n+2 ) k7K4K6K 3 PH23PCO PH2O

(Rj) j)2

] ]

(59)

On the basis of eqs 63-66, the concentrations of the surface intermediates [CO-s2], [OH-s2], and [COOHs2] can be derived as follows:

[CO-s2] ) KWGS1PCO[s2]

(67)

Rj / j)2

[OH-s2] ) KWGS2KWGS5-0.5PH2OPH2-0.5[s2] (68) K4PH2 + K1PCO + PH22.5PCO PH2O


i 2

1+

[COOH-s2] ) KWGS1KWGS2KWGS3KWGS5-0.5PH2OPCOPH2-0.5[s2] (69)


If Kp is used to represent the equilibrium constant of the WGS reaction, it can be expressed by the equilibrium partial pressures of CO, CO2, H2, and H2O.

K 3

PH22PCO PH2O

+ K1K2PCOPH2 + K6K4 K 3

0.5

PH22PCO n (1 + K6 K4PH2) K 3 PH2O i)2

(Rj) j)2

(60)

Kp )

/ / PCO2PH2 / P/ PH2O CO

(70)

5074 Ind. Eng. Chem. Res., Vol. 42, No. 21, 2003
Table 2. Elementary Steps and Corresponding Expressions of Rates and Equilibrium Constants for the WGS Reaction step I II III IV V elementary reaction CO + s2 ) COs2 H2O + 2s2 ) OHs2 + Hs2 COs2 + OHs2 ) COOHs2 + s2 COOHs2 ) CO2 + Hs2 2Hs2 ) H2 + 2s2 expressions of rates and equilibrium constants KWGS1 ) [COs2]/PCO[s2] KWGS2 ) [OHs2][Hs2]/PH2O[s2]2 KWGS3 ) [COOHs2][s2]/[COs2][OHs2] r4 ) kWGS4[COOHs2] - k-WGS4PCO2[Hs2] KWGS4 ) kWGS4/k-WGS4 1/KWGS5 ) PH2[s2]2/[Hs2]2

Table 3. Rate Expressions for the WGS Reaction model WGS1 WGS2 WGS3 RDS step I step III step IV rate expression RCO2 ) kv(PCO - PCO2PH2PH2O-1/Kp)/(1 + KvPCOPH2/PH2O) RCO2 ) kv(PCOPH2O/PH20.5 - PCO2PH20.5/Kp)/(1 + KvPH2O/PH20.5)2 RCO2 ) kv(PCOPH2O/PH20.5 - PCO2PH20.5/Kp)/(1 + KvPCOPH2O/PH20.5)

According to eqs 63, 64, and 66, the partial pressures of CO, H2O, and H2 can be written as follows:

Thus, the expressions of [H-s2] and [COOH-s2] can be obtained by substituting eq 80 into eqs 66 and 78.

P/ ) [CO-s2]/(KWGS1[s2]) CO
/ PH 2 O

(71)
2

[H-s2] )

KWGS5PH2 1 + KpKWGS4KWGS50.5PCOPH2O/PH20.5 KpKWGS4KWGS50.5PCOPH2O/PH20.5 1 + KpKWGS4KWGS50.5PCOPH2O/PH20.5

(81)

) [OH-s2][H-s2]/(KWGS2[s2] )

(72) (73)

/ PH2 ) [H-s2]2/(KWGS5[s2]2)

[COOH-s2] ) (82)

If we assumed that step IV reached the equilibrium state, the expression of partial pressure of CO2 at the / equilibrium state (PCO2) can be obtained:
/ PCO2

) [COOH-s2]*/(KWGS4[H-s2]*)

(74)

Finally, the rate expression of CO2 formation can be expressed as

Substituting eq 71-74 into eq 70 gives

RCO2 ) (75)

kv(PCOPH2O/PH20.5 - PCO2PH20.5/Kp) 1 + KvPCOPH2O/PH20.5

(83)

Kp )

KWGS1KWGS2[COOH-s2]*[s2]* KWGS4KWGS5[CO-s2]*[OH-s2]*

When eq 65 is substituted into eq 75, Kp can be expressed as

Kp )

KWGS1KWGS2KWGS3 KWGS4KWGS5

where Kv ) KpKWGS4KWGS50.5 and kv ) kWGS4Kv. By different assumptions of the RDS in WGS, other rate expressions of CO2 of the WGS reaction can be derived, and the results are tabulated in Table 3. The equilibrium constant Kp of the WGS reaction can be calculated by the following relation:11-13

(76) Kp )

Equation 76 can rewritten as follows:

5078.0045 - 5.8972089 + 13.958689 T 10-4T - 27.592844 10-8T 2 (84)

KWGS3 )

KWGS4KWGS5KP KWGS1KWGS2

(77)

A total of 39 kinetic models of FTS can be obtained by the combination of 13 FTS models with 3 WGS models. 4. Results and Discussion 4.1. Experimental Results. Iron-based FischerTropsch catalysts, once exposed in the synthesis gas under typical reduction or initial reaction conditions, are normally reduced by H2 and CO, transforming from their oxide phases to metallic and carbide phases.40,41-45 During this transformation stage, the phases of catalysts greatly change with time on stream as well as changes in operation conditions, leading to irreproducible experimental data. To minimize (it can never be completely eliminated in FTS over iron catalysts) the effect of a sharp phase change on the quality of kinetic data, the catalyst used for kinetic tests was reduced and then stabilized at a fixed operation condition for about 1000 h, aiming at the complete establishment of the catalyst phase for FTS. After this stabilization stage, operation conditions were switched according to kinetic

Substituting eq 77 into eq 69 yields

[COOH-s2] ) KpKWGS4KWGS50.5PCOPH2OPH2-0.5[s2] (78)


As a result of the assumption that the RDS is step IV, it can be considered that the concentration of the adsorbed species [COOH-s2] is much larger than those of the other adsorbed species.

[COOH-s2] + [s2] ) 1

(79)

The concentration of the empty active site, [s2], can be obtained by substituting eq 78 into eq 79

[s2] )

1 (80) 1 + KpKWGS4KWGS50.5PCOPH2O/PH20.5

Ind. Eng. Chem. Res., Vol. 42, No. 21, 2003 5075

Figure 2. Catalyst stability with time on stream (H2/CO ) 2.0; P ) 23 bar; GHSV ) 5.1 10-4 Nm3 kg of catalyst-1 s-1). Table 4. Catalyst Performance with Time on Stream during the Stability Test time on stream (h) 34 temperature (K) pressure (MPa) GHSV (10-4 Nm3 kg of catalyst-1 s-1) CO conversion % H2 conversion % H2/CO (in feed gas) H2/CO (in tail gas) selectivity CO2 (%) selectivity (wt %) C1 C20 C2) C30 C3) C40 C4) C2-4)/tolHC (wt %) C2-4)/C2-40 C5+ 543 2.30 5.1 94.8 43.7 2.0 21.55 36.09 7.51 3.00 5.31 1.76 7.63 1.51 6.04 18.99 2.89 67.24 83 543 2.30 5.1 94.6 40.1 2.0 22.13 40.88 9.39 4.26 5.18 2.18 8.15 1.83 6.19 19.51 2.19 62.82 180 556 2.30 5.1 96.1 44.3 2.0 28.33 38.71 13.35 6.17 5.35 2.77 8.82 2.07 6.58 20.75 1.70 54.89 273 556 2.30 5.1 96.9 47.6 2.0 34.35 39.13 14.44 7.04 4.81 2.94 8.91 2.11 6.46 20.18 1.48 53.29 404 556 2.30 5.1 95.9 41.7 2.0 28.61 36.73 15.62 7.41 4.87 3.32 8.61 2.31 6.68 20.16 1.38 51.19 516 556 2.30 5.1 96.4 44.6 2.0 30.43 39.27 16.80 8.20 4.91 3.55 9.08 2.51 6.63 20.62 1.28 48.33 641 556 2.30 5.1 96.0 47.4 2.0 25.95 40.37 16.76 8.20 4.85 3.63 9.28 2.51 7.05 21.18 1.30 47.72 845 556 2.30 5.1 96.1 45.8 2.0 27.25 38.70 16.84 8.24 4.39 3.59 8.47 2.49 6.37 19.23 1.18 49.61 960 556 2.30 5.1 96.0 44.8 2.0 28.21 38.80 16.90 8.30 4.41 3.63 8.50 2.48 6.40 19.21 1.17 49.40

sampling arrangements. The reaction results during the stabilization stage are summarized in Table 4 and Figure 2. From these results, it can be found that the catalyst reached a stable stage after about 500 h on stream under the conditions used. The selectivity to CO2 was maintained at the level of 40 ( 2%, that to CH4 at 16 ( 1%, that to C5+ at 50 ( 2%. For the activity, CO conversion reached to a level of 95 ( 1%. This confirmed that kinetic data sampling may be based on a rather stable stage of the catalyst in the following kinetic experimental stage. Kinetic sampling conditions were arranged according to orthogonal arrangement of sampling points, enabling efficient and optimal distribution of experimental points.13 The reaction condition variables that need to be considered are temperature, pressure, and H2/CO ratio. For each variable, four values were planned, bringing about 16 experimental points (corresponding to an orthogonal

table of L16(43)). The experimental results obtained in this investigation for different reaction temperatures, total reaction pressures, space velocities, and H2/CO ratios are given in Table 5. Five more points were added to consider more reaction conditions, among which one point (no. 16 in Table 5) was the reference point (no. 12 in Table 5). The order of kinetic experiments was from no. 1 to no. 21 listed in Table 5. From the results listed, it can be found that the CO and H2 conversions at no. 16 well reproduced those at no. 12, indicating that the catalytic phase reached a stable state. 4.2. Estimation of the Kinetic Parameters. 4.2.1. Reactor Model of the Fixed Bed. The reactor model for describing the kinetic experimental conditions is assumed to be a plug-flow homogeneous state. At this stage, the transportation in catalyst pores and the solubilities of different hydrocarbons are not considered in order to avoid the unsolvable difficulties in the

5076 Ind. Eng. Chem. Res., Vol. 42, No. 21, 2003
Table 5. Operation Conditions and Resulting Product Quantities no. 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 T (C) 283.0 283.2 283.3 283.1 283.2 297.0 297.2 297.1 297.2 297.1 297.0 312.2 312.3 312.1 312.0 312.2 328.5 328.4 328.3 328.4 328.1 P (MPa) 1.98 2.51 3.05 1.50 3.05 2.02 2.51 3.01 3.02 1.50 2.05 2.02 3.02 2.50 1.50 2.02 2.50 2.02 1.51 3.02 2.50 H2/CO ratio 2.05 2.62 3.08 1.03 3.06 1.03 3.13 2.58 2.59 2.07 2.05 3.05 2.04 1.02 2.55 3.05 2.04 2.55 3.05 1.02 2.55 Fin (mL/min) 342.0 348.6 435.6 343.6 349.3 449.7 462.9 456.1 606.9 454.7 342.0 684.6 686.9 678.6 686.9 684.8 892.6 894.1 896.1 908.4 1378.9 reaction time (h) 10.60 10.67 7.01 12.86 11.33 6.02 10.83 7.51 10.22 10.51 6.02 7.03 13.25 11.04 9.01 7.02 5.51 7.01 8.05 8.04 4.50 CO convn (%) 40.9 57.0 57.6 21.8 70.1 35.8 69.0 74.9 59.2 35.2 66.1 63.3 70.9 42.6 38.9 63.8 67.4 59.3 45.0 63.9 66.1 H2 convn (%) 28.2 28.9 26.8 21.0 28.9 27.6 25.8 36.6 33.7 26.8 34.5 27.7 38.0 31.1 17.1 27.1 35.0 25.5 17.3 47.2 32.4 Vexit (mL/min) 308.9 292.5 375.8 341.6 287.8 404.3 408.7 346.7 482.6 411.5 270.2 575.0 492.5 599.4 672.4 569.2 713.7 788.7 866.4 665.2 1123.5 oil (g) 2.46 3.62 2.16 1.92 3.52 2.28 4.21 3.89 5.83 2.71 2.46 3.02 12.56 9.18 3.23 3.53 5.75 4.60 2.82 13.33 5.25 wax (g) 2.14 2.83 1.85 3.36 3.24 3.06 3.41 2.64 4.27 2.02 1.62 0.78 4.55 7.49 0.58 0.68 1.96 1.20 0.83 6.72 1.20 water (g) 7.20 8.04 8.40 5.10 14.30 4.50 15.52 10.02 15.77 8.20 4.60 12.70 27.59 12.60 11.50 13.23 14.00 15.40 13.30 15.10 16.50

Table 6. Stoichiometric Coefficients of the Matrix for FTS Reactions component reactants products CO H2 CO2 H2O CH4 C2H4 C2H6 ... CnH2n CnH2n+2 CO + H2O CO2 + H2 -1 1 1 -1 0 0 0 0 0 CO + 3H2 CH4 + H2O -1 -3 0 1 1 0 0 0 0 2CO + 4H2 C2H4 + 2H2O -2 -4 0 2 0 1 0 0 0 reaction path 2CO + 5H2 C2H6 + 2H2O -2 -5 0 2 0 0 1 0 0 nO + 2nH2 CnH2n + nH2O -n -2n 0 n 0 0 0 1 0 nO + (2n + 1)H2 CnH2n+2 + nH2O -n -(2n + 1) 0 n 0 0 0 0 1

integration of the reactor model embedded in a parameter optimization procedure. This simplification may lose the transportation-enhanced readsorption phenomena, and further modification to the kinetic models obtained is needed. This will be discussed elsewhere. From the group of equations (1)-(3), the total number of the components, Nc, and of the reactions, NR, are 2N + 3 and 2N, respectively, where N is the maximum carbon number of hydrocarbons considered. The matrix of the stoichiometric coefficients between the products and reactions according to the FTS and the WGS reaction is listed in Table 6.11-13 The model of the fixed-bed reactor used in our kinetic study can be described as follows:11-13

(RCnH2n+2, RCnH2n, RCO2, ...). The partial pressure of the ith component can be calculated by using the following formula:

Pi )

mi

mi i)1
where PT is the total pressure in the reactor. Numerical integration of the continuity equation (85) is performed by using Gears method.65 4.2.2. Optimization Method. In the estimation of parameters of the kinetic model, the Levenberg-Marquardt (LM) algorithm still plays an important role. However, in most cases, the objective function based upon the nonlinear and experimental data frequently contains more than one minimum.66 As a general algorithm, LM-type continuum methods often fail in locating global minima. In our paper, to avoid getting trapped in local minima, the parameters of the various rival models in this paper were estimated in a first step using the genetic algorithm (GA) approach that was developed in this group67 and then using the LM algorithm to make refined optimization, after which the statistical tests and the physiochemical constraints are used to evaluate the significance of models and parameters. The GA algorithm encoded in this group using a real-number code chromosome representation to

Nc

PT

(i ) 1, 2, 3, ..., Nc)

(87)

dmi dW

RijRj j)1

NR

(i ) 1, 2, ..., Nc; j ) 1, 2, ..., NR) (85)

with the initial condition

W ) 0; mi ) mi,0

(86)

where W is the weight of the catalyst used, mi and mi,0 are the mole flow rates of component i along the reactor axis and at the inlet of the reactor, NR is the number of reactions involved, Nc is the total number of components, Rij is the stoichiometric coefficient for the ith component in reaction j, and Rj is the rate of reaction j

Ind. Eng. Chem. Res., Vol. 42, No. 21, 2003 5077
Table 7. Values of the Parameters for the Mechanism FT III WGS3a parameter k5,0 E5 k7M,0 E7M k7,0 E7 k8,0 E8 kv,0
a

value 7.88 75.52 2.01 106 97.39 1.10 106 111.48 8.79 103 97.37 3.42 103

unit mol kJ mol-1 mol g-1 s-1 bar-1 kJ mol-1 mol g-1 s-1 bar-1 kJ mol-1 mol g-1 s-1 kJ mol-1 mol g-1 s-1 bar-1.5 g-1 s-1 bar-1

parameter Ev k-8 Kv K1 K2 K3 K4 K6

value 58.43 2.77 10-5 2.76 10-2 2.59 1.67 10-3 8.34 10-2 1.21 0.10

unit kJ mol-1 mol g-1 s-1 bar-1 bar-0.5 bar-1 bar-1 bar-1

All of the energetic values are estimated to be in the 95% confidence level, and frequencies may be lowered to 90%.

ensure a wide searching space for optimal kinetic parameters. The GA procedure consists mainly of four parts: randomly producing an initial population, randomly selecting two individuals from the population space to crossover (using certain operator), producing the next generation of population, nonuniform mutation, and repeating these steps to scan the whole searching space. For estimation of the models, 20 model responses in the regression procedure were the outlet concentrations of 15 paraffins and olefins (such as CH4, C2H4, C2H6, C3H6, C3H8, C4H8, C4H10, etc.) selected as the most significant Fischer-Tropsch products, along with the concentrations of CO2, the conversions of CO and H2, and the overall concentrations of C5+ and C11+ hydrocarbons. The objective function is defined as follows:

The relative residual (RR) between experimental and calculated values of responses will be used to show the deviations between the model and experiment

RR )

mi,exp - mi,cal 100 mi,exp

(91)

The dependence of the reaction rate parameters on temperature can be described by the Arrhenius law.

ki(T) ) ki,0 exp(-Ei/RT)

(92)

fi,obj )

mi,exp - mi,cal mi,exp

(i ) 1, 2, ..., Nexp) (88)

where mi,exp and mi,cal are the experimental and calculated values of conversions of reactants or the selectivities of products, respectively, and Nexp is the total number of experimental runs. Because of the complexity of the models, a multiresponse objective function should be introduced, which is expressed in the following form:
Nresp Nexp

FTol,obj )

i)1 j)1

Wi

mij,exp - mij,cal mij,exp

(i ) 1, 2, ..., Nresp; j ) 1, 2, ..., Nexp) (89)


where Wi represents the weighting factor of the ith objective function (it expresses the relative importance of the ith response). Those responses with the most accurate measurement and/or with special significance in the regression are provided with greater weights): Nresp is the number of responses in the system, mij,exp is the experimental value of the ith response for the jth kinetic experiment, and mij,cal is the calculated value of the ith response for the jth kinetic experiment. The adjustable model parameters for the several kinetic models were calculated by minimizing FTol,obj with the GA and LM methods. The accuracy of the fitted model relative to the experimental data was obtained from the MARR (mean absolute relative residuals) function:
Nresp Nexp

MARR )

j)1 i)1

mi,exp - mi,cal mi,exp

NrespNexp

100 (90)

4.2.3. Parameter Estimation. For scanning the models by parameter optimization, several basic physical criteria are applied: the rate constants and equilibrium constants should be positive, and the activation energies for the paraffins and olefin formation and for the WGS reaction should be in the range of values reported by other researchers. The optimization procedure by the GA and LM methods has been set to find the minimum, which provided (1) a reasonable fit to the experimental data, (2) physically meaningful values of the model parameters, and (3) acceptable values of statistical parameters, such as F values for the models and t values for the parameters. A total of 25 kinetic models are rejected because of unreasonable values of the parameters. The discrimination of the remaining 14 rival models is performed by comparing first their MARR and second statistical tests on models (F test) and parameters (t test). The result shows that only two kinetic models MARR are within the 20%. The two models are FT III WGS3 (MARR 18.6%) and FT VI WGS3 (MARR 19.2%). 4.3. Results and Discussion. The parameters of the remained best model are listed in Table 7, and all of them are statistically significant. The estimated activation energy for chain growth, E5, is 75.52 kJ mol-1, indicating that the methylene insertion into the metal-alkylidene bond has a moderate height of energy barrier that it needs to overcome. The activation energy for methane formation, E7M, is 97.39 kJ mol-1, which is in good agreement with that reported by Vannice58 (89 kJ mol-1) and smaller than that for other paraffin formation, E7, with a value of 111.48 kJ mol-1, which can explain the higher selectivity of methane than those of other paraffins. The activation energy of olefin formation is 97.37 kJ mol-1, much smaller than that of paraffins, which can interpret the general fact that the selectivity is much higher to alkenes on the Fe-Mn catalysts than on other iron-based catalysts.67 This is different from Dictor and Bell69 (80-90 kJ mol-1 for paraffins and 100-110 kJ mol-1 for olefins) as well as Lox and Froment11,12 (94.5 kJ mol-1 for paraffins and

5078 Ind. Eng. Chem. Res., Vol. 42, No. 21, 2003

Figure 3. Comparison of calculated and experimental CO conversions (FT III WGS3).

Figure 5. Comparison of the calculated and experimental product distribution (FT III WGS3 reaction conditions: T ) 556 K, P ) 2.51 MPa, H2/CO ) 2.62, GHSV ) 1.6 10-3 Nm3 kg of catalyst-1 s-1).

Figure 4. RRs of calculated and experimental CO2 selectivity (FT III WGS3).

Figure 6. Comparison of the calculated and experimental product distribution (Lox and Froments model reaction conditions: T ) 556 K, P ) 2.51 MPa, H2/CO ) 2.62, GHSV ) 1.6 10-3 Nm3 kg of catalyst-1 s-1).

132.3 kJ mol-1 for olefins) over a Fe-Cu-K catalyst. It is evident that all of the estimated activation energies of the hydrocarbon formation are mostly within those reported in the literature (all between 70 and 105 kJ mol-1).11,12,17,69 The activation energy of 58.43 kJ mol-1 for the WGS reaction in this work is comparable with the values reported in the literature.6,8,11,12,70 A comparison between the experimental and calculated values of conversion of CO and selectivities of CO2 are presented in Figures 3 and 4. The figures show that the RRs between the model and experiment are mostly within 20%. Figures 5-10 show the comparison of experimental and calculated product distributions. Figure 5, 7, and 9 are predicted by model FT III WGS3 and Figure 6, 8, and 10 by Lox and Froments ASF model. The ASF model appears to give a strong deviation for the selectivity to hydrocarbons, lower to methane and higher to other hydrocarbons. As shown in these figures, the selectivities to olefins predicted with the ASF type model are lower than those to paraffins, in contrast with the experimental results. The modeled product distributions using FT III WGS3 are in good agreement with the experimental selectivities, and the deviation for methane is described fairly accurately. It should be noted that the current kinetic model has not considered the effect of diffusivity and solubility of olefins with different chain sizes on the paraffin/olefin ratio on the basis of the understanding of different

Figure 7. Comparison of the calculated and experimental product distribution (FT III WGS3 reaction conditions: T ) 585 K, P ) 3.02 MPa, H2/CO ) 2.04, GHSV ) 3.2 10-3 Nm3 kg of catalyst-1 s-1).

orientations of kinetic and reactor models in considering the intrinsic mechanism information and the transportation effects. The kinetic model developed is significant only for the cases excluding the diffusivity and solubility factors, which are believed to be significant in enhancing the olefin readsorption and secondary reaction and, therefore, in changing the olefin and paraffin selectivities. However, kinetic experiments can never exclude the diffusivity and solubility (mobility) especially of

Ind. Eng. Chem. Res., Vol. 42, No. 21, 2003 5079

Figure 8. Comparison of the calculated and experimental product distribution (Lox and Froments model reaction conditions: T ) 585 K, P ) 3.02 MPa, H2/CO ) 2.04, GHSV ) 3.2 10-3 Nm3 kg of catalyst-1 s-1).

Figure 11. Comparison of the calculated and experimental olefin/ paraffin ratios for different carbon numbers (FT III WGS3 reaction conditions: T ) 585 K, P ) 3.02 MPa, H2/CO ) 2.04, GHSV ) 3.2 10-3 Nm3 kg of catalyst-1 s-1).

paraffin ratio dependence on the carbon number if the transportation enhancement is not considered.19 We have recently tested the kinetic models developed in this group in a reactor simulation task, in which all of the transportation effects are comprehensively considered thanks to the nonintrinsic factor excluded kinetic models, and the primary results recovered the experimentally observed olefin/paraffin ratio dependence correctly. This later work is too extensive to present here and will be discussed elsewhere.

Figure 9. Comparison of the calculated and experimental product distribution (FT III WGS3 reaction conditions: T ) 601 K, P ) 1.51 MPa, H2/CO ) 3.05, GHSV ) 4.2 10-3 Nm3 kg of catalyst-1 s-1).

5. Conclusions Kinetic experiments of the Fischer-Tropsch reaction over an industrial Fe-Mn ultrafine particle catalyst are carried out for a wide range of industrially relevant conditions. Different reaction equation combinations were evaluated by global parameter optimization, which involved the minimization of the multiresponse objective function by a GA approach and a LM method. It was found that a kinetic model for the FTS based on the alkylidene mechanism gives the best fit of the experimental data. The best model shows that two elementary steps, the insertion of methylene into the metalalkylidene bond and the desorption of hydrocarbon products, are intrinsically slower than the others in FTS and the RDS of the WGS reaction is the desorption of CO2 via formate intermediate species. The activation energy for alkene formation is 97.37 kJ mol-1 and is much smaller than that for paraffins formation, 111.48 kJ mol-1, which can interpret the higher alkene selectivity on Fe-Mn catalysts than on other iron-based catalysts.

Figure 10. Comparison of the calculated and experimental product distribution (Lox and Froments model reaction conditions: T ) 601 K, P ) 1.51 MPa, H2/CO ) 3.05, GHSV ) 4.2 10-3 Nm3 kg of catalyst-1 s-1).

Acknowledgment heavy olefins. Figure 11 shows the dependence of the olefin-to-paraffin ratio on the carbon number. It can be seen that the kinetic model without considering the transportation effect (diffusivities and solubilities) has not predicted the correct dependence compared with experimental results. This again proved the understanding in FTS that pure olefin readsorption and secondary reaction have little consequence to the observed olefin/ Financial support from the Chinese Academy of Sciences (Project No. KGCX1-SW-02), Committee of Science and Technology of China via 863 plan (Project No. 2001AA523010), Shanxi Natural Science Foundation (20031032), and the National Natural Sciences Foundation of China (Project No. 29673054) is gratefully acknowledged.

5080 Ind. Eng. Chem. Res., Vol. 42, No. 21, 2003

Appendix: Rate Expressions of the Different Models of FTS Model FT I: RDS, steps 5-7.

RCH4 ) k6MK4PH2R1/(1 + RCnH2n+2 ) k6K4PH2

K4PH2 + K1PCO + K PH22PCO/PH2O + K1K2PH2PCO + 3

K4PH2
n

Rj))2 i)1 j)1


(
i

(A1)

j)1

Rj /(1 +
n

K4PH2 + K1PCO + K PH22PCO/PH2O + K1K2PH2PCO + 3

K4PH2

Rj))2 i)1 j)1


(

(n g 2) (A2)

RCnH2n ) k7+(1 - n) K4PH2

Rj /[1 + j)1

K4PH2 + K1PCO + K PH22PCO/PH2O + K1K2PH2PCO + 3 K4PH2 k7-

n ) (k7 /k7 ) PCnH2n/ R1RA

{[

(Rj)] i)1 j)1


)

n-1

+ k5K 3 PCOPH2 PH2O


2

(RAi-2PC i)2

(n-i+2)H2(n-i+2)

[s1] + k6 K4PH2[s1] + k7+

Rn )

k5K PH22PCO/PH2O 3 k5K PH22PCO/PH2O + k6 K4PH2 + k7+(1 - n)/[s1] 3 R1 ) k5K PH21.5PCO/PH2O 3 k5K PH21.5PCO/PH2O + k6 K4 3 k5K 3 RA ) PH22PCO [s1] PH2O

]}

(n g 2) (A3)

(n g 2) (A4)

(n g 2)

(A5)

(A6)

PH22PCO k5K [s1] + k6 K4PH2[s1] + k7+ 3 PH2O

(A7)

Model FT II: RDS, steps 5-7.


0.5 1.5

RCH4 ) k6MK4 PH2 R1/ 1 + RCnH2n+2 ) k6K4 PH2 RCnH2n ) k7 (1 - n) K4PH2


+ 0.5 1.5

K4PH2 + K1PCO + K 3

PH22PCO PH2O

+ K1K2PH2PCO +

K4PH2

Rj) i)1 j)1


(

(A8)

j)1

Rj / 1 +

K4PH2 + K1PCO + K 3

PH22PCO PH2O

+ K1K2PH2PCO +

K4PH2

Rj) i)1 j)1


(
n

]
i

(n g 2) (A9)

j)1

Rj / 1 +

K4PH2 + K1PCO + K 3

PH22PCO PH2O

+ K1K2PH2PCO +

K4PH2

Rj) i)1 j)1


(

(n g 2) (A10)

k5K 3 Rn )

PH22PCO [s1] PH2O

PH22PCO k5K [s1] + k6PH2 + k7+(1 - n) 3 PH2O

(n g 2)

(A11)

Ind. Eng. Chem. Res., Vol. 42, No. 21, 2003 5081

k5K 3 R1 )

PH22PCO [s1] PH2O

PH22PCO k5K [s1] + k6 3 PH2O PH22PCO [s1] k5K 3 PH2O

(A12)

RA )

n ) (k7 /k7 ) PCnH2n/ R1RA

Model FT III: RDS, steps 5, 7, and 8.

{ [
PH2O

PH22PCO k5K [s1] + k6PH2 + k7+ 3 PH2O k7-

(A13)

n-1

+ k5K 3 PCOPH2 PH2O


2

(RAi-2PC i)2

(n-i+2)H2(n-i+2)

[s1] + k6PH2 + k7+

]}
]
2

(n g 2)

(A14)

PH23PCO RCH4 ) k7MK4K6K / 1+ 3 PH2O

K4PH2 + K1PCO + K 3

PH22PCO PH2O

+ K1K2PCOPH2 + K6K40.5K 3

PH22.5PCO PH2O
i

RCnH2n+2 ) k7K4K6K 3

PH23PCO

Rj / j)2

PH22PCO n (1 + K6 K4PH2) K 3 PH2O i)2 1+ K4PH2 + K1PCO + K 3 PH22PCO PH2O

(Rj) j)2

(A15)

+ K1K2PCOPH2 + K6K40.5K 3

PH22.5PCO PH2O

RCnH2n ) [k8 (1 - n)K PH2 PCO/PH2O 3

Rj]/ 1 + j)2
0.5

PH22PCO n (1 + K6 K4PH2) K 3 PH2O i)2 K4PH2 + K1PCO + K 3 PH22PCO PH2O

(Rj) j)2

(n g 2) (A16)

+ K1K2PCOPH2 +

K6K4 K 3

PH22.5PCO PH2O k5K 3

PH22PCO n + K (1 + K6 K4PH2) 3 PH2O i)2

(Rj) j)2

(n g 2) (A17)

PH22PCO PH2O (n g 2) (A18)

Rn ) k5K 3

PH22PCO PH2O
2

+ k7K6K4PH2 + k8+(1 - n)/[s1] k8k5K PCOPH22/PH2O[s1] + k7K6K4PH2[s1] + k8+ 3

n ) (k8 /k8 ) PCnH2n/ RA

{ [

n-1

K PCOPH2 /PH2O + 3

(RAi-2PC i)2
RA ) k5K PH22PCO/PH2O[s1] 3 k5K PH22PCO/PH2O[s1] + k7K6K4PH2[s1] + k8+ 3

(n-i+2)H2(n-i+2)

]}

(n g 2) (A19)

(A20)

5082 Ind. Eng. Chem. Res., Vol. 42, No. 21, 2003

Model FT IV: RDS, steps 7-9.

RCH4 ) k8MK0PH2R1/(1 +

K0PH2 + K1PCO(1 + K2 K0PH2) + PH2O/K4 + K4K3K2K1K0PCOPH2/PH2O(1 + K5 K0PH2 + K6K5K0PH2) + K0PH2

(Rj))2 i)1 j)1

(A21)

RCnH2n+2 ) k8K0PH2

Rj /(1 + j)1

K0PH2 + K1PCO(1 + K2 K0PH2) + PH2O/K4 + K0PH2

K4K3K2K1K0PCOPH2/PH2O(1 + K5 K0PH2 + K6K5K0PH2) + RCnH2n ) k9+(1 - n) K0PH2

Rj))2 i)1 j)1


(

(n g 2) (A22)

Rj /[1 + j)1

K0PH2 + K1PCO(1 + K2 K0PH2) + PH2O/K4 + K0PH2

K4K3K2K1K0PCOPH2/PH2O(1 + K5 K0PH2 + K6K5K0PH2) + k7K PH22PCO/PH2O 3 k7K PH22PCO/PH2O + k8 K0PH2 + k9+(1 - n)/[s1] 3 R1 ) k7K PH22PCO/PH2O 3 k7K PH22PCO/PH2O + k8 K1PH2 3 k9k7K PH2 PCO/PH2O[s1] + k8 K0PH2[s1] + k9 3 k7K PH22PCO/PH2O[s1] 3
2

(Rj)] i)1 j)1

(n g 2) (A23)

Rn )

(n g 2)

(A24)

(n ) 1)

(A25)

n ) (k9 /k9 ) PCnH2n/ R1RA

{ [

n-1

+ i)2

(RAi-2PC

(n-i+2)H2(n-i+2)

]}
]

(n g 2) (A26) (A27)

RA )

k7K PH22PCO/PH2O[s1] + k8 K0PH2[s1] + k9+ 3

Model FT V(I): RDS, steps 1, 5, and 6.

RCH4 ) k5MPH2 K0PH2R1/ 1 +

K0PH2 + 1 +

PH2O

K2K3K4 P 2 H2
PH2O 1

K3K4 PH2

K4

K0PH2

Rj) i)1 j)1


(

(A28)

RCnH2n+2 ) k5PH2 K0PH2

j)1

Rj / 1 +

K0PH2 + 1 +

RCnH2n ) k6+(1 - n) K4PH2

j)1

Rj / 1 +

1 K4

K2K3K4 P 2 H2

K0PH2 + 1 +

K3K4 PH2

K0PH2

R )
(
j i)1 j)1

PH2O

K2K3K4 P 2 H2

K3K4 PH2

K4

]
i

(n g 2) (A29)

K0PH2

Rj) i)1 j)1


(

(n g 2) (A30) (A31)

Rn )

k1PCO k1PCO + k5PH2 + k6+(1 - n) R1 ) k1PCO k1PCO + k5PH2 k1PCO k1PCO + k5PH2 + k6+ (n ) 1)

(n g 2)

(A32)

RA )

(A33)

n ) (k6 /k6 ) PCnH2n / R1RA

{ [
[ [

Ind. Eng. Chem. Res., Vol. 42, No. 21, 2003 5083

n-1

k6k1PCO + k5PH2 + k6
+

(RAi-2PC i)2
n i

(n-i+2)H2(n-i+2)

]}

(n g 2)

(A34)

Model FT V(II): RDS, steps 2, 5, and 6.

RCH4 ) k5MPH2 K0PH2R1 / 1 +

K0PH2 + 1 +

1 K3K4PH2

1 K4

K0PH2

Rj) + i)1 j)1


(
n i

K1PCO K0PH2(1 + RCnH2n+2 ) k5PH2 K0PH2

j)1

Rj / 1 +

K0PH2 + 1 +

1 K3K4PH2

K4

Rj)) i)1 j)1


(

(A35)

K0PH2

Rj) + i)1 j)1


(
n i

K1PCO K0PH2(1 + RCnH2n ) k6+(1 - n) K0PH2

j)1

Rj / 1 +

K0PH2 + 1 +

1 K3K4PH2

1 K4

Rj)) i)1 j)1


(

] ]

(n g 2) (A36)

K0PH2

Rj) + i)1 j)1


(
n i

K1PCO K0PH2(1 + Rn ) k2PCOPH2 k2PCOPH2 + k5PH2 + k6+(1 - n) R1 ) k2PCO k2PCO + k5 (n ) 1) (n g 2)

Rj)) i)1 j)1


(

(n g 2) (A37)

(A38)

(A39)

RA )

k2PCOPH2 k2PCOPH2 + k5PH2 + k6+ k6-

(A40)

n ) (k6 /k6 ) PCnH2n/ R1RA

{ [
[ [

n-1

k2PCOPH2 + k5PH2 + k6

(RAi-2PC + i)2

(n-i+2)H2(n-i+2)

]}
)
(1 +

(n g 2)

(A41)

Model FT V(III): RDS, steps 3, 5, and 6.

RCH4 ) k5MPH2 K0PH2R1/ 1 +

K0PH2 + 1 +

( )
1 K4

K0PH2

Rj) + i)1 j)1


( K0PH2 K1PCO +

K1K2PCOPH2 PH2O

(Rj)) i)1 j)1

RCnH2n+2 ) k5PH2 K0PH2

j)1

Rj / 1 +

K0PH2 + 1 +

( )
1 K4

(A42)

K0PH2

K0PH2 K1PCO +
n

RCnH2n ) k6 (1 - n) K0PH2

Rj / j)1

( (

Rj) + i)1 j)1


( PH2O

K1K2PCOPH2

) )

(1 +

Rj)) i)1 j)1


(

1+

K0PH2 + 1 +

( )
1 K4

] ]

(n g 2) (A43)

K0PH2

(Rj) + i)1 j)1


PH2O

K0PH2 K1PCO +

K1K2PCOPH2

(1 +

(Rj)) i)1 j)1

(n g 2) (A44)

5084 Ind. Eng. Chem. Res., Vol. 42, No. 21, 2003

Rn )

k3PCOPH22/PH2O k3PCOPH22/PH2O + k5PH2 + k6+(1 - n) R1 ) k3PCOPH22/PH2O k3PCOPH22/PH2O + k5M k3PCOPH22/PH2O k3PCOPH22/PH2O + k5PH2 + k6+ k6k3PCOPH2 /PH2O + k5PH2 + k6
2 +

(n g 2)

(A45)

(n ) 1)

(A46)

RA )

(A47)

n ) (k6-/k6+) PCnH2n/ R1RAn-1 +

{ [

(RAi-2PC i)2

(n-i+2)H2(n-i+2)

]}
n

(n g 2) (A48)

Model FT VI: RDS, steps 6-8.

RCH4 ) k7MK50.5PH21.5R1/ 1 +

K5PH2 + K1PCO + K4-1

PH2O PH2

+ K1K2K4

PH2PCO PH2O

+ K 3

PH22PCO PH2O

+
i

K5PH2

(Rj) i)1 j)1


+

RCnH2n+2 ) k7K5 PH2

0.5

1.5

Rj / j)1

(A49)

1+

K5PH2 + K1PCO + K4

-1

PH2O PH2

+ K1K2K4

PH2PCO PH2O

+ K 3

PH22PCO PH2O

K5PH2

Rj) i)1 j)1


( + K 3
i

RCnH2n ) k8 (1 - n) K5PH2

Rj / 1 + j)1

(n g 2) (A50)

K5PH2 + K1PCO + K4

-1

PH2O PH2

+ K1K2K4

PH2PCO PH2O
n

PH22PCO

K5PH2 k6K 3 Rn ) PH22PCO [s1] PH2O

Rj) i)1 j)1


(

PH2O (n g 2) (A51)

PH22PCO k6K [s1] + k7PH2 + k8+(1 - n) 3 PH2O k6K 3 R1 ) PH2PCO [s ] PH2O 1

(n g 2)

(A52)

PH2PCO k6K [s ] + k7 3 PH2O 1

(n ) 1)

(A53)

RA )

k6K PH22PCO/PH2O[s1] 3 k6K PH22PCO/PH2O[s1] + k7PH2 + k8+ 3 k8-

(A54)

n ) (k8 /k8 ) PCnH2n/ R1RA

{ [

n-1

+
2

k6K PCOPH2 /PH2O[s1] + k7PH2 + k8 3

(RAi-2PC i)2

(n-i+2)H2(n-i+2)

]}

(n g 2)

(A55)

Ind. Eng. Chem. Res., Vol. 42, No. 21, 2003 5085

Model FT VII(I): RDS, steps 1, 7, and 8.

RCH4 ) k7MK0PH2R1/ 1 +

K0PH2 +

PH2O K6 K0PH2

RCnH2n+2 ) k7K0PH2

Rj / 1 + j)1
PH2O

PH2O

1 K3K4K5K0PH2

1 K4K5 K0PH2

1 K5

K2K3K4K5K02PH22 K0PH2 +

+1

K0PH2

(Rj) i)1 j)1

(A56)

PH 2 O K6 K0PH2

1 K3K4K5K0PH2

1 K4K5 K0PH2

1 K5

K2K3K4K5K02PH22

+1

RCnH2n ) k8+(1 - n) K0PH2

j)1

Rj / 1 +

PH2O

) )

K0PH2

(Rj) i)1 j)1

(n g 2) (A57)

K0PH2 +

PH 2 O K6 K0PH2

1 K3K4K5K0PH2

1 K4K5 K0PH2

1 K5

K2K3K4K5K02PH22 Rn )

+1

K0PH2

(Rj) i)1 j)1

(n g 2) (A58)

k1PCO k1PCO + k7 K0PH2[s1] + k8+(1 - n) R1 ) k1PCO k1PCO + k7 K0PH2[s1] k1PCO k1PCO + k7 K0PH2[s1] + k8+ k8k1PCO + k7 K0PH2[s1] + k8

(n g 2)

(A59)

(A60)

RA )

(A61)

n ) (k8 /k8 ) PCnH2n/ R1RA

{ [

n-1

(RAi-2PC + i)2

(n-i+2)H2(n-i+2)

]}
n

(n g 2)

(A62)

Model FT VII(II): RDS, steps 2, 7, and 8.

RCH4 ) k7MK0PH2R1/ 1 +

K0PH2 +

PH2O K6 K0PH2

K1PCO +

1 K3K4K5K0PH2

1 K4K5 K0PH2

1 K5

+1

K0PH2

(Rj) i)1 j)1

RCnH2n+2 ) k7K0PH2

j)1

Rj / 1 +

(A63)

K0PH2 +

PH 2 O K6 K0PH2 1

K1PCO +

1 K4K5 K0PH2

1 K5

K3K4K5K0PH2

+1

K0PH2

(Rj) i)1 j)1

(n g 2) (A64)

5086 Ind. Eng. Chem. Res., Vol. 42, No. 21, 2003

RCnH2n ) k8+(1 - n) K0PH2

j)1

Rj / 1 +

K0PH2 + 1

PH2O K6 K0PH2 +

K1PCO +

1 K4K5 K0PH2

1 K5

K3K4K5K0PH2

+1

K0PH2

Rj) i)1 j)1


(

(n g 2) (A65)

Rn )

k2K1 K0PH2PCO[s1] k2K1 K0PH2PCO[s1] + k7 K0PH2[s1] + k8+(1 - n) R1 ) k2K1 K0PH2PCO k2K1 K0PH2PCO + k7 K0PH2 k2K1 K0PH2PCO[s1] k2K1 K0PH2PCO[s1] + k7 K0PH2[s1] + k8+ k8k2K1 K0PH2PCO[s1] + k7 K0PH2[s1] + k8 (n ) 1)

(n g 2)

(A66)

(A67)

RA )

(A68)

n ) (k8 /k8 ) PCnH2n / R1RA

{ [

n-1

(RAi-2PC + i)2

(n-i+2)H2(n-i+2)

]}
+
i

(n g 2) (A69)

Model FT VIII: RDS, steps 5, 9, and 10.

RCH4 ) k9MK6PH2R1/ 1 +

K6PH2 + K1PCO + K2

PH2PCO PH2O

PH2O

PH 2 O K8 K6PH2

K6K7K8 PH2

+ K 3

PH21.5PCO PH2O

K 4

PH22PCO PH2O

K6PH2

(Rj) i)1 j)1

RCnH2n+2 ) k9K6PH2

Rj / 1 + j)1

(A70)

K6PH2 + K1PCO + K2

PH2PCO PH2O

PH2O

PH 2 O K8 K6PH2
n i

K6K7K8 PH2 K 4 PH22PCO PH2O 1 +

+ K 3

PH21.5PCO PH2O

K6PH2

(Rj) i)1 j)1


+

RCnH2n ) k10+(1 - n) K6PH2

j)1

Rj / 1 +

(n g 2) (A71)

K6PH2 + K1PCO + K21 PH21.5PCO PH2O

PH2PCO PH2O

PH2O

PH 2 O K8 K6PH2

K6K7K8 PH2
n

K 3

+ K 4

PH22PCO PH2O

K6PH2

Rj) i)1 j)1


(

(n g 2) (A72)

Rn )

k5K PH22PCO/PH2O 4 k5K PH22PCO/PH2O + k9 K6PH2 + k10+(1 - n)/[s1] 4 k5K PH21.5PCO/PH2O 4 k5K PH21.5PCO/PH2O + k9 K6 4

(n g 2)

(A73)

R1 )

(n ) 1)

(A74)

Ind. Eng. Chem. Res., Vol. 42, No. 21, 2003 5087

RA )

k5K PH22PCO/PH2O[s1] 4 k5K PH22PCO/PH2O[s1] + k9 K6PH2[s1] + k10+ 4 k10-

(A75)

n ) (k10 /k10 ) PCnH2n/ R1RA

{ [

n-1

+
2

k5K PH2 PCO/PH2O[s1] + k9 K6PH2[s1] + k10 4

+ i)2

(RAi-2PC

(n-i+2)H2(n-i+2)

]}

(n g 2) (A76)

Model FT IX: RDS, steps 5, 10, and 11.

RCH4 ) k10K9K6 K 4

0.5

PH22.5PCO PH2O

/ 1+

K6PH2 + K1PCO + K2

PH2PCO PH2O

+ K2

PH2O PH2

PH 2 O K8 K6PH2

+ K 3

PH21.5PCO PH2O
n i

K 4

PH22PCO PH2O

RCnH2n+2 ) k10K9K6K PH23PCO/PH2O 4

i)2

Ri / 1 +

K 3

PH21.5PCO PH2O
n

+ (1 + K9 K6PH2)K 4 PH2PCO PH2O PH2O P H2


n

PH22PCO PH2O +

Rj i)2 j)2
+

(A77)

K6PH2 + K1PCO + K21 PH22PCO PH2O

+ K22

PH 2 O K8 K6PH2

+ K 4

+ (1 + K9 K6PH2)K 4

PH22PCO PH2O + K2
2

RCnH2n ) k11 (1 - n)K PH2 PCO/PH2O 4

Ri / i)2

K 3 n )
+

PH21.5PCO PH2O

Rj i)2 j)2
PH2O PH2
n i

(n g 2) (A78) PH 2 O

1+

K6PH2 + K1PCO + K2 PH22PCO PH2O

PH2PCO PH2O

K8 K6PH2

+ K 4

+ (1 + K9 K6PH2)K 4

PH22PCO PH2O

Rj i)2 j)2
n

(k11 /k11 ) PCnH2n/ K PH2 PCO/PH2ORA 4

{ [

(n g 2) (A79)

n-1

+
2

k11k5K PH2 PCO/PH2O[s1] + k10K9K6PH2[s1] + k11 4


+

RAi-2PC i)2

(n-i+2)H2(n-i+2)

(n g 2) (A80) Rn ) k5K PH22PCO/PH2O 4 k5K PH22PCO/PH2O + k10K9K6PH2 + k11+(1 - n)/[s1] 4 RA ) k5K PH22PCO/PH2O[s1] 4 k5K PH22PCO/PH2O[s1] + k10K9K6PH2[s1] + k11+ 4 (n g 2) (A81)

]}

(A82)

where K ) K1K2K3K4K62K7K8. 4 Model FT X: RDS, steps 4, 8, and 9.

RCH4 ) k8MK7K5 K PH2 PCO/PH2O/ 1 + 3

0.5

2.5

K5PH2 + K1PCO + K2 PH22PCO PH2O

PH2PCO PH2O

+ K2

PH2O P H2

+ K7K K5 3

0.5

PH22.5PCO PH2O
n i

K 3

+ (1 + K7 K5PH2)K PH22PCO/PH2O 3

Rj i)2 j)2

(A83)

5088 Ind. Eng. Chem. Res., Vol. 42, No. 21, 2003

RCnH2n+2 ) k8K7K50.5K PH23.5PCO/PH2O 3


0.5

i)2

Ri/ 1 +

K7K K5 3

PH22.5PCO PH2O
n

K5PH2 + K1PCO + K21

PH2PCO PH2O

+ K22

PH2O PH2
n

+
i

+ K 3

PH22PCO PH2O

+ (1 + K7 K5PH2)K PH22PCO/PH2O 3 PH2PCO PH2O

RCnH2n ) k9+(1 - n)K PH22PCO/PH2O 3 K7K K50.5 3

Ri/ 1 + i)2
+ K 3

PH22.5PCO PH2O

Rj i)2 j)2
+
i

] ]

(n g 2) (A84)

K5PH2 + K1PCO + K21

+ K22

PH2O PH2
n

PH22PCO PH2O

+ (1 + K7 K5PH2)K PH22PCO/PH2O 3

Rj i)2 j)2

(n g 2) (A85)

Rn )

k4K PH22PCO/PH2O 3 k4K PH22PCO/PH2O + k8K7K5PH2 + k9+(1 - n)/[s1] 3 k4K 3 RA ) PH22PCO [s1] PH2O

(n g 2)

(A86)

PH22PCO k4K [s1] + k8K7K5PH2[s1] + k9+ 3 PH2O

(A87)

n ) (k9-/k9+) PCnH2n/ K 3 PH22PCO PH2O RA


n-1

{[

[s1] +

k9PH22PCO k4K [s1] + k8K7K5PH2[s1] + k9+ 3 PH2O

[s1]

RAi-2PC i)2

(n-i+2)H2(n-i+2)

]}

(n g 2)

(A88)

Nomenclature
E5 ) activation energy for chain growth (kJ mol-1) E7M ) activation energy for methane formation (kJ mol-1) E7 ) activation energy for paraffin formation (kJ mol-1) E8 ) activation energy for olefin formation (kJ mol-1) Ev ) activation energy for the WGS reaction (kJ mol-1) fi,obj ) objective function for the ith response FTot,obj ) multiresponse objective function kWGS4 ) rate constant of the forward reaction in step IV (mol g-1 s-1) k-WGS4 ) rate constant of the reversible reaction in step IV (mol g-1 s-1) k5 ) rate constant of chain growth (mol g-1 s-1 bar-1) k5,0 ) preexponential factor of chain growth (mol g-1 s-1 bar-1) k7M ) rate constant of methane formation (mol g-1 s-1 bar-1) k7M,0 ) preexponential factor of methane formation (n g 2) (mol g-1 s-1 bar-1) k7 ) rate constant of paraffin formation (mol g-1 s-1 bar-1) k7,0 ) preexponential factor of paraffin formation (n g 2) (mol g-1 s-1 bar-1) k8 ) rate constant of olefin formation (mol g-1 s-1) k8,0 ) preexponential factor of olefin formation (n g 2) (mol g-1 s-1) kv ) rate constant of CO2 formation (mol g-1 s-1 bar-1.5) kv,0 ) preexponential factor of CO2 formation (mol g-1 s-1 bar-1.5) k-8 ) rate constant of olefin readsorption reaction (mol g-1 s-1 bar-1)

K1 ) equilibrium constant of the elementary reaction 1 for FTS (bar-1) K2 ) equilibrium constant of the elementary reaction 2 for FTS (bar-1) K3 ) equilibrium constant the elementary reaction 3 for FTS K4 ) equilibrium constant the elementary reaction 4 for FTS (bar-1) K6 ) equilibrium constant the elementary reaction 6 for FTS Kp ) equilibrium constant of the WGS reaction Kv ) group of constants of the WGS reaction (bar-0.5) KWGS1 ) equilibrium constant of CO adsorption elementary step KWGS2 ) equilibrium constant of H2O adsorption elementary step KWGS3 ) equilibrium constant of surface reaction elementary step KWGS4 ) equilibrium constant of CO2 desorption elementary step KWGS5 ) equilibrium constant of H2 desorption elementary reaction mi ) mole flow rate of component i (mol s-1) MARR ) mean absolute relative residuals N ) maximum carbon number of the hydrocarbons involved Nc ) total number of components involved NR ) total number of reactions involved Nexp ) total number of experiments Nresp ) total number of responses for parameter estimization

Ind. Eng. Chem. Res., Vol. 42, No. 21, 2003 5089 Pi ) partial pressure of component i (bar) PT ) total pressure of the reaction system (bar) R ) gas constant (J mol-1 K-1) Rj ) overall reaction rate of reaction path j (mol g-1 s-1) Ri ) rate of formation of component i (mol g-1 s-1) RR ) relative residual between the experimental and calculated values of the response s1 ) active site for hydrocarbon formation s2 ) active site for the WGS reaction T ) temperature of the reaction system (K) Wi ) weight of the ith response W ) weight of the catalyst used (g) Greek Symbols R1 ) chain growth factor for carbon number 1 Rn ) chain growth factor for carbon number n (n g 2) RA ) chain growth factor without olefin readsorption Ri,j ) stoichiometric coefficient for the ith component in the jth reaction n ) readsorption factor of 1-olefin with carbon number n (n g 2) Superscripts and Subscripts cal ) calculated value exp ) experimental value i ) index indicating reactions j ) index indicating components M ) methane n ) number of carbon atoms * ) equilibrium state
(12) Lox, E. S.; Froment, G. F. Kinetics of the Fischer-Tropsch reaction on a precipitated promoted iron catalyst. 2. kinetic modeling. Ind. Eng. Chem. Res. 1993, 32, 71. (13) Wang, Y.-N.; Ma, W.-P.; Lu, Y.-J.; Yang, J.; Xu, Y.-Y.; Xiang, H.-W.; Li, Y.-W.; Zhao, Y.-L.; Zhang, B.-J. Kinetics modeling of Fischer-Tropsch synthesis over an industrial FeCu-K catalyst. Fuel 2003, 82, 195. (14) Ponec, V. In Handbook of Heterogeneous Catalysis; Knoezinger, G., Ertl, H., Weitkamp, J., Eds.; VCH: Weinheim, Germany, Vol. 4, pp 1879-1883. (15) Herington, E. F. G. The Fischer-Tropsch synthesis considered as a polymerization reaction. Chem. Ind. (London) 1946, 65, 346. (16) Friedel, R. A.; Anderson, R. B. Composition of Synthetic Liquid Fuels, I. Product Distribution and Analysis of C5-C8 Paraffin Isomers from Cobalt Catalyst. J. Am. Chem. Soc. 1950, 72, 1212. (17) Huff, G. A.; Satterfield, C. N. Liquid accumulation in a Fischer-Tropsch fixed bed reactor. Ind. Eng. Chem., Process Des. Dev. 1985, 24, 986. (18) Wojciechowski, B. W. The kinetics of the Fischer-Tropsch synthesis. Catal. Rev.-Sci. Eng. 1988, 30, 629. (19) (a) Novak, S.; Madon, R. J.; Suhl, H. Models of hydrocarbon product distributions in Fischer-Tropsch synthesis. J. Chem. Phys. 1981, 74 (11), 6083. (b) Novak, S.; Madon, R. J.; Suhl, H. Secondary effects in the Fischer-Tropsch synthesis. J. Catal. 1982, 77, 141. (20) Iglesia, E.; Soled, S. L.; Baumgartner, J. E.; Reyes, S. C. Synthesis and catalytic properties of eggshell cobalt catalysts for the Fischer-Tropsch synthesis. J. Catal. 1995, 153, 108. (21) Pichler, H.; Schultz, H.; Elstner, M. Gesetzmassigheiten bei der synthese von kohlenwasserstoffen aus kohlenoxide und wasserstoff (Some laws of the synthesis of hydrocarbons from carbon oxide and hydrogen). Brennst. Chem. 1967, 48, 78. (22) Satterfield, C. N.; Huff, G. A. Carbon number distribution of Fischer-Tropsch products formed on an iron catalyst in a slurry reactor. J. Catal. 1982, 73, 187. (23) Konig, L.; Gaube, J. Fischer-Tropsch-synthesis, recent investigations and developments. Chem.-Ing.-Tech. 1983, 55, 14. (24) Satterfield, C. N.; Huff, G. A.; Longwell, J. P. Product distribution from iron catalysts in Fischer-Tropsch slurry reactors. Ind. Eng. Chem., Process Des. Dev. 1982, 21 (3), 465. (25) Egiebor, N. O.; Cooper, W. C.; Wojciechowski, B. W. Carbon number distribution of Fischer-Tropsch CO hydrogenation products from precipitated iron catalysts. Can. J. Chem. Eng. 1985, 63, 826. (26) Huff, G. A.; Satterfield, C. N. Evidence for two chain growth probabilities on iron catalysts in Fischer-Tropsch synthesis. J. Catal. 1984, 85, 370. (27) Jordan, D. S.; Bell, A. T. Catalytic reactions of carbon monoxide. J. Phys. Chem. 1986, 90, 4797. (28) Iglesia, E.; Reyes, S. C.; Madon, R. J. Selectivity control and catalyst design in the Fischer-Tropsch synthesis. sites, pellets, and reactors. In Advances in catalysis and related subjects; Eley, D. D., Weisz, P. B., Pines, H., Eds.; Academic Press: New York, 1993; Vol. 39, p 221. (29) Kuipers, E. W.; Vinkenburg, I. H.; Oosterbeek, H. Chain length dependence of R-olefin readsorption in Fischer-Tropsch synthesis. J. Catal. 1995, 152 (1), 137. (30) Iglesia, E.; Reyes, S. C.; Madon, R. J. Transport enhanced R-olefin readsorption pathways in Ru-catalyzed hydrocarbon synthesis. J. Catal. 1991, 129, 238. (31) Zimmerman, W. H.; Bukur, D.; Ledakowicz, S. Kinetic Model of Fischer-Tropsch Synthesis Selectivity in the Slurry Phase. Chem. Eng. Sci. 1992, 47, 2707. (32) Kuipers, E. W.; Scheper, C.; Wilson, J. H.; Vinkenburg, I. H.; Oosterbeek, H. Non-ASF Product Distributions Due to Secondary Reactions during Fischer-Tropsch Synthesis. J. Catal. 1996, 158 (1), 288. (33) Schulz, H.; Claeys, M. Kinetic modelling of FischerTropsch product distribution. Appl. Catal. A 1999, 186 (1-2), 91. (34) Overett, M. J.; Hill, R. O.; Moss, J. R. Organometallic chemistry and surface science: mechanistic models for the Fischer-Tropsch synthesis, Coord. Chem. Rev. 2000, 206-207, 581. (35) Ponec, V. Some aspects of the mechanism of methanation and Fischer-Tropsch synthesis. Catal. Rev.-Sci. Eng. 1978, 18 (1), 151.

Literature Cited
(1) Schulz, H. Short history and present trends of FischerTropsch synthesis. Appl. Catal. A 1999, 186, 3. (2) (a) Dry, M. E. The Fischer-Tropsch process: 1950-2000. Catal. Today 2002, 71, 227. (b) Dry, M. E. Fischer-Tropsch reactions and the environment. Appl. Catal. A 1999, 189 (2), 185. (3) Dry, M. E. High quality diesel via the Fischer-Tropsch processsa review. J. Chem. Technol. Biotechnol. 2001, 77, 43. (4) (a) van der Laan, G. P.; Beenackers, A. A. C. M. Intrinsic kinetics of the gas-solid Fischer-Tropsch and water gas shift reactions over a precipitated iron catalyst. Appl. Catal. A 2000, 193, 39. (b) van der Laan, G. P. Kinetics, selectivity and scale up of the Fischer-Tropsch synthesis. Ph.D. Dissertation, University of Groningen, Groningen, Germany, 1999. (5) Liu, Z.-T.; Li, Y.-W.; Zhou, J.-L.; Zhang, B.-J. Intrinsic kinetics of Fischer-Tropsch synthesis over an Fe-Cu-K catalyst. J. Chem. Soc., Faraday Trans. 1995, 91 (18), 3255. (6) (a) Zimmerman, W. H.; Bukur, D. B. Reaction kinetics over iron catalysts used for Fischer-Tropsch synthesis. Can. J. Chem. Eng. 1990, 68, 292. (b) Zimmerman, W. H. Kinetic modeling of the Fischer-Tropsch synthesis. Ph.D. Thesis, TAMU, 1990. (7) Anderson, R. B. Catalysis for the Fischer-Tropsch synthesis; Van Nostrand Reinhold: New York, 1956; Vol. 4, p 29. (8) Dry, M. E. Advances in Fischer-Tropsch chemistry. Ind. Eng. Chem., Prod. Res. Dev. 1976, 15, 282. (9) (a) Sarup, B.; Wojciechowski, B. W. Studies of the FischerTropsch Synthesis on a Cobalt Catalyst. I. Evaluation of Product Distribution Parameters from Experimental Data. Can. J. Chem. Eng. 1988, 66, 831. (b) Sarup, B.; Wojciechowski, B. W. Studies of the Fischer-Tropsch Synthesis on a Cobalt Catalyst. II. Kinetics of Carbon Monoxide Conversion to Methane and to Higher Hydrocarbons. Can. J. Chem. Eng. 1989, 67, 62. (c) Sarup, B.; Wojciechowski, B. W. Studies of the Fischer-Tropsch Synthesis on a Cobalt Catalyst. III. Mechanistic Formulation of the Kinetics of Selectivity for Higher Hydrocarbon Formation. Can. J. Chem. Eng. 1989, 67, 620. (10) van der Laan, G. P.; Beenackers, A. A. C. M. Kinetics and selectivity of the Fischer-Tropsch synthesis: A literature review. Catal. Rev.-Sci. Eng. 1999, 41 (3-4), 255 and the literature herein. (11) Lox, E. S.; Froment, G. F. Kinetic of the Fischer-Tropsch reaction on a precipitated promoted iron catalyst. 1. experimental procedure and results. Ind. Eng. Chem. Res. 1993, 32, 61.

5090 Ind. Eng. Chem. Res., Vol. 42, No. 21, 2003
(36) Brady, R.; Pettit, R. Reactions of diazomethane on transition-metal surfaces and their relationship to the mechanism of the Fischer-Tropsch reaction. J. Am. Chem. Soc. 1980, 102, 6181. (37) Martinez, J. M.; Adams, H.; Bailey, N. A.; Maitlis, P. M. The Coupling of Vinyl and -Methylene Ligands: A New View of the Mechanism of the Fischer-Tropsch Polymerisation Reaction. J. Chem. Soc., Chem. Commun. 1989, 5, 286. (38) Joyner, R. W. The mechanism of chain growth in the Fischer-Tropsch hydrocarbon synthesis. Catal. Lett. 1988, 1, 307. (39) (a) Bai, L.; Xiang, H.-W.; Li, Y.-W.; Han, Y.-Z.; Zhong, B. Slurry phase Fischer-Tropsch synthesis over manganese-promoted iron ultrafine particle catalyst. Fuel 2002, 81, 1577. (b) Zhong, B.; Wang, Q.; Peng, S. Y. Chinese Patent ZL95106156.9, 1995. (40) Itoh, H.; Nagano, S. Catalytic properties and crystalline structures of manganese-promoted iron ultrafine particles for liquid-phase hydrogenation of carbon monoxide. Appl. Catal. A 1993, 96, 125. (41) Li, X. G.; Zhong, B. Fischer-Trpoch synthesis on Fe-Mn ultrafine catalysts. Catal. Lett. 1994, 23, 245. (42) (a) Maiti, G. C.; Malessa, R.; Baerns, M. Iron/manganese oxide catalysts for Fischer-Tropsch synthesis. Part I: structural and textural changes by calcination, reduction, and synthesis. Appl. Catal. 1983, 5, 151. (b) Maiti, G. C.; Malessa, R.; Lochner, U.; Papp, H.; Baerns, M. Iron/manganese oxide catalysts for Fischer-Tropsch synthesis. Part II: crystal phase composition, activity and selectivity. Appl. Catal. 1985, 16, 215. (c) Lochner, U.; Papp, H.; Baerns, M. Iron/manganese oxide catalysts for Fischer-Tropsch synthesis. Part III: Phase changes in iron/ manganese oxide Fischer-Tropsch catalysts during start-up and synthesis process. Appl. Catal. 1986, 23, 339. (d) Malessa, R.; Baerns, M. Iron/manganese oxide catalysts for Fischer-Tropsch synthesis. 4: activity and selectivity. Ind. Eng. Chem. Res. 1988, 27, 279. (e) Grzybek, T.; Papp, H.; Baerns, M. Iron/manganese oxide catalysts for Fischer-Tropsch synthesis. Part V: XPS surface characterization of calcined and reduced samples. Appl. Catal. 1987, 29, 335. (f) Grzybek, T.; Papp, H.; Baerns, M. Iron/ manganese oxide catalysts for Fischer-Tropsch synthesis. Part VI: surface characterization during start-up period and pseudosteady-state synthesis. Appl. Catal. 1987, 29, 351. (43) (a) Burkur, D. B.; Lang, X.; Ding, Y. Pretreatment effect studies with a precipitated iron Fischer-Tropsch catalyst in a slurry reactor. Appl. Catal. A 1999, 186, 255. (b) Mansker, L. D.; Jin, Y.; Burkur, D. B.; Datye, A. K. Characterization of slurry phase iron catalysts for Fischer-Tropsch synthesis. Appl. Catal. A 1999, 186, 277. (44) (a) Jin, Y.; Datye, A. K. Phase transformations in iron Fischer-Tropsch catalysts during temperature programmed reduction. J. Catal. 2000, 196, 8. (b) Sault, A. G. An Auger electron spectroscopy study of the activation of iron Fischer-Tropsch catalysts. Part I: Hydrogen activation. J. Catal. 1993, 140, 121. (c) Sault, A. G.; Datye, A. K. An Auger electron spectroscopy study of the activation of iron Fischer Tropsch catalysts: Part II: Carbon monoxide activation. J. Catal. 1993, 140, 136. (45) Itoh, H.; Nagano, S.; Takeda, K.; Kikuchi, E. Catalytic properties and crystalline structures of manganese-promoted iron ultrafine particles for liquid-phase hydrogenation of carbon monoxide. Appl. Catal. A 1993, 186, 125. (46) (a) Liu, Z.-P.; Hu, P. A new insight into Fischer-Tropsch synthesis. J. Am. Chem. Soc. 2002, 124, 11568. (b) Ciobca, I. M. The molecular basis of the Fischer-Tropsch reaction. Ph.D. Thesis, Technical University Eindhoven, Eindhoven, The Netherlands, 2002. (47) Lox, E. S.; Marin, G. B.; de Graeve, E.; Bussier, P. Characterization of promoted precipitated iron catalyst for Fischer-Tropsch synthesis. Appl. Catal. A 1988, 40, 197. (48) Madon, R. J.; Taylor, W. F. Fischer-Tropsch synthesis on a precipitated iron catalyst. J. Catal. 1981, 69, 23. (49) Zhang, H. B.; Schrader, G. L. Characterization of a fused iron catalyst for Fischer-Tropsch synthesis by in-situ laser Raman spectroscopy. J. Catal. 1985, 95, 325. (50) Adesina, A. A. Hydrocarbon synthesis via Fischer-Tropsch reaction: travails and triumphs. Appl. Catal. A 1996, 138, 345. (51) Ponec, V.; Van Barneveld, W. A. The role of chemisorption in Fischer Tropsch synthesis. Ind. Eng. Chem., Process Des. Des. 1979, 4, 268. (52) Li, S.; Meitzner, G. D.; Iglesia, E. Structure and site evolution of iron oxide catalyst precursors during the FischerTropsch synthesis. J. Phys. Chem. B 2001, 105, 5743. (53) (a) Jiang, M.; Koizumi, N.; Yamada, M. Adsorption properties of iron-manganese catalysts investigated by in-situ diffuse reflectance FTIR spectroscopy. J. Phys. Chem. B 2000, 104, 7636. (b) Bian, G.; Oonuki, A.; Kobayashi, Y.; Koizumi, N.; Yamada, M. Syngas adsorption on precipitated iron catalysts reduced by H2, syngas, or CO and on those used for high-pressure FT synthesis by in-situ diffuse reflectance FTIR spectroscopy. Appl. Catal. A 2001, 219, 13. (54) Dry, M. E. Practical and theoretical aspects of the catalytic Fischer-Tropsch process. Appl. Catal. A 1996, 138, 319. (55) Huff, G. A.; Satterfield, C. N. Intrinsic kinetics of the Fischer-Tropsch synthesis on a reduced fused-magnetite catalyst. Ind. Eng. Chem., Process Des. Dev. 1984, 23, 696. (56) Bell, A. T. Catalytic synthesis of hydrocarbons over group VIII metals. A discussion of the reaction mechanism. Catal. Rev.Sci. Eng. 1981, 23 (1-2), 203. (57) Turner, M. L.; Marsih, N.; Mann, B. E.; Quyoum, R.; Long, H. C.; Maitlis, P. M. Investigations by 13C NMR spectroscopy of ethene-initiated catalytic CO hydrogenation. J. Am. Chem. Soc. 2002, 124, 10456. (58) Vannice, M. A. Synthesis of hydrocarbons from CO and H2. Catal. Rev.-Sci. Eng. 1976, 14 (2), 153. (59) Hanlon, R. T.; Satterfield, C. N. Reactions of selected 1-olefins and ethanol added during the Fischer-Tropsch synthesis. Energy Fuels 1988, 2, 196. (60) Newsome, D. S. The water-gas shift reaction. Catal. Rev.Sci. Eng. 1980, 21, 275. (61) Rethwisch, D. G.; Dumesic, J. A. Adsorptive and catalytic properties of supported metal oxides. III. Water gas shift over supported iron and zinc oxides. J. Catal. 1986, 101, 35. (62) Ovesen, C. V.; Stoltze, P.; Norskov, J. K.; Campbell, C. T. A Kinetic Model of the Water-Gas Shift Reaction. J. Catal. 1992, 134, 445. (63) Ovesen, C. V.; Clausen, B. S.; Hammershi, B.; Askgrd, T.; Chorkendorff, I.; Nrskov, J. K.; Rasmussen, P. B.; Stoltze, P.; Taylor, P. A. Kinetics of the water-gas shift reaction at high pressures. J. Catal. 1996, 158, 170. (64) Graaf, G. H.; Winkelman, J. G. M.; Stamhuis, E. J.; Beenackers, A. A. C. M.; Kinetics of the three phase methanol synthesis. Chem. Eng. Sci. 1988, 43, 2161. (65) Gear, G. W. Numerical initial value problem in ordinary differential equations; Prentice Hall: Englewood Cliffs, NJ, 1971. (66) Park, T. Y.; Froment, G. F. A hybrid genetic algorithm for the estimation of the parameters in detailed kinetics models. Comput. Chem. Eng. 1998, 22, S103. (67) Han, R. F.; Zhang, Y. K.; Wang, Y. N.; Xu, Y. Y.; LI, Y. W. Optimization of parameters of FTS kinetic model by genetic algorithm. J. Fuel Chem. Technol. (China) 2001, 29 (4), 371. (68) Ji, Y.-Y.; Xiang, H.-W.; Yang, J.-L.; Xu, Y.-Y.; Li, Y.-W.; Zhong, B. Effect of reaction conditions on the product distribution during Fischer-Tropsch synthesis over an industrial Fe-Mn catalyst. Appl. Catal. A 2001, 214 (1), 77. (69) Dictor, R. A.; Bell, A. T. Fischer-Tropsch synthesis over reduced and unreduced iron oxide catalysts. J. Catal. 1986, 97, 121. (70) Feimer, J. L.; Silveston, P. L.; Hudgins, R. R. Steady-state study of the Fischer-Tropsch reaction. Ind. Eng. Chem., Prod. Res. Dev. 1981, 20, 609.

Received for review February 14, 2003 Revised manuscript received June 12, 2003 Accepted June 13, 2003 IE030135O

Você também pode gostar