You are on page 1of 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, People’s Republic of China, and
Department of Chemistry, Jinan University, Guangzhou 510632, People’s 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
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

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
Downloaded via CHULALONGKORN UNIV on June 2, 2020 at 03:39:11 (UTC).

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 rate-
determining 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 transportation-
enhanced readsorption and secondary reaction of olefins more practically in the reactor modeling
stage.

1. Introduction The quality of a detailed kinetic model is closely


Fischer-Tropsch synthesis (FTS) is an industrially related to the understanding of the mechanism in the
important process for the conversion of syngas (H2/CO) FTS catalytic reaction system, in which a polymeriza-
derived from carbon sources such as coal, peat, biomass, tion process has been recognized to be dominant;
and natural gas into hydrocarbons and oxygenates. however, sufficient details on the FTS mechanistic
Today, it continuously attracts interest as an option for aspect are not yet fully understood.10,14,15 In 1946,
the production of clean transportation fuels and chemi- Herington first treated the molar distribution of hydro-
cal feedstocks.1-3 The FTS product is composed of a carbons from FTS in terms of a polymerization mech-
complex multicomponent mixture of linear and branched anism.15,16 The same formulation was rediscovered by
hydrocarbons and oxygenated products, the majority of Anderson et al. in 1951 and named the ASF distribu-
which are linear hydrocarbons.3,4 tion.7,16 In the ASF model, the formation of hydrocarbon
The FTS kinetics has extensively been studied, and chains was assumed as a stepwise polymerization pro-
many attempts have been made for the rate equations cedure and the chain growth probability was assumed
describing the FTS reactions.4-13 In most cases, the to be independent of the carbon number. However,
hydrocarbon products were lumped according to the significant deviations from the ideal ASF distribution
carbon number of hydrocarbon molecules with an ideal have been observed in many studies.17-20 Pichler et al.21
Anderson-Schulz-Flory (ASF) distribution. Although for the first time reported the deviations of experimental
a few of the available kinetic models were developed results from the ASF distribution. The usual deviations
based on the detailed mechanism of the Langmuir- of the distribution of the linear hydrocarbons are a
Hinshelwood-Hougen-Watson (LHHW) type,4,10-13 only relatively higher selectivity to methane, a relatively
two of them simultaneously considered both the syn- lower selectivity to ethane, and an increase in the
gas conversion rates and the hydrocarbon formation chain growth probability with increasing molecular size
rates.11-13 We have pointed out the deficiencies existing in comparison to the ideal ASF distribution. Several
in the conventional lumped models and those tailing different explanations about the cause of these devia-
syngas conversion rates with a carbon number distribu- tions have been proposed.22-24 Some authors25,26 inter-
tion formula when a comprehensive simulation is preted the deviations from the standard ASF distribu-
required.13 tion by the superposition of two ASF distributions. They
suspected the existence of two sorts of sites for the chain

Chinese Academy of Sciences. growth on the catalyst surface and, therefore, proposed

Jinan University. 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. the carbide mechanism for the FTS and the formate
However, this explanation cannot interpret the increase intermediate scheme for the WGS reaction. The rates
of the paraffin/olefin ratio with the chain length. On the of CO or H2 consumption and product formation were
basis of the experiments with cofed olefins, it was noted unified in these models. Experimental data were re-
that the readsorption and secondary reaction of olefins gressed with large-scale nonlinear optimization ap-
had a great influence on the products distribution of proaches for determining the “best” kinetic parameters,
FTS.27 Some researchers28-30 proposed a more plausible and statistical analysis was applied to validate the
explanation for these deviations and suggested that the feasibility of both models and parameters in them. The
occurrence of secondary reactions of the olefins caused final model could describe the distribution of linear
the deviations from the ASF distribution. Iglesia et al.30 paraffins and olefins of FTS obeying the ideal ASF
developed a model describing the olefin readsorption distribution at the level of surface reaction. However,
effect enhanced by intraparticle and interparticle trans- the deviations, observed in many experiments, from the
port processes. They suggested that the diffusion limita- ideal ASF distribution were totally neglected in their
tion within liquid-filled pores slowed the removal of model. Very recently, we have proposed a systematic
1-olefins, which caused an increase of their residence approach for considering more complicated olefin read-
time within the catalyst pores. In their model, the CO sorption phenomena in kinetic modeling on the basis
hydrogenation model and the olefin readsorption model of ideas similar to those of Froment.13 Although pri-
were treated separately. In 1992, Zimmerman and mary, namely, a limited number of mechanism sets,
Bukur31 developed a model for both the formation of from which only about 10 kinetic models were scanned,
linear hydrocarbons and the water-gas-shift (WGS) were considered, our optimal model for an industrial
reaction over iron catalysts. In their model, the read- iron catalyst showed a better description for the non-
sorption and secondary reaction of olefins were taken ideal ASF distribution than that from Lox and Froment.
into account and the increase of the paraffin/olefin ratio However, the models based on the detailed mechanism
with the carbon number can be predicted. However, the scan for FTS processes need to first describe the most
results displayed significant deviations between model intrinsic factors involved, namely, those defined at the
predicted and experimental mole fractions, especially surface mechanism level, and the solubility/diffusivity-
for methane and ethene. Kuipers et al.32 proposed a enhanced paraffin/olefin ratio dependence on the carbon
chain-length-dependent olefin readsorption mechanism number should be described through modification of the
to explain the fact that the paraffin/olefin ratio increases models with intrinsic significances. This modification
with the carbon number. On the basis of the olefin needs to be carefully arranged at both the surface
cofeeding kinetic experiments over cobalt catalyst in a mechanism and the reactor simulation levels because
continuously operated and well-mixed slurry reactor, the transportation-enhanced phenomena may not fun-
Schulz and Claeys33 developed a kinetic model for the damentally be reflected solely either in an “intrinsic”
Fischer-Tropsch reaction system. kinetic model or in reactor models using the intrinsic
It is well accepted that the FTS reaction is a surface kinetics correctly. This is because reactor models can
polymerization reaction, but its detailed mechanism is fundamentally account for transportation phenomena,
still not fully understood. Recent mechanistic studies and at the same time long-chain olefins staying in the
show new evidence in favor of a mechanism start with catalyst pores filled with wax are difficult to remove
CO dissociation. It is generally agreed that the FTS instantly as “defined” even in the case of kinetic
proceeds via the dissociation of CO, further forming experimental conditions. In fact, this is a dilemma for
carbide on the surface in sequence of the hydrogenation the kinetics modeling of FTS. The difficulties arise,
of this carbide.10,34,35 Despite considerable research making kinetic modeling impossible, if one wants to
efforts, uncertainties still remain about the mechanism fundamentally take the transportation phenomena into
of chain growth in the FTS reaction. Brady and Pettit36 account for the modeling of a kinetic reactor. On the
proposed an alkyl mechanism and suggested that the basis of this understanding, we first make an attempt
chain growth is initiated by the insertion of methylene to grasp the most significant intrinsic information at
into the adsorbed alkyl. Martinez et al.37 proposed a the mechanism kinetic level, and later efforts will be
chain growth mechanism, in which the chain growth is devoted to solving the modification problem for consid-
initiated by an R-vinyl and the polymerization process ering the transportation-enhanced phenomena with the
proceeds through the reaction between methylene and combination of kinetics and reactor simulations.
an alkenyl. Joyner38 developed an alkylidene mecha- The goal of this paper is, therefore, to establish and
nism and assumed the chain growth process propagated test detailed mechanistic kinetic models for the Fis-
via the successive insertion of methylene into the cher-Tropsch system over an industrial Mn-Fe cata-
metal-alkylidene bond. With this model, one can easily lyst while considering many more possibilities in mech-
explain the formation of R-olefins as the primary anism combinations than before. The new mechanistic
products in FTS. These proposed models have provided model, in which the olefin readsorption and secondary
a basis for a workable kinetic model derived from reactions are taken into account, combines all of the FTS
experimental data. However, models are still far from reactions with WGS reaction in a self-consistent way,
meeting the demands for both better mechanism un- leaving the transportation-enhanced olefin readsorption
derstanding and applications in engineering scale-up, factor for further work in the combination of kinetics
leaving a great deal of work under the expectation for and reactor simulations.
a better treatment in considering both the fundamental
mechanism and self-consistency in rate expressions.10,13 2. Experimental Section
Lox and Froment11,12 studied the FTS kinetics over
an iron catalyst. On the basis of the LHHW scheme, The experimental setup used in this study is shown
they developed several sets of “elementary steps”, from in Figure 1. A micro-fixed-bed reactor (a 1.0 m long
which detailed kinetic models were derived by assuming stainless steel tube of 0.012 m in internal diameter with
5068 Ind. Eng. Chem. Res., Vol. 42, No. 21, 2003

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 Ki-
netic 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
Figure 1. Experimental setup: (1) gas cylinders; (2) pressure on a 13× molecular sieve packed column (1.5 m × 3 mm
meters; (3) pressure regulators; (4) pressure regulators before the i.d., Ar carrier flow) and detected with a thermal
reactor; (5) mass flow controllers; (6) the reactor; (7) wet flowmeter; conductivity detector (TCD). C1-C5 hydrocarbons in the
(8) needle valves; (9) backpressure regulator; (10) purification
columns; (11) wax condenser; (12) oil/water condenser; (13) wax gas phase were analyzed on a C220/C-22 (170-250 µm)
trap (hot trap); (14) cold trap; (15) high-temperature ball valves; packed column (7.2 m × 3.2 mm i.d.) flame ionization
(16) ball valves; (17) salt bath. detector (FID) with N2 as the carrier. CO2 was measured
on a silica gel packed column (1 m × 3 mm i.d, H2
a temperature-controlled molten salt bath to achieve a carrier) with TCD and quantified by an external stan-
uniform temperature profile in the catalyst bed) was dard method. The oil product from the cold trap was
used for all kinetic experiments. The catalyst used in separated on a 60 m × 0.25 mm (i.d.) OV-101 capillary
the kinetic study is an industrial manganese-promoted column (N2 carrier, FID) with the temperature pro-
iron catalyst prepared by Institute of Coal Chemistry grammed from 333 K (maintained for 20 min) to 563 K
(ICC), Chinese Academy of Sciences. The preparation at the rate of 3 K min-1. The wax product from the hot
method has been patented, and some open information trap was first dissolved in CS2 (0.5-1.0 mass %) and
can be found in the literature.39,41 The fresh catalyst then analyzed on a OV-101 capillary column, FID, and
was crushed and sieved to particles with diameters of N2 carrier with the temperature programmed (1 K
0.25-0.36 mm (40-60 ASTM mesh), which has been min-1) from 343 to 563 K. The water phase was
proved to be a compromising particle size safe for separated by using a BD-wax (GW, US) capillary
neglectable intraparticle transfer limitations and prom- column (N2 carrier, FID). All of the data in steady-state
isingly easy operations to the reactor during the reactions showed promising material balances with
experiment.10-13 A 3.58 g (3 cm3) catalyst was used for carbon, hydrogen, and oxygen material balances be-
the kinetic experiments. The catalyst was diluted by 24 tween 95 and 104%.
cm3 inert silica sand [(catalyst:sand ) 1:8 (v/v)] with The experiences from our intensive experimental
the same mesh size range. Before the reaction, the investigations of the catalytic performances of various
catalyst was reduced in situ with syngas (H2/CO ) 2.0) FTS catalysts showed that better material balances are
in the reactor. difficult to reach in practice. This is partially due to the
fact that small fluctuations during a sampling time
In a typical reduction procedure, operation conditions
brought about certain experimental errors for the final
were adjusted to 0.25-0.30 MPa and 1000 h-1 at 513 results and partially due to the fact that the complexity
K, and then the temperature was raised to 548 K at of the FTS products caused difficulties in analyses. In
the rate of 1 K h-1 and retained at that temperature addition, detailed kinetic modeling requires the infor-
for the whole reduction stage. After reduction for 36 h, mation of the compositions of the reactants and almost
the temperature was lowered to 513 K and then the all FTS products (hydrocarbons relevant) in each stream.
operation conditions were adjusted to the desired value It is estimated that about 5% error in the total material
for the FTS. For intrinsic kinetic studies, a high space balances could produce very large errors (sometimes
velocity is needed to exclude the external diffusion several times) for the contents of components with very
limitation. In our work, the space velocity of (1.6-4.2) small mole fractions in a stream. This error amplifica-
× 10-3 Nm3 kg of catalyst-1 s-1 is much higher than tion phenomenon in FTS makes it even more difficult
those reported by others,5,13 ensuring the kinetic experi- to achieve high modeling accuracy in detailed kinetics.
ments with the neglected external diffusion effect.
The syngas was prepared by blending of pure CO and 3. Kinetic Models
H2 (>99.99% purity). The CO and H2 passed through a
series of an oxygen-removal trap, a heated silica gel 3.1. Active Site Assumption. The overall FTS
molecular sieve trap, and an activated charcoal trap to reactions can be simplified as the combination of FTS
remove tiny amounts of oxygen, water, and other reactions and the WGS reaction.11,12
impurities. The flow rates of CO and H2 were controlled
by two Brooks 5850E mass flow controllers. The outlet paraffin formation: nCO + (2n + 1)H2 )
of the reactor is connected with a hot trap (420 K) and CnH2n+2 + nH2O (1)
then an ice trap (273 K) at the system pressure. After
these product collectors, the pressure was released olefin formation: nCO + 2nH2 ) CnH2n + nH2O (2)
through a backpressure regulator. The flow rate of the
tail gas was monitored by a wet test gas meter. Gas 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 Fischer-Tropsch catalysis cycling. The precise defini-
not clear especially in the cases of iron catalysts. This tion of the catalysis cycle during FTS on iron catalysts
is because of the fact that the iron-based catalysts is, therefore, impossible with the current understanding
starting from their oxide precursors have experimen- of the FTS catalysis. Nevertheless, some conventional
tally proved to have complex phase transfer during the ideas/treatments may serve for the development of
reduction as well as synthesis operation.40,41-45 The kinetic models. We here take advantage of the conven-
controversy in the active sites (or phases) in the FTS tional idea that the building block “CH2” is formed by
system appeals the need for much attention to be paid the reaction of a surface carbon with dissociated
to the understanding of the FTS mechanism at a deep hydrogen4,10-13,50,54,56 to build the kinetic models. It is
level; for example, the work for Co or Ru catalysts, for thus assumed that surface carbon species undergoes a
which metal phases are believed to be active phases for reaction with surface hydrogen:
FTS, has been considered at even a molecular level with
many interesting results achieved, casting an encourag- Cs1 + Hs1 ) CHs1 (7)
ing light on the detailed mechanism understanding.46
For iron catalysts, however, we expect many more CHs1 + Hs1 ) CH2s1 (8)
challenges in this direction.41-49 Here, we will not go
into these controversies and instead will use experi- or with molecular hydrogen according to an Eley-Rideal
mental data to fit many possible mechanism-derived (ER) mechanism:
kinetic models, which are anyhow expected to reflect
the most important points of FTS catalysis and chem- Cs1 + H2 ) CH2s1 (9)
istry. Beyond the complexities in the FTS catalysis with
iron catalysts, it is generally accepted that reactions in Another possible pathway of the formation of “CH2”
eqs 1-3 can be assumed to take place on two kinds of starts with molecularly adsorbed carbon monoxide and
active sites. The active sites for the hydrocarbon forma- successive hydrogen-assisted dissociation,4,18,55
tion and the WGS reaction are iron carbides and
magnetite (Fe3O4), respectively.4,11-13,47-49 COs1 + Hs1 ) HCOs1 + s1 (10)
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 HCOs1 + Hs1 ) Cs1 + H2Os1 (11)
series reactions involving different extents and deter-
mining altogether the overall catalyst performance. The Cs1 + Hs1 ) CHs1 (12)
whole synthesis reaction can be simplified as the
combination of FTS reactions and the WGS reaction. CHs1 + Hs1 ) CH2s1 (13)
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) or assisted by molecular hydrogen via the ER mecha-
chain termination and desorption of products; (5) read- nism:
sorption and secondary reaction of olefins.
The CO adsorbed on the catalyst surface either in the COs1 + H2 ) H2COs1 (14)
molecule state or in the dissociated state18,51-53
H2COs1 + H2 ) CH2s1 + H2O (15)
CO + s1 ) COs1 (4)
There is still a controversy about the mechanism of
COs1 + s1 ) Cs1 + Os1 (5) chain growth in the FTS. The alkyl mechanism proposes
that the reaction is initiated by the formation of a
where s1 is an empty catalytic site, on which hydro- methyl species and that chain growth takes place by
carbon can be formed. The dissociative adsorption of the successive insertion of methylene into the metal-
hydrogen takes place on two adjacent free active alkyl bond:36
sites.51-53
CH2s1 + CnH2n+1s1 ) Cn+1H2n+3s1 + s1 (16)
H2 + 2s1 ) 2Hs1 (6)
The alkylidene mechanism proposes that the forma-
For the further formation of hydrogenated carbon tion of the adsorbed ethylidene initiates the chain
species, mechanism studies for FTS often assumed the formation and that chain growth is facilitated by meth-
formation of CH2 and CH3 species,4,10-13,55 while recent ylene insertion into the metal-alkylidene bond:38
characterization and theoretical studies in energetics
debate for the increased possible formation of CH CH2s1 + CnH2ns1 ) Cn+1H2n+2s1 + s1 (17)
species on several transition-metal surfaces.46,57 For iron
catalysts, the situation considered in theoretical models Another mechanism assumes that chain growth is
(or the model catalysts and operation conditions in initiated by the adsorption of CO on the active sites
characterization) describing the catalyst surfaces is already containing a hydrocarbon intermediate and then
definitely too far from a real working catalyst, which by a sequence of hydrogenation:56,58
continuously changes in both the bulk phase and
surfaces due to the carbonization of the iron and the CO + CnH2n+1s1 ) CnH2n+1s1CO (18)
hydrogenation of carbon species on the surface during
reduction and reactions.40-45,47,52 The carbides on the CnH2n+1s1CO + H2 ) CnH2n+1s1C + H2O (19)
catalyst surface may exchange carbon and hydrogen
sources with reactants and surface intermediates during CnH2n+1s1C + H2 ) CnH2n+1s1CH2 (20)
5070 Ind. Eng. Chem. Res., Vol. 42, No. 21, 2003

Table 1. Elementary Reaction Sets for FTS


FT I 1 CO + s1 ) COs1 FT VI 1 CO + s1 ) COs1
2 COs1 + H2 ) H2COs1 2 COs1 + s1 ) Cs1 + Os1
3 H2COs1 + H2 ) CH2s1 + H2O 3 Cs1 + H2 ) CH2s1
4 H2 + 2s1 ) 2Hs1 4 Os1 + H2 ) H2O + s1
5(n) CH2s1 + Hs1 ) CH3s1 + s1 5 H2 + 2s1 ) 2Hs1
CH2s1 + CH3s1 ) CH3CH2s1 + s1 6(n) CH2s1 + Hs1 ) CH3s1 + s1
CH2s1 + CnH2n+1s1 ) CnH2n+1CH2s1 + s1 CH2s1 + CH3s1 ) CH3 CH2s1 + s1
6(n) CnH2n+1s1 + Hs1 ) CnH2n+2 + 2s1 CH2s1 + CnH2n+1s1 ) CnH2n+1CH2s1 + s1
7(n) CnH2n+1s1 ) CnH2n + Hs1 7(n) CnH2n+1s1 + H2 ) CnH2n+2 + Hs1
FT II 1 CO + s1 ) COs1 8(n) CnH2n+1s1 ) CnH2n + Hs1
2 COs1 + H2 ) H2COs1 FT VII 0 H2 + 2s1 ) 2Hs1
3 H2COs1 + H2 ) CH2s1 + H2O 1(n) CO + Hs1 ) Hs1CO
4 H2 + 2s1 ) 2Hs1 CO + CH3s1 ) CH3s1CO
5(n) CH2s1 + Hs1 ) CH3s1 + s1 CO + CnH2n+1s1 ) CnH2n+1s1CO
CH2s1 + CH3s1 ) CH3CH2s1 + s1 2(n) Hs1CO + Hs1 ) Hs1C + HOs1
CH2s1 + CnH2n+1s1 ) CnH2n+1CH2s1 + s1 CH3s1CO + Hs1 ) CH3s1C + HOs1
6 (n) CnH2n+1s1 + H2 ) CnH2n+2 + Hs1 CnH2n+1s1CO + Hs1 ) CnH2n+1s1C + HOs1
7 (n) CnH2n+1s1 ) CnH2n + Hs1 3(n) CnH2n+1s1C + Hs1 ) CnH2n+1s1CH + s1
FT III 1 CO + s1 ) COs1 4(n) CnH2n+1s1CH + Hs1 ) CnH2n+1s1CH2 + s1
2 COs1 + H2 ) H2COs1 5(n) CnH2n+1s1CH2 ) CnH2n+1CH2s1
3 H2COs1 + H2 ) CH2s1 + H2O 6 HOs1 + Hs1 ) H2O + 2s1
4 H2 + 2s1 ) 2Hs1 7(n) CnH2n+1s1 + Hs1 ) CnH2n+2 + 2s1
5(n) CH2s1 + CH2s1 ) CH2CH2s1 + s1 8(n) CnH2n+1s1 ) CnH2n + Hs1
CH2s1 + CnH2ns1 ) CnH2nCH2s1 + s1 FT VIII 1 CO + s1 ) COs1
6(n) CnH2ns1 + Hs1 ) CnH2n+1s1 + s1 2 COs1 + s1 ) Cs1 + Os1
7(n) CH3s1 + Hs1 ) CH4 + 2s1 3 Cs1 + Hs1 ) CHs1 + s1
CnH2n+1s1 + Hs1 ) CnH2n+2 + 2s1 4 CHs1 + Hs1 ) CH2s1 + s1
8(n) CnH2ns1 ) CnH2n + s1 5(n) CH2s1 + Hs1 ) CH3s1 + s1
FT IV 0 H2 + 2s1 ) 2Hs1 CH2s1 + CH3s1 ) CH3CH2s1 + s1
1 CO + s1 ) COs1 CH2s1 + CnH2n+1s1 ) CnH2n+1CH2s1 + s1
2 COs1 + Hs1 ) HCO s1 + s1 6 H2 + 2s1 ) 2Hs1
3 HCOs1 + Hs1 ) Cs1 + H2Os1 7 Os1 + Hs1 ) HOs1 + s1
4 H2Os1 ) H2O + s1 8 HOs1 + Hs1 ) H2O + 2s1
5 Cs1 + Hs1 ) CHs1 + s1 9(n) CnH2n+1s1 + Hs1 ) CnH2n+2 + 2s1
6 CHs1 + Hs1 ) CH2s1 + s1 10(n) CnH2n+1s1 ) CnH2n + Hs1
7(n) CH2s1 + Hs1 ) CH3s1 + s1 FT IX 1 CO + s1 ) COs1
CH2s1 + CH3s1 ) CH3CH2s1 + s1 2 COs1 + s1 ) Cs1 + Os1
CH2s1 + CnH2n+1s1 ) CnH2n+1CH2s1 + s1 3 Cs1 + Hs1 ) CHs1 + s1
8(n) CnH2n+1s1 + Hs1 ) CnH2n+2 + 2s1 4 CHs1 + Hs1 ) CH2s1 + s1
9(n) CnH2n+1s1 ) CnH2n + Hs1 5(n) CH2s1 + CH2s1 ) C2H4s1 + s1
FT V 0 H2 + 2s1 ) 2Hs1 CH2s1 + CnH2ns1 ) CnH2nCH2s1 + s1
1(n) CO + Hs1 ) Hs1CO 6 H2 + 2s1 ) 2Hs1
CO + CH3s1 ) CH3s1CO 7 Os1 + Hs1 ) HOs1 + s1
CO + CnH2n+1s1 ) CnH2n+1s1CO 8 HOs1 + Hs1 ) H2O + 2s1
2(n) Hs1CO + H2 ) Hs1C + H2O 9(n) CnH2ns1 + Hs1 ) CnH2n+1s1 + s1
CH3s1CO + H2 ) CH3s1C + H2O 10(n) CnH2n+1s1 + Hs1 ) CnH2n+2 + 2s1
CnH2n+1s1CO + H2 ) CnH2n+1s1C + H2O 11(n) CnH2ns1 ) CnH2n + s1
3(n) Hs1C + H2 ) Hs1CH2 FT X 1 CO + s1 ) COs1
CH3s1C + H2 ) CH3s1CH2 2 COs1 + s1 ) Cs1 + Os1
CnH2n+1s1C + H2 ) CnH2n+1s1CH2 3 Cs1 + H2 ) CH2s1
4(n) CnH2n+1s1CH2 ) CnH2n+1CH2s1 4(n) CH2s1 + CH2s1 ) CH2CH2s1 + s1
5(n) CnH2n+1s1 + H2 ) CnH2n+2 + Hs1 CH2s1 + CnH2ns1 ) CnH2nCH2s1 + s1
6(n) CnH2n+1s1 ) CnH2n + Hs1 5 H2 + 2s1 ) 2Hs1
6 Os1 + H2 ) H2O + s1
7(n) CnH2ns1 + Hs1 ) CnH2n+1s1 + s1
8(n) CnH2n+1s1 + H2 ) CnH2n+2 + Hs1
9(n) CnH2ns1 ) CnH2n + s1

Termination of the chain growth can occur via and a dual-site reaction with an adsorbed hydrogen
several routes. For the desorption of an olefin, a atom
single-site mechanism was assumed that R-olefins (1-
alkenes) may be formed by a β-hydride elimination CnH2n+1s1 + Hs1 ) CnH2n+2 + 2s1 (24)
reaction
were considered.
CnH2n+1s1 ) CnH2n + Hs1 (21) It is generally agreed that the R-olefins desorbed from
the catalyst surface can readsorb on the active sites and
or by the desorption of adsorbed alkylidene take place as secondary reactions, which have signifi-
cant influences on the product distribution of the
FTS.29,59
CnH2ns1 ) CnH2n + s1 (22)
On the basis of the above five categories of elementary
reactions, we can define 10 possible mechanisms for
For the desorption of a paraffin, both a single-site FTS, as shown in Table 1.
reaction with molecular hydrogen 3.2.2. Rate Expressions of the Hydrocarbon
Formation. On the basis of the detailed elementary
CnH2n+1s1 + H2 ) CnH2n+2 + Hs1 (23) 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. atoms. Here we introduce a readsorption factor βn,
For each model, the possible rate-determining steps which is defined as follows:
(RDSs) were identified, while all other steps were
assumed to be at quasi-equilibrium. k8- PCnH2n[s1]
βn ) (n g 2) (30)
Before we derive the rate expression of hydrocarbon k8+ [CnH2ns1]
formation, first, it is assumed that the steady-state
conditions are reached for both the surface composition The chain growth probability for the carbon chain
of catalyst and the concentration of all of the intermedi- with n carbon atoms is
ates involved. Second, it is assumed that the rate
constant of the elementary steps for the formation of k5[CH2s1]
hydrocarbons is independent of the carbon number of Rn )
the intermediates involved in the elementary reactions k5[CH2s1] + k7K6[Hs1]2/[s1] + k8+(1 - βn)
except for methane. Third, there are two types of (n g 2) (31)
uniformly distributed active sites respectively for FTS
and WGS reactions on the catalyst surface.4 The concentration of surface intermediates can be
For the derivation of the rate expressions, FT III expressed as a function of the partial pressure of CO,
H2, and H2O by applying the pseudo-equilibrium rela-
in Table 1 will be demonstrated. It is assumed
tion
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
PH22PCO PH22PCO
[CH2s1] ) K1K2K3 [s1] ) K′3 [s1] (32)
and olefins with n carbon atoms can thus be written PH2O PH2O
as
[Hs1] ) xK4PH2[s1] (33)
RCH4 ) k7M[CH3s1][Hs1] )
k7MK6[CH2s1][Hs1]2/[s1] (25) Substitution of eqs 32 and 33 in eq 31 yields

[CnH2ns1]
RCnH2n+2 ) k7[CnH2n+1s1][Hs1] ) Rn ) )
[Cn-1H2n-2s1]
k7K6[CnH2ns1][Hs1]2/[s1] (n g 2) (26) PH22PCO
k5K′3 [s1]
PH2O
RCnH2n ) k8+[CnH2ns1] - k8-PCnH2n[s1] (n g 2) (27)
PH22PCO
k5K′3 [s1] + k7K6K4PH2[s1] + k8+(1 - βn)
The pseudo-steady-state conditions are applied to the PH2O
concentration of surface intermediates [CnH2ns1]: (n g 2) (34)

d[CnH2nS1] Equation 28 can be rearranged as


- ) k5[CnH2ns1][CH2s1] -
dt k5[CH2s1][Cn-1H2n-2s1]
k5[Cn-1H2n-2s1][CH2s1] + k7[CnH2n+1s1][Hs1] + [CnH2ns1] ) +
k5[CH2s1] + k7K6[Hs1]2/[s1] + k8+
k8+[CnH2nS1] - k8-PCnH2n[s1] ) k5[CnH2ns1][CH2s1] -
k5[Cn-1H2n-2s1][CH2s1] + k7K6[CnH2ns1][Hs1]2/[s1] + k8-PCnH2n[s1]
(n g 2) (35)
k8+[CnH2ns1] - k8-PCnH2n[s1] ) 0 (n g 2) (28) k5[CH2s1] + k7K6[Hs1]2/[s1] + k8+

Equation 35 can be rewritten in the following form:13


After rearrangement, eq 28 yields
Xn ) RAXn-1 + BYn (36)
[CnH2ns1]
) where Xn, RA, Yn, and B are defined as follows:
[Cn-1H2n-2s1]
k5[CH2s1] Xn ) [CnH2ns1]

( )
) (n g 2) (37)
k8-PCnH2n[s1]
k5[CH2s1] + k7K6[Hs1]2/[s1] + k8+ 1 - k5[CH2s1]
k8+[CnH2ns1] RA ) )
k5[CH2s1]
k5[CH2s1] + k7K6K4PH2[s1] + k8+
(n g 2) (29)
k5[CH2s1] + k7K6[Hs1]2/[s1] + k8+(1 - βn) PH22PCO
k5K′3 [s1]
PH2O
where k5[CH2s1] is related to the chain growth of (38)
hydrocarbon and k8+(1 - k8-PCnH2n[s1]/k8+[CnH2ns1]) is PH22PCO
the rate of olefin formation, which is the net effect of k5K′3 [s1] + k7K6K4PH2[s1] + k8+
desorption and readsorption of olefins with n carbon PH2O
5072 Ind. Eng. Chem. Res., Vol. 42, No. 21, 2003

k8- Thus, the expression of the concentration of [CnH2n+1s1]


B) +
) is
k5[CH2s1] + k7K6K4PH2[s1] + k8
[CnH2n+1s1] ) K6[CnH2ns1][Hs1]/[s1] )
k8-
(39) PH22.5PCO n
PH22PCO
k5K′3 [s1] + k7K6K4PH2[s1] + k8+
K6K4 0.5
K′3
PH2O
[s1] ∏
i)2
Ri (n g 2) (50)
PH2O

Yn ) PCnH2n[s1] (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.
From eq 36, we get

X2 ) RAX1 + BY2 (41) PH22.5PCO


[CH3s1] ) K6K1K2K3K40.5 [s1] )
PH2O
X3 ) RA2X1 + RABY2 + BY3 (42)
PH22.5PCO
0.5
K6K′3K4 [s1] (51)
X4 ) RA3X1 + RA2BY2 + RABY3 + BY4 (43) PH2O

............ [COs1] ) K1PCO[s1] (52)


n [H2COs1] )
Xn ) RAn-1X1 + B ∑
i)2
RAi-2Y(n-i+2) (n g 2) (44) K2PH2[COs1] ) K1K2PCOPH2[s1] ) K′2PCOPH2[s1] (53)

where X1 and Yn-i+2 are defined as follows: Normalization of the concentration of the sites on the
catalyst surface leads to
X1 ) [CH2s1] ) K3′PH22PCO/PH2O[s1] (45)
1 ) [s1] + [Hs1] + [COs1] + [CH2s1] + [H2COs1] +
n n
Yn-i+2 ) PC(n-i+2)H2(n-i+2)[s1] (n g 2) (46)
[CH3s1] + ∑
i)2
[CnH2ns1] + ∑[CnH2n+1s1]
i)2
(54)
Thus, the concentration of [CnH2ns1] can be rewritten
as Combining eqs 32, 33, and 49-53 with eq 54 gives
[CnH2ns1] )
n
1 ) [s1] + xK P
4 H2 + K1PCO[s1] +
RA n-1
[CH2s1] + B[s1] ∑ RA i-2
PC(n-i+2)H2(n-i+2) ) 2
PH2 PCO
i)2
K′3 [s1] + K1K2PCOPH2[s1] +
n
PH2O
RAn-1K′3PH22PCO/PH2O[s1] + B[s1] ∑
i)2
RAi-2PC (n-i+2)H2(n-i+2)
PH22.5PCO PH22PCO n i
(47) K6K40.5K′3
PH2O
[s1] + K′3
PH2O i)2
∑∏
j)2
(Rj)[s1] +
Substituting eqs 32 and 47 into eq 30 yields

{ [
PH22.5PCO n i
K6K4 K′30.5
∑ ∏(Rj)[s1] (55)
βn ) (k8-/k8+) PCnH2n/ RAn-1K′3PCOPH22/PH2O + PH2O i)2 j)2

The concentration of free active site [s1] can thus be


k8- expressed as follows:

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

]} xK P
n [s1] ) 1/ 1 + 4 H2 + K1PCO +
∑ (RAi-2PC(n-i+2)H2(n-i+2)) (n g 2) (48)
i)2
PH22PCO
According to the definition of chain growth probability K′3 + K1K2PCOPH2 +
Rn in eq 34, we have PH2O

n
PH22.5PCO PH22PCO n i

[CnH2ns1] ) [CH2s1] ∏ Ri )
0.5
K6K4 K′3
PH2O
+ K′3
PH2O
∑ ∏(Rj) +

]
i)2 j)2
i)2
PH22PCO n PH22.5PCO n i
K′3
PH2O
[s1] Ri
i)2
∏ (n g 2) (49) K6K4 K′30.5

PH2O
∑ ∏(Rj)
i)2 j)2
(56)
Ind. Eng. Chem. Res., Vol. 42, No. 21, 2003 5073

Substituting eq 56 into eqs 49 and 50 yields n


RCnH2n ) k8+(1 - βn)K′3PH22PCO/PH2O ∏ Rj /

[
[CnH2ns1] ) j)2

[
PH22PCO
xK P
2
PH2 PCO n 1+ + K1PCO + K′3 +

4 H2
K′3 Ri/ 1 + xK4PH 2
+ K1PCO + PH2O
PH2O i)2
PH22.5PCO
PH22PCO PH22.5PCO K1K2PCOPH2 + K6K40.5K′3 +
PH2O

]
0.5
K′3 + K1K2PCOPH2 + K6K4 K′3 +
PH2O PH2O PH22PCO n i

]
PH22PCO PH22.5PCO
K′3
P H 2O x
(1 + K6 K4PH2)
i)2
∑∏
j)2
(Rj) (61)
n i n i
K′3
PH2O
∑ ∏
i)2 j)2
(Rj) + K6K40.5K′3
PH2O
∑ ∏(Rj)
i)2 j)2 3.3. Kinetic Models for the WGS Reaction. The
kinetics of the WGS reaction has been intensively
(n g 2) (57) studied by many researchers, and several mechanisms
were proposed.60-64 It is generally accepted that the
[CnH2n+1s1] ) WGS reaction over supported iron catalysts proceeds via

[
a mechanism of formate species due to a limited change
PH22.5PCO n of oxidation states of the iron cations.4,60-64 Rethwisch
0.5
K6K4 K′3
PH2O

i)2
Ri/ 1 + xK4PH 2
+ K1PCO + and Dumesic61 suggested that the WGS reaction pro-
ceeds on active sites different from those for FTS, and
for supported iron catalysts, the magnetite is the most
PH22PCO PH22.5PCO active phase for the WGS. On the basis of the formate
0.5 intermediate mechanism, the elementary steps of the
K′3 + K1K2PCOPH2 + K6K4 K′3 + WGS, and corresponding expressions of rates, equilib-
PH2O PH2O
rium constants are summarized in Table 2.

]
If we assumed that the slowest step in the WGS is
PH22PCO n i PH22.5PCO n i step IV (WGS 3), the rate of CO2 formation can be
K′3
PH2O
∑ ∏
i)2 j)2
(Rj) + K6K4 K′3 0.5

PΗ2Ο
∑ ∏(Rj)
i)2 j)2
written as follows:

RCO2 ) r4 ) kWGS4[COOH-s2] -
(n g 2) (58)
k-WGS4PCO2[H-s2] (62)
Substituting eqs 57 and 58 into eqs 25-27, we have
From the elementary step listed in Table 3, we have
RCH4 ) [CO-s2] ) KWGS1PCO[s2] (63)

[
3
PH2 PCO
KWGS2PH2O[s2]2 ) [OH-s2][H-s2] (64)
k7MK4K6K′3
PH2O
/ 1+ x K4PH2 + K1PCO +
KWGS3[OH-s2][CO-s2] ) [COOH-s2][s2] (65)
2 2.5
PH2 PCO PH2 PCO
K′3 + K1K2PCOPH2 + K6K40.5K′3 + [H-s2] ) xKWGS5PH2[s2] (66)
PH2O PH2O

]
On the basis of eqs 63-66, the concentrations of the
PH22PCO n i
2 surface intermediates [CO-s2], [OH-s2], and [COOH-
K′3
PH2O
(1 + K6 K4PH2)
i)2
x ∑∏
j)2
(Rj) (59) s2] can be derived as follows:

[CO-s2] ) KWGS1PCO[s2] (67)


RCnH2n+2 )
[OH-s2] ) KWGS2KWGS5-0.5PH2OPH2-0.5[s2] (68)

k7K4K6K′3
PH23PCO

PH2O

n

j)2
Rj /
[
1+ xK P
4 H2 + K1PCO + [COOH-s2] )
KWGS1KWGS2KWGS3KWGS5-0.5PH2OPCOPH2-0.5[s2] (69)
PH22PCO PH22.5PCO
0.5 If Kp is used to represent the equilibrium constant of
K′3 + K1K2PCOPH2 + K6K4 K′3 + the WGS reaction, it can be expressed by the equilib-
PH2O PH2O rium partial pressures of CO, CO2, H2, and H2O.

PH22PCO

]
2 /
n i PCO P/
x ∑∏
2 H2
K′3 (1 + K6 K4PH2) (Rj) (60) Kp ) (70)
PH2O i)2 j)2 P/COPH
/
2O
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 elementary reaction expressions of rates and equilibrium constants
I CO + s2 ) COs2 KWGS1 ) [COs2]/PCO[s2]
II H2O + 2s2 ) OHs2 + Hs2 KWGS2 ) [OHs2][Hs2]/PH2O[s2]2
III COs2 + OHs2 ) COOHs2 + s2 KWGS3 ) [COOHs2][s2]/[COs2][OHs2]
IV COOHs2 ) CO2 + Hs2 r4 ) kWGS4[COOHs2] - k-WGS4PCO2[Hs2]
KWGS4 ) kWGS4/k-WGS4
V 2Hs2 ) H2 + 2s2 1/KWGS5 ) PH2[s2]2/[Hs2]2

Table 3. Rate Expressions for the WGS Reaction


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

According to eqs 63, 64, and 66, the partial pressures Thus, the expressions of [H-s2] and [COOH-s2] can be
of CO, H2O, and H2 can be written as follows: obtained by substituting eq 80 into eqs 66 and 78.

P/CO ) [CO-s2]/(KWGS1[s2]) (71)


[H-s2] )
xKWGS5PH 2
(81)
/ 2 1 + KpKWGS4KWGS5 PCOPH2O/PH20.5
0.5
PH 2O
) [OH-s2][H-s2]/(KWGS2[s2] ) (72)

/
[COOH-s2] )
PH ) [H-s2]2/(KWGS5[s2]2) (73)
2
KpKWGS4KWGS50.5PCOPH2O/PH20.5
(82)
If we assumed that step IV reached the equilibrium 1 + KpKWGS4KWGS50.5PCOPH2O/PH20.5
state, the expression of partial pressure of CO2 at the
/
equilibrium state (PCO 2
) can be obtained: Finally, the rate expression of CO2 formation can be
/ expressed as
PCO 2
) [COOH-s2]*/(KWGS4[H-s2]*) (74)
kv(PCOPH2O/PH20.5 - PCO2PH20.5/Kp)
Substituting eq 71-74 into eq 70 gives RCO2 ) (83)
1 + KvPCOPH2O/PH20.5
KWGS1KWGS2[COOH-s2]*[s2]*
Kp ) (75) where Kv ) KpKWGS4KWGS50.5 and kv ) kWGS4Kv.
KWGS4KWGS5[CO-s2]*[OH-s2]*
By different assumptions of the RDS in WGS, other
When eq 65 is substituted into eq 75, Kp can be rate expressions of CO2 of the WGS reaction can be
expressed as derived, and the results are tabulated in Table 3. The
equilibrium constant Kp of the WGS reaction can be
KWGS1KWGS2KWGS3 calculated by the following relation:11-13
Kp ) (76)
KWGS4KWGS5 5078.0045
Kp ) - 5.8972089 + 13.958689 ×
T
Equation 76 can rewritten as follows:
10-4T - 27.592844 × 10-8T 2 (84)
KWGS4KWGS5KP A total of 39 kinetic models of FTS can be obtained by
KWGS3 ) (77)
KWGS1KWGS2 the combination of 13 FTS models with 3 WGS models.

Substituting eq 77 into eq 69 yields


4. Results and Discussion
[COOH-s2] )
4.1. Experimental Results. Iron-based Fischer-
KpKWGS4KWGS50.5PCOPH2OPH2-0.5[s2] (78) Tropsch catalysts, once exposed in the synthesis gas
under typical reduction or initial reaction conditions,
As a result of the assumption that the RDS is step are normally reduced by H2 and CO, transforming from
IV, it can be considered that the concentration of the their oxide phases to metallic and carbide phases.40,41-45
adsorbed species [COOH-s2] is much larger than those During this transformation stage, the phases of cata-
of the other adsorbed species. lysts greatly change with time on stream as well as
changes in operation conditions, leading to irreproduc-
[COOH-s2] + [s2] ) 1 (79) ible experimental data. To minimize (it can never be
completely eliminated in FTS over iron catalysts) the
The concentration of the empty active site, [s2], can be effect of a “sharp” phase change on the quality of kinetic
obtained by substituting eq 78 into eq 79 data, the catalyst used for kinetic tests was reduced and
then stabilized at a fixed operation condition for about
1 1000 h, aiming at the complete establishment of the
[s2] ) (80) catalyst phase for FTS. After this stabilization stage,
1 + KpKWGS4KWGS50.5PCOPH2O/PH20.5 operation conditions were switched according to kinetic
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 83 180 273 404 516 641 845 960
temperature (K) 543 543 556 556 556 556 556 556 556
pressure (MPa) 2.30 2.30 2.30 2.30 2.30 2.30 2.30 2.30 2.30
GHSV (10-4 Nm3 kg of catalyst-1 s-1) 5.1 5.1 5.1 5.1 5.1 5.1 5.1 5.1 5.1
CO conversion % 94.8 94.6 96.1 96.9 95.9 96.4 96.0 96.1 96.0
H2 conversion % 43.7 40.1 44.3 47.6 41.7 44.6 47.4 45.8 44.8
H2/CO (in feed gas) 2.0 2.0 2.0 2.0 2.0 2.0 2.0 2.0 2.0
H2/CO (in tail gas) 21.55 22.13 28.33 34.35 28.61 30.43 25.95 27.25 28.21
selectivity CO2 (%) 36.09 40.88 38.71 39.13 36.73 39.27 40.37 38.70 38.80
selectivity (wt %)
C1 7.51 9.39 13.35 14.44 15.62 16.80 16.76 16.84 16.90
C20 3.00 4.26 6.17 7.04 7.41 8.20 8.20 8.24 8.30
C2) 5.31 5.18 5.35 4.81 4.87 4.91 4.85 4.39 4.41
C30 1.76 2.18 2.77 2.94 3.32 3.55 3.63 3.59 3.63
C3) 7.63 8.15 8.82 8.91 8.61 9.08 9.28 8.47 8.50
C40 1.51 1.83 2.07 2.11 2.31 2.51 2.51 2.49 2.48
C4) 6.04 6.19 6.58 6.46 6.68 6.63 7.05 6.37 6.40
C2-4)/tol‚HC (wt %) 18.99 19.51 20.75 20.18 20.16 20.62 21.18 19.23 19.21
C2-4)/C2-40 2.89 2.19 1.70 1.48 1.38 1.28 1.30 1.18 1.17
C5+ 67.24 62.82 54.89 53.29 51.19 48.33 47.72 49.61 49.40

sampling arrangements. The reaction results during the table of L16(43)). The experimental results obtained in
stabilization stage are summarized in Table 4 and this investigation for different reaction temperatures,
Figure 2. From these results, it can be found that the total reaction pressures, space velocities, and H2/CO
catalyst reached a stable stage after about 500 h on ratios are given in Table 5. Five more points were added
stream under the conditions used. The selectivity to CO2 to consider more reaction conditions, among which one
was maintained at the level of 40 ( 2%, that to CH4 at point (no. 16 in Table 5) was the reference point (no. 12
16 ( 1%, that to C5+ at 50 ( 2%. For the activity, CO in Table 5). The order of kinetic experiments was from
conversion reached to a level of 95 ( 1%. This confirmed no. 1 to no. 21 listed in Table 5. From the results listed,
that kinetic data sampling may be based on a rather it can be found that the CO and H2 conversions at no.
stable stage of the catalyst in the following kinetic 16 well reproduced those at no. 12, indicating that the
experimental stage. catalytic phase reached a stable state.
Kinetic sampling conditions were arranged according 4.2. Estimation of the Kinetic Parameters. 4.2.1.
to orthogonal arrangement of sampling points, enabling Reactor Model of the Fixed Bed. The reactor model
efficient and optimal distribution of experimental points.13 for describing the kinetic experimental conditions is
The reaction condition variables that need to be con- assumed to be a plug-flow homogeneous state. At this
sidered are temperature, pressure, and H2/CO ratio. For stage, the transportation in catalyst pores and the
each variable, four values were planned, bringing about solubilities of different hydrocarbons are not considered
16 experimental points (corresponding to an orthogonal 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


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

Table 6. Stoichiometric Coefficients of the Matrix for FTS Reactions


reaction path
component
reactants CO + H2O CO + 3H2 2CO + 4H2 2CO + 5H2 nO + 2nH2 nO + (2n + 1)H2
products CO2 + H2 CH4 + H2O C2H4 + 2H2O C2H6 + 2H2O CnH2n + nH2O CnH2n+2 + nH2O
CO -1 -1 -2 -2 -n -n
H2 1 -3 -4 -5 -2n -(2n + 1)
CO2 1 0 0 0 0 0
H2O -1 1 2 2 n n
CH4 0 1 0 0 0 0
C2H4 0 0 1 0 0 0
C2H6 0 0 0 1 0 0
...
CnH2n 0 0 0 0 1 0
CnH2n+2 0 0 0 0 0 1

integration of the reactor model embedded in a param- (RCnH2n+2, RCnH2n, RCO2, ...). The partial pressure of the
eter optimization procedure. This simplification may ith component can be calculated by using the following
lose the transportation-enhanced readsorption phenom- formula:
ena, and further modification to the kinetic models
obtained is needed. This will be discussed elsewhere. mi
From the group of equations (1)-(3), the total number Pi ) PT (i ) 1, 2, 3, ..., Nc) (87)
Nc
of the components, Nc, and of the reactions, NR, are 2N
+ 3 and 2N, respectively, where N is the maximum ∑ mi
carbon number of hydrocarbons considered. The matrix i)1
of the stoichiometric coefficients between the products
and reactions according to the FTS and the WGS where PT is the total pressure in the reactor.
reaction is listed in Table 6.11-13 Numerical integration of the continuity equation (85)
The model of the fixed-bed reactor used in our kinetic is performed by using Gear’s method.65
study can be described as follows:11-13 4.2.2. Optimization Method. In the estimation of
parameters of the kinetic model, the Levenberg-Mar-
dmi NR quardt (LM) algorithm still plays an important role.

dW
) ∑
j)1
RijRj (i ) 1, 2, ..., Nc; j ) 1, 2, ..., NR) (85) 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
with the initial condition
locating global minima. In our paper, to avoid getting
trapped in local minima, the parameters of the various
W ) 0; mi ) mi,0 (86) rival models in this paper were estimated in a first step
using the genetic algorithm (GA) approach that was
where W is the weight of the catalyst used, mi and mi,0 developed in this group67 and then using the LM
are the mole flow rates of component i along the reactor algorithm to make refined optimization, after which
axis and at the inlet of the reactor, NR is the number of the statistical tests and the physiochemical constraints
reactions involved, Nc is the total number of compo- are used to evaluate the significance of models and
nents, Rij is the stoichiometric coefficient for the ith parameters. The GA algorithm encoded in this group
component in reaction j, and Rj is the rate of reaction j using a real-number code chromosome representation to
Ind. Eng. Chem. Res., Vol. 42, No. 21, 2003 5077

Table 7. Values of the Parameters for the Mechanism FT III WGS3a


parameter value unit parameter value unit
k5,0 7.88 × 103 mol g-1 s-1 bar-1 Ev 58.43 kJ mol-1
E5 75.52 kJ mol-1 k-8 2.77 × 10-5 mol g-1 s-1 bar-1
k7M,0 2.01 × 106 mol g-1 s-1 bar-1 Kv 2.76 × 10-2 bar-0.5
E7M 97.39 kJ mol-1 K1 2.59 bar-1
k7,0 1.10 × 106 mol g-1 s-1 bar-1 K2 1.67 × 10-3 bar-1
E7 111.48 kJ mol-1 K3 8.34 × 10-2
k8,0 8.79 × 103 mol g-1 s-1 K4 1.21 bar-1
E8 97.37 kJ mol-1 K6 0.10
kv,0 3.42 mol g-1 s-1 bar-1.5
a 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 The relative residual (RR) between experimental and
parameters. The GA procedure consists mainly of four calculated values of responses will be used to show the
parts: randomly producing an initial population, ran- deviations between the model and experiment
domly selecting two individuals from the population
space to crossover (using certain operator), producing the mi,exp - mi,cal
next generation of population, nonuniform mutation, and RR ) × 100 (91)
mi,exp
repeating these steps to scan the whole searching space.
For estimation of the models, 20 model responses
in the regression procedure were the outlet concen- The dependence of the reaction rate parameters on
trations of 15 paraffins and olefins (such as CH4, C2H4, temperature can be described by the Arrhenius law.
C2H6, C3H6, C3H8, C4H8, C4H10, etc.) selected as the
most significant Fischer-Tropsch products, along with ki(T) ) ki,0 exp(-Ei/RT) (92)
the concentrations of CO2, the conversions of CO and
H2, and the overall concentrations of C5+ and C11+ 4.2.3. Parameter Estimation. For scanning the
hydrocarbons. The objective function is defined as models by parameter optimization, several basic physi-
follows: cal criteria are applied: the rate constants and equi-
librium constants should be positive, and the activation

(mi,exp - mi,cal
) energies for the paraffins and olefin formation and for
2
fi,obj ) (i ) 1, 2, ..., Nexp) (88) the WGS reaction should be in the range of values
mi,exp reported by other researchers. The optimization proce-
dure by the GA and LM methods has been set to find
where mi,exp and mi,cal are the experimental and calcu- the minimum, which provided (1) a reasonable fit to the
lated values of conversions of reactants or the selectivi- experimental data, (2) physically meaningful values of
ties of products, respectively, and Nexp is the total the model parameters, and (3) acceptable values of
number of experimental runs. Because of the com- statistical parameters, such as F values for the models
plexity of the models, a multiresponse objective function and t values for the parameters.
should be introduced, which is expressed in the follow- A total of 25 kinetic models are rejected because of
ing form: unreasonable values of the parameters. The discrimina-

( )
tion of the remaining 14 rival models is performed by
Nresp Nexp mij,exp - mij,cal 2 comparing first their MARR and second statistical tests
FTol,obj ) ∑ ∑ Wi
mij,exp
on models (F test) and parameters (t test). The result
i)1 j)1 shows that only two kinetic models’ MARR are within
(i ) 1, 2, ..., Nresp; j ) 1, 2, ..., Nexp) (89) 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
where Wi represents the weighting factor of the ith the remained best model are listed in Table 7, and
objective function (it expresses the relative importance all of them are statistically significant. The estimated
of the ith response). Those responses with the most activation energy for chain growth, E5, is 75.52 kJ
accurate measurement and/or with special significance mol-1, indicating that the methylene insertion into
in the regression are provided with greater weights): the metal-alkylidene bond has a moderate height of
Nresp is the number of responses in the system, mij,exp is energy barrier that it needs to overcome. The activation
the experimental value of the ith response for the jth energy for methane formation, E7M, is 97.39 kJ mol-1,
kinetic experiment, and mij,cal is the calculated value of which is in good agreement with that reported by
the ith response for the jth kinetic experiment. Vannice58 (89 kJ mol-1) and smaller than that for other
The adjustable model parameters for the several paraffin formation, E7, with a value of 111.48 kJ mol-1,
kinetic models were calculated by minimizing FTol,obj which can explain the higher selectivity of methane than
with the GA and LM methods. those of other paraffins. The activation energy of olefin
The accuracy of the fitted model relative to the formation is 97.37 kJ mol-1, much smaller than that of
experimental data was obtained from the MARR (mean paraffins, which can interpret the general fact that the
absolute relative residuals) function: selectivity is much higher to alkenes on the Fe-Mn

( )
catalysts than on other iron-based catalysts.67 This is
Nresp Nexp mi,exp - mi,cal 1 different from Dictor and Bell69 (80-90 kJ mol-1 for
MARR ) ∑ ∑ mi,exp NrespNexp
× 100 (90) paraffins and 100-110 kJ mol-1 for olefins) as well as
j)1 i)1 Lox and Froment11,12 (94.5 kJ mol-1 for paraffins and
5078 Ind. Eng. Chem. Res., Vol. 42, No. 21, 2003

Figure 5. Comparison of the calculated and experimental product


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

Figure 6. Comparison of the calculated and experimental product


Figure 4. RRs of calculated and experimental CO2 selectivity (FT distribution (Lox and Froment’s model reaction conditions: T )
III WGS3). 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 calcu-
lated 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,
Figure 7. Comparison of the calculated and experimental product
and 10 by Lox and Froment’s ASF model. The ASF distribution (FT III WGS3 reaction conditions: T ) 585 K, P )
model appears to give a strong deviation for the selec- 3.02 MPa, H2/CO ) 2.04, GHSV ) 3.2 × 10-3 Nm3 kg of catalyst-1
tivity to hydrocarbons, lower to methane and higher to s-1).
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 orientations of kinetic and reactor models in considering
experimental results. The modeled product distributions the intrinsic mechanism information and the transpor-
using FT III WGS3 are in good agreement with the tation effects. The kinetic model developed is significant
experimental selectivities, and the deviation for meth- only for the cases excluding the diffusivity and solubility
ane is described fairly accurately. factors, which are believed to be significant in enhancing
It should be noted that the current kinetic model has the olefin readsorption and secondary reaction and,
not considered the effect of diffusivity and solubility of therefore, in changing the olefin and paraffin selectivi-
olefins with different chain sizes on the paraffin/olefin ties. However, kinetic experiments can never exclude
ratio on the basis of the understanding of different 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 Froment’s model reaction conditions: T ) Figure 11. Comparison of the calculated and experimental olefin/
585 K, P ) 3.02 MPa, H2/CO ) 2.04, GHSV ) 3.2 × 10-3 Nm3 kg paraffin ratios for different carbon numbers (FT III WGS3 reaction
of catalyst-1 s-1). 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 experi-
mentally observed olefin/paraffin ratio dependence cor-
rectly. This later work is too extensive to present here
and will be discussed elsewhere.

5. Conclusions
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 Kinetic experiments of the Fischer-Tropsch reaction
catalyst-1 s-1). 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 experi-
mental data. The best model shows that two elementary
steps, the insertion of methylene into the metal-
alkylidene 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 selec-
Figure 10. Comparison of the calculated and experimental
product distribution (Lox and Froment’s model reaction condi-
tivity on Fe-Mn catalysts than on other iron-based
tions: T ) 601 K, P ) 1.51 MPa, H2/CO ) 3.05, GHSV ) catalysts.
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 Financial support from the Chinese Academy of
seen that the kinetic model without considering the Sciences (Project No. KGCX1-SW-02), Committee of
transportation effect (diffusivities and solubilities) has Science and Technology of China via 863 plan (Project
not predicted the correct dependence compared with No. 2001AA523010), Shanxi Natural Science Founda-
experimental results. This again proved the understand- tion (20031032), and the National Natural Sciences
ing in FTS that pure olefin readsorption and secondary Foundation of China (Project No. 29673054) is gratefully
reaction have little consequence to the observed olefin/ 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.
n i
RCH4 ) k6MK4PH2R1/(1 + x K4PH2 + K1PCO + K′3PH22PCO/PH2O + K1K2PH2PCO + x K4PH2 ∑ ∏
i)1 j)1
Rj))2
( (A1)

n n i
RCnH2n+2 ) k6K4PH2 ∏
j)1
Rj /(1 + x K4PH2 + K1PCO + K′3PH22PCO/PH2O + K1K2PH2PCO + x K4PH2 ∑ ∏
i)1 j)1
Rj))2
( (n g 2)
(A2)
n
RCnH2n ) k7+(1 - βn) K4PH2x ∏
j)1
Rj /[1 + xK4PH 2
+ K1PCO + K′3PH22PCO/PH2O + K1K2PH2PCO +
n i

xK P ∑(∏R )] (n g 2) (A3)

{[ ]}
4 H2 j
i)1 j)1

k7- n
- +
βn ) (k7 /k7 ) PCnH2n/ R1RA n-1
+
2
∑ (RAi-2PC (n-i+2)H2(n-i+2)
) (n g 2) (A4)
PCOPH2 i)2

k5K′3
PH2O x
[s1] + k6 K4PH2[s1] + k7+

k5K′3PH22PCO/PH2O
Rn ) (n g 2) (A5)
k5K′3PH22PCO/PH2O + k6xK4PH2 + k7+(1 - βn)/[s1]

k5K′3PH21.5PCO/PH2O
R1 ) (A6)
k5K′3PH21.5PCO/PH2O + k6xK4

PH22PCO
k5K′3 [s1]
PH2O
RA ) (A7)
PH22PCO
k5K′3 [s1] + k6xK4PH2[s1] + k7+
PH2O

Model FT II: RDS, steps 5-7.

0.5
RCH4 ) k6MK4 PH2 R1/ 1 + 1.5
[ x K4PH2 + K1PCO + K′3
PH22PCO

PH2O
+ K1K2PH2PCO + x K4PH2 ∑
n

∏Rj)
i)1 j)1
(
i

] (A8)

RCnH2n+2 )

0.5
k6K4 PH2 1.5

n

j)1 [
Rj / 1 + x K4PH2 + K1PCO + K′3
PH22PCO

PH2O
+ K1K2PH2PCO + x K4PH2 ∑
n

∏Rj)
(
i)1 j)1
i

] (n g 2) (A9)

RCnH2n )

+
x
k7 (1 - βn) K4PH2 ∏
j)1
n

[
Rj / 1 + x K4PH2 + K1PCO + K′3
PH22PCO

PH2O
+ K1K2PH2PCO + x K4PH2 ∑
n

∏Rj)
i)1 j)1
(
i

] (n g 2)

(A10)

PH22PCO
k5K′3 [s1]
PH2O
Rn ) (n g 2) (A11)
PH22PCO
k5K′3 [s1] + k6PH2 + k7+(1 - βn)
PH2O
Ind. Eng. Chem. Res., Vol. 42, No. 21, 2003 5081

PH22PCO
k5K′3 [s1]
PH2O
R1 ) (A12)
PH22PCO
k5K′3 [s1] + k6
PH2O

PH22PCO
k5K′3 [s1]
PH2O
RA ) (A13)
PH22PCO
k5K′3 [s1] + k6PH2 + k7+

{ [ ]}
PH2O

k7- n
- +
βn ) (k7 /k7 ) PCnH2n/ R1RA n-1
+
2
∑ (RAi-2PC (n-i+2)H2(n-i+2)
) (n g 2) (A14)
PCOPH2 i)2

k5K′3 [s1] + k6PH2 + k7+


PH2O

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

RCH4 ) k7MK4K6K′3
PH23PCO

PH2O
/ 1+
[ xK P 4 H2 + K1PCO + K′3
PH22PCO

PH2O
+ K1K2PCOPH2 + K6K40.5K′3
PH22.5PCO

PH2O
+

PH22PCO

]
2
n i
K′3
PH2O
(1 + K6 K4PH2) x
i)2
∑∏
j)2
(Rj) (A15)

RCnH2n+2 ) k7K4K6K′3
PH23PCO

PH2O

n
Rj /
[ 1+ xK P 4 H2 + K1PCO + K′3
PH22PCO

PH2O
+ K1K2PCOPH2 + K6K40.5K′3
PH22.5PCO

PH2O
+

]
j)2

PH22PCO n i
2

K′3
PH2O
(1 + K6 K4PH2)xi)2
∑∏
j)2
(Rj) (n g 2) (A16)

+ 2
RCnH2n ) [k8 (1 - βn)K′3PH2 PCO/PH2O ∏
n

[
Rj]/ 1 + xK4PH 2
+ K1PCO + K′3
PH22PCO

PH2O
+ K1K2PCOPH2 +

]
j)2

PH22.5PCO PH22PCO n i
K6K4 K′3 0.5

PH2O
+ K′3
PH2O x
(1 + K6 K4PH2)
i)2
∑∏
j)2
(Rj) (n g 2) (A17)

PH22PCO
k5K′3
PH2O
Rn ) (n g 2) (A18)
PH22PCO
k5K′3 + k7K6K4PH2 + k8+(1 - βn)/[s1]
PH2O

- +

{ [
βn ) (k8 /k8 ) PCnH2n/ RA n-1
K′3PCOPH2 /PH2O + 2
k8-

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


×


i)2
n
(RAi-2PC (n-i+2)H2(n-i+2)
)
]} (n g 2) (A19)

k5K′3PH22PCO/PH2O[s1]
RA ) (A20)
k5K′3PH22PCO/PH2O[s1] + k7K6K4PH2[s1] + k8+
5082 Ind. Eng. Chem. Res., Vol. 42, No. 21, 2003

Model FT IV: RDS, steps 7-9.

RCH4 ) k8MK0PH2R1/(1 + xK P 0 H2 x
+ K1PCO(1 + K2 K0PH2) + PH2O/K4 +
n i
K4K3K2K1K0PCOPH2/PH2O(1 + K5 K0PH2 + K6K5K0PH2) + x xK P ∑(∏R ))
0 H2
i)1 j)1
j
2
(A21)

n
RCnH2n+2 ) k8K0PH2 ∏
j)1
Rj /(1 + xK0PH 2 x
+ K1PCO(1 + K2 K0PH2) + PH2O/K4 +
n i
K4K3K2K1K0PCOPH2/PH2O(1 + K5 K0PH2 + K6K5K0PH2) + x x K0PH2 ∑ ∏(
i)1 j)1
Rj))2 (n g 2) (A22)

x
RCnH2n ) k9+(1 - βn) K0PH2 ∏
j)1
Rj /[1 + xK0PH 2 x
+ K1PCO(1 + K2 K0PH2) + PH2O/K4 +
n i
K4K3K2K1K0PCOPH2/PH2O(1 + K5 K0PH2 + K6K5K0PH2) + x x K0PH2 ∑ (∏Rj)]
i)1 j)1
(n g 2) (A23)

k7K′3PH22PCO/PH2O
Rn ) (n g 2) (A24)
k7K′3PH22PCO/PH2O + k8xK0PH2 + k9+(1 - βn)/[s1]

k7K′3PH22PCO/PH2O
R1 ) (n ) 1) (A25)
k7K′3PH22PCO/PH2O + k8xK1PH2

{ [ k9-

]}
n
- +
βn ) (k9 /k9 ) PCnH2n/ R1RA n-1
+ ∑(RAi-2PC (n-i+2)H2(n-i+2)
) (n g 2)
2
k7K′3PH2 PCO/PH2O[s1] + k8 K0PH2[s1] + k9 x + i)2

(A26)

k7K′3PH22PCO/PH2O[s1]
RA ) (A27)
k7K′3PH22PCO/PH2O[s1] + k8xK0PH2[s1] + k9+

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

[ ( )x ]
1 PH2O 1 1 1 n i

x
RCH4 ) k5MPH2 K0PH2R1/ 1 + x K0PH2 + 1 +
K2K3K4 P 2
+
K3K4 PH2
+
K4
K0PH2 ∑ ∏Rj)
i)1 j)1
( (A28)
H2

[ ( )x ]
n
1 PH2O 1 1 1 n i

x
RCnH2n+2 ) k5PH2 K0PH2 ∏
j)1
Rj / 1 + x K0PH2 + 1 +
K2K3K4 P 2
+
K3K4 PH2
+
K4
K0PH2 ∑ ∏R ) (
i)1 j)1
j (n g 2) (A29)
H2

[ ( )x ]
n
1 PH2O 1 1 1 n i

x
RCnH2n ) k6+(1 - βn) K4PH2 ∏
j)1
Rj / 1 + xK0PH2 + 1 +
K2K3K4 P 2
+
K3K4 PH2
+
K4
K0PH2 ∑ ∏Rj)
i)1 j)1
( (n g 2)
H2
(A30)
k1PCO
Rn ) (n g 2) (A31)
k1PCO + k5PH2 + k6+(1 - βn)

k1PCO
R1 ) (n ) 1) (A32)
k1PCO + k5PH2

k1PCO
RA ) (A33)
k1PCO + k5PH2 + k6+
Ind. Eng. Chem. Res., Vol. 42, No. 21, 2003 5083

- +

{ [
βn ) (k6 /k6 ) PCnH2n / R1RA n-1
+
k6-

k1PCO + k5PH2 + k6 +
n


i)2
(RAi-2PC (n-i+2)H2(n-i+2)
]}
) (n g 2) (A34)

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

[ ( )x
n i
1 1
x
RCH4 ) k5MPH2 K0PH2R1 / 1 + xK0PH2 + 1 +
K3K4PH2
+
K4
K0PH2 ∑ ∏Rj) +
(

]
i)1 j)1
n i

x
K1PCO K0PH2(1 + ∑ ∏Rj))
(
i)1 j)1
(A35)

[ ( )x
n n i
1 1
x
RCnH2n+2 ) k5PH2 K0PH2 ∏ Rj / 1 + x
K0PH2 + 1 +
K3K4PH2
+
K4
K0PH2 ∑ ∏Rj) +
(

]
j)1 i)1 j)1
n i
K1PCO K0PH2(1 + x ∑ ∏ Rj))
(
i)1 j)1
(n g 2) (A36)

[ ( )x
n n i
1 1
x
RCnH2n ) k6+(1 - βn) K0PH2 ∏ Rj / 1 + x K0PH2 + 1 +
K3K4PH2
+
K4
K0PH2 ∑ ∏Rj) +
(

]
j)1 i)1 j)1
n i
K1PCO K0PH2(1 + x ∑ ∏Rj))
(
i)1 j)1
(n g 2) (A37)

k2PCOPH2
Rn ) (n g 2) (A38)
k2PCOPH2 + k5PH2 + k6+(1 - βn)

k2PCO
R1 ) (n ) 1) (A39)
k2PCO + k5

k2PCOPH2
RA ) (A40)
k2PCOPH2 + k5PH2 + k6+

- +

{ [
βn ) (k6 /k6 ) PCnH2n/ R1RA n-1
+
k6-

k2PCOPH2 + k5PH2 + k6 +

i)2
n
(RAi-2PC (n-i+2)H2(n-i+2)
)
]} (n g 2) (A41)

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

x
RCH4 ) k5MPH2 K0PH2R1/ 1 +
[ x K0PH2 + 1 +
( )x 1
K4
K0PH2
n

∑ (
i

∏Rj) +

( ) ]
i)1 j)1

K1K2PCOPH2 n i

x K0PH2 K1PCO +
PH2O
(1 + ∑ (∏Rj))
i)1 j)1
(A42)

x
RCnH2n+2 ) k5PH2 K0PH2 ∏
n

[
Rj / 1 + x
K0PH2 + 1 +
( )x 1
K4
K0PH2
n

∑ ∏Rj) +
(
i

( ) ]
j)1 i)1 j)1

K1K2PCOPH2 n i

x K0PH2 K1PCO +
PH2O
(1 + ∑ ∏Rj))
(
i)1 j)1
(n g 2) (A43)

+
x
RCnH2n ) k6 (1 - βn) K0PH2 ∏
n
Rj /
[
1+ xK P 0 H2
( )x
+ 1+
K4
1
K0PH2 ∑
n
(∏Rj) +
i

( ) ]
j)1 i)1 j)1

K1K2PCOPH2 n i

x K0PH2 K1PCO +
PH2O
(1 + ∑ (∏Rj))
i)1 j)1
(n g 2) (A44)
5084 Ind. Eng. Chem. Res., Vol. 42, No. 21, 2003

k3PCOPH22/PH2O
Rn ) (n g 2) (A45)
k3PCOPH22/PH2O + k5PH2 + k6+(1 - βn)

k3PCOPH22/PH2O
R1 ) (n ) 1) (A46)
k3PCOPH22/PH2O + k5M

k3PCOPH22/PH2O
RA ) (A47)
k3PCOPH22/PH2O + k5PH2 + k6+

{ [
βn ) (k6-/k6+) PCnH2n/ R1RAn-1 +
2
k6-

k3PCOPH2 /PH2O + k5PH2 + k6 +



i)2
n
(RAi-2PC (n-i+2)H2(n-i+2)
)
]} (n g 2) (A48)

Model FT VI: RDS, steps 6-8.

RCH4 ) k7MK50.5PH21.5R1/ 1 +
[ xK P 5 H2 + K1PCO + K4-1
PH2O

PH2
+ K1K2K4
PH2PCO

PH2O
+ K′3
PH22PCO

PH2O
+

x K5PH2 ∑
n
(∏Rj)
i)1 j)1
i

] (A49)

0.5
RCnH2n+2 ) k7K5 PH2 1.5

n
Rj /
[
1+ xK P 5 H2 + K1PCO + K4 -1
PH2O

PH2
+ K1K2K4
PH2PCO

PH2O
+ K′3
PH22PCO

PH2O
+

]
j)1

n i

x K5PH2 ∑ ∏Rj)
(
i)1 j)1
(n g 2) (A50)

+
x
RCnH2n ) k8 (1 - βn) K5PH2 ∏
n

[
Rj / 1 + xK5PH 2
+ K1PCO + K4 -1
PH2O

PH2
+ K1K2K4
PH2PCO

PH2O
+ K′3
PH22PCO

PH2O
+

]
j)1

n i

xK5PH2 ∑ ∏Rj)
(
i)1 j)1
(n g 2) (A51)

PH22PCO
k6K′3 [s1]
PH2O
Rn ) (n g 2) (A52)
PH22PCO
k6K′3 [s1] + k7PH2 + k8+(1 - βn)
PH2O

PH2PCO
k6K′3 [s ]
PH2O 1
R1 ) (n ) 1) (A53)
PH2PCO
k6K′3 [s ] + k7
PH2O 1

k6K′3PH22PCO/PH2O[s1]
RA ) (A54)
k6K′3PH22PCO/PH2O[s1] + k7PH2 + k8+

- +

{ [
βn ) (k8 /k8 ) PCnH2n/ R1RA n-1
+
2
k8-

k6K′3PCOPH2 /PH2O[s1] + k7PH2 + k8 +



i)2
n
(RAi-2PC (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.

(
PH2O
RCH4 ) k7MK0PH2R1/ 1 + x K0PH2 + +
K6 K0PH2x

( ) )
PH2O 1 1 1 n i 2

+ + + +1 xK P ∑(∏R )0 H2 j (A56)
x
K2K3K4K5K02PH22 K3K4K5K0PH2 K5 i)1 j)1
K4K5 K0PH2

(
n PH 2 O
RCnH2n+2 ) k7K0PH2 ∏ Rj / 1 + xK0PH 2
+ +
j)1
x
K6 K0PH2

( ) )
PH2O 1 1 1 n i 2

+ + + +1 xK P ∑(∏R )
0 H2 j (n g 2) (A57)
x
K2K3K4K5K02PH22 K3K4K5K0PH2 K5 i)1 j)1
K4K5 K0PH2

[
n PH 2 O
x
RCnH2n ) k8+(1 - βn) K0PH2 ∏ Rj / 1 + xK0PH2 + +
j)1
x
K6 K0PH2

( ) ]
PH2O 1 1 1 n i
+ + + +1 xK0PH2 ∑ (∏Rj) (n g 2) (A58)
x
K2K3K4K5K02PH22 K3K4K5K0PH2 K5 i)1 j)1
K4K5 K0PH2

k1PCO
Rn ) (n g 2) (A59)
k1PCO + k7xK0PH2[s1] + k8+(1 - βn)

k1PCO
R1 ) (A60)
k1PCO + k7xK0PH2[s1]

k1PCO
RA ) (A61)
k1PCO + k7xK0PH2[s1] + k8+

{ [ k8-

]}
n
- +
βn ) (k8 /k8 ) PCnH2n/ R1RA n-1
+ ∑ (RAi-2PC (n-i+2)H2(n-i+2)
) (n g 2) (A62)
x
k1PCO + k7 K0PH2[s1] + k8 + i)2

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

(
PH2O
RCH4 ) k7MK0PH2R1/ 1 + x K0PH2 + +
K6 K0PH2x

( )x )
n i 2
1 1 1
K1PCO + + + +1 K0PH2 ∑ (∏Rj) (A63)
x
K3K4K5K0PH2 K5 i)1 j)1
K4K5 K0PH2

(
i PH 2 O
RCnH2n+2 ) k7K0PH2 ∏ Rj / 1 + x K0PH2 + +
j)1
x
K6 K0PH2

( )x )
n i 2
1 1 1
K1PCO + + + +1 K0PH2 ∑ (∏Rj) (n g 2) (A64)
x
K3K4K5K0PH2 K5 i)1 j)1
K4K5 K0PH2
5086 Ind. Eng. Chem. Res., Vol. 42, No. 21, 2003

[
i PH2O
x
RCnH2n ) k8+(1 - βn) K0PH2 ∏ Rj / 1 + x K0PH2 + +
j)1
x

]
K6 K0PH2

( )x
n i
1 1 1
K1PCO + + + +1 K0PH2 ∑ ∏Rj)
( (n g 2) (A65)
x
K3K4K5K0PH2 K5 i)1 j)1
K4K5 K0PH2

k2K1xK0PH2PCO[s1]
Rn ) (n g 2) (A66)
k2K1xK0PH2PCO[s1] + k7xK0PH2[s1] + k8+(1 - βn)

k2K1xK0PH2PCO
R1 ) (n ) 1) (A67)
k2K1xK0PH2PCO + k7xK0PH2

k2K1xK0PH2PCO[s1]
RA ) (A68)
k2K1xK0PH2PCO[s1] + k7xK0PH2[s1] + k8+

{ [ k8-

]}
n
- +
βn ) (k8 /k8 ) PCnH2n / R1RA n-1
+ ∑(RAi-2PC (n-i+2)H2(n-i+2)
) (n g 2)
x
k2K1 K0PH2PCO[s1] + k7 K0PH2[s1] + k8 x + i)2

(A69)

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

PH2PCO 1 PH2O PH 2 O PH21.5PCO


RCH4 ) k9MK6PH2R1/ 1 + xK P 6 H2 + K1PCO + K2 1
+ + + K′3 +
x
PH2O K6K7K8 PH2 PH2O

)
K8 K6PH2
2
PH22PCO n i
K′4
PH2O
+ xK P ∑(∏R ) 6 H2
i)1 j)1
j (A70)

RCnH2n+2 ) k9K6PH2 ∏
i

(
Rj / 1 + xK6PH 2
+ K1PCO + K2 1
PH2PCO

PH2O
+
1
K6K7K8 PH2
PH2O
+
PH 2 O

x
+ K′3
PH21.5PCO

PH2O
+

)
j)1
K8 K6PH2
2 2
PH2 PCO n i
K′4
PH2O
+ xK P ∑(∏R )
6 H2
i)1 j)1
j (n g 2) (A71)

x
RCnH2n ) k10+(1 - βn) K6PH2 ∏
i

j)1 [
Rj / 1 + x K6PH2 + K1PCO + K21
PH2PCO

PH2O
+
1
K6K7K8 PH2
PH2O
+
PH 2 O

x
+

]
K8 K6PH2
1.5 2
PH2 PCO PH2 PCO n i
K′3
PH2O
+ K′4
PH2O
+ x K6PH2 ∑ ∏Rj)
i)1 j)1
( (n g 2) (A72)

k5K′4PH22PCO/PH2O
Rn ) (n g 2) (A73)
k5K′4PH22PCO/PH2O + k9xK6PH2 + k10+(1 - βn)/[s1]

k5K′4PH21.5PCO/PH2O
R1 ) (n ) 1) (A74)
k5K′4PH21.5PCO/PH2O + k9xK6
Ind. Eng. Chem. Res., Vol. 42, No. 21, 2003 5087

k5K′4PH22PCO/PH2O[s1]
RA ) (A75)
k5K′4PH22PCO/PH2O[s1] + k9xK6PH2[s1] + k10+

{ [ k10-

]}
n
- +
βn ) (k10 /k10 ) PCnH2n/ R1RA n-1
+ ∑(RAi-2PC (n-i+2)H2(n-i+2)
) (n g 2)
2
k5K′4PH2 PCO/PH2O[s1] + k9 K6PH2[s1] + k10 x + i)2

(A76)

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

RCH4 ) k10K9K6 K′4 0.5


PH22.5PCO

PH2O
/ 1+
( xK P 6 H2 + K1PCO + K2 1
PH2PCO

PH2O
+ K2 2
PH2O

PH2
+
x
PH 2 O
+ K′3
PH21.5PCO

PH2O
+

)
K8 K6PH2
2 2
PH2 PCO PH22PCO n i
K′4
PH2O x
+ (1 + K9 K6PH2)K′4
PH2O
∑ ∏ Rj
i)2 j)2
(A77)

RCnH2n+2 ) k10K9K6K′4PH23PCO/PH2O ∏
i)2
n

(
Ri / 1 + x K6PH2 + K1PCO + K21
PH2PCO

PH2O
+ K22
PH2O

P H2
+
x
PH 2 O
+

)
K8 K6PH2
1.5 2 2 2
PH2 PCO PH2 PCO PH2 PCO n i
K′3
PH2O
+ K′4
PH2O
+ (1 + K9 K6PH2)K′4x PH2O
∑ ∏ Rj
i)2 j)2
(n g 2) (A78)

+ 2
RCnH2n ) k11 (1 - βn)K′4PH2 PCO/PH2O ∏
n

i)2
Ri /
[ 1+ xK P 6 H2 + K1PCO + K2 1
PH2PCO

PH2O
+ K2 2
PH2O

PH2
+
x
PH 2 O

K8 K6PH2
+

]
1.5 2 2
PH2 PCO PH2 PCO PH2 PCO n i
K′3
PH2O
+ K′4
PH2O
+ (1 + K9 K6PH2)K′4x PH2O
∑ ∏ Rj
i)2 j)2
(n g 2) (A79)

βn )

- +

{ [ 2
(k11 /k11 ) PCnH2n/ K′4PH2 PCO/PH2ORA n-1
+
2
k11-

k5K′4PH2 PCO/PH2O[s1] + k10K9K6PH2[s1] + k11 +



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

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

k5K′4PH22PCO/PH2O[s1]
RA ) (A82)
k5K′4PH22PCO/PH2O[s1] + k10K9K6PH2[s1] + k11+

where K′4 ) K1K2K3K4K62K7K8.


Model FT X: RDS, steps 4, 8, and 9.

0.5 2.5
RCH4 ) k8MK7K5 K′3PH2 PCO/PH2O/ 1 +
[ xK P 5 H2 + K1PCO + K2 1
PH2PCO

PH2O
+ K2 2
PH2O

P H2
+ K7K′3K5 0.5
PH22.5PCO

PH2O
+

K′3
PH22PCO

PH2O x
+ (1 + K7 K5PH2)K′3PH22PCO/PH2O ∑
n

∏Rj
i)2 j)2
i

] (A83)
5088 Ind. Eng. Chem. Res., Vol. 42, No. 21, 2003

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

[
Ri/ 1 + xK5PH2 + K1PCO + K21
PH2PCO

PH2O
+ K22
PH2O

PH2
+

]
i)2

PH22.5PCO PH22PCO n i
K7K′3K5 0.5

PH2O
+ K′3
PH2O x
+ (1 + K7 K5PH2)K′3PH22PCO/PH2O ∑ ∏ Rj
i)2 j)2
(n g 2) (A84)

RCnH2n ) k9+(1 - βn)K′3PH22PCO/PH2O


n

∏ [
Ri/ 1 + xK5PH 2
+ K1PCO + K21
PH2PCO

PH2O
+ K22
PH2O

PH2
+

]
i)2

PH22.5PCO PH22PCO n i
K7K′3K50.5
PH2O
+ K′3
PH2O x
+ (1 + K7 K5PH2)K′3PH22PCO/PH2O ∑ ∏ Rj
i)2 j)2
(n g 2) (A85)

k4K′3PH22PCO/PH2O
Rn ) (n g 2) (A86)
k4K′3PH22PCO/PH2O + k8K7K5PH2 + k9+(1 - βn)/[s1]

PH22PCO
k4K′3 [s1]
PH2O
RA ) (A87)
PH22PCO
k4K′3 [s1] + k8K7K5PH2[s1] + k9+
PH2O

{[ ]}
βn ) (k9-/k9+) ×
PH22PCO k9- n
PCnH2n/ K′3
PH2O
RAn-1
[s1] + [s1] ∑ RAi-2PC (n-i+2)H2(n-i+2)
(n g 2)
PH22PCO i)2

k4K′3 [s1] + k8K7K5PH2[s1] + k9+


PH2O
(A88)

Nomenclature K1 ) equilibrium constant of the elementary reaction 1 for


FTS (bar-1)
E5 ) activation energy for chain growth (kJ mol-1) K2 ) equilibrium constant of the elementary reaction 2 for
E7M ) activation energy for methane formation (kJ mol-1) FTS (bar-1)
E7 ) activation energy for paraffin formation (kJ mol-1) K3 ) equilibrium constant the elementary reaction 3 for
E8 ) activation energy for olefin formation (kJ mol-1) FTS
Ev ) activation energy for the WGS reaction (kJ mol-1) K4 ) equilibrium constant the elementary reaction 4 for
fi,obj ) objective function for the ith response FTS (bar-1)
FTot,obj ) multiresponse objective function K6 ) equilibrium constant the elementary reaction 6 for
kWGS4 ) rate constant of the forward reaction in step IV FTS
(mol g-1 s-1) Kp ) equilibrium constant of the WGS reaction
k-WGS4 ) rate constant of the reversible reaction in step Kv ) group of constants of the WGS reaction (bar-0.5)
IV (mol g-1 s-1) KWGS1 ) equilibrium constant of CO adsorption elementary
k5 ) rate constant of chain growth (mol g-1 s-1 bar-1) step
k5,0 ) preexponential factor of chain growth (mol g-1 s-1 KWGS2 ) equilibrium constant of H2O adsorption elemen-
bar-1) tary step
k7M ) rate constant of methane formation (mol g-1 s-1 KWGS3 ) equilibrium constant of surface reaction elemen-
bar-1) tary step
k7M,0 ) preexponential factor of methane formation (n g KWGS4 ) equilibrium constant of CO2 desorption elementary
2) (mol g-1 s-1 bar-1) step
k7 ) rate constant of paraffin formation (mol g-1 s-1 bar-1) KWGS5 ) equilibrium constant of H2 desorption elementary
k7,0 ) preexponential factor of paraffin formation (n g 2) reaction
(mol g-1 s-1 bar-1) mi ) mole flow rate of component i (mol s-1)
k8 ) rate constant of olefin formation (mol g-1 s-1) MARR ) mean absolute relative residuals
k8,0 ) preexponential factor of olefin formation (n g 2) (mol N ) maximum carbon number of the hydrocarbons in-
g-1 s-1) volved
kv ) rate constant of CO2 formation (mol g-1 s-1 bar-1.5) Nc ) total number of components involved
kv,0 ) preexponential factor of CO2 formation (mol g-1 s-1 NR ) total number of reactions involved
bar-1.5) Nexp ) total number of experiments
k-8 ) rate constant of olefin readsorption reaction (mol g-1 Nresp ) total number of responses for parameter estimiza-
s-1 bar-1) tion
Ind. Eng. Chem. Res., Vol. 42, No. 21, 2003 5089

Pi ) partial pressure of component i (bar) (12) Lox, E. S.; Froment, G. F. Kinetics of the Fischer-Tropsch
PT ) total pressure of the reaction system (bar) reaction on a precipitated promoted iron catalyst. 2. kinetic
R ) gas constant (J mol-1 K-1) modeling. Ind. Eng. Chem. Res. 1993, 32, 71.
(13) Wang, Y.-N.; Ma, W.-P.; Lu, Y.-J.; Yang, J.; Xu, Y.-Y.;
Rj ) overall reaction rate of reaction path j (mol g-1 s-1)
Xiang, H.-W.; Li, Y.-W.; Zhao, Y.-L.; Zhang, B.-J. Kinetics
Ri ) rate of formation of component i (mol g-1 s-1) modeling of Fischer-Tropsch synthesis over an industrial Fe-
RR ) relative residual between the experimental and Cu-K catalyst. Fuel 2003, 82, 195.
calculated values of the response (14) Ponec, V. In Handbook of Heterogeneous Catalysis;
s1 ) active site for hydrocarbon formation Knoezinger, G., Ertl, H., Weitkamp, J., Eds.; VCH: Weinheim,
s2 ) active site for the WGS reaction Germany, Vol. 4, pp 1879-1883.
T ) temperature of the reaction system (K) (15) Herington, E. F. G. The Fischer-Tropsch synthesis con-
sidered as a polymerization reaction. Chem. Ind. (London) 1946,
Wi ) weight of the ith response
65, 346.
W ) weight of the catalyst used (g) (16) Friedel, R. A.; Anderson, R. B. Composition of Synthetic
Greek Symbols Liquid Fuels, I. Product Distribution and Analysis of C5-C8
Paraffin Isomers from Cobalt Catalyst. J. Am. Chem. Soc. 1950,
R1 ) chain growth factor for carbon number 1 72, 1212.
Rn ) chain growth factor for carbon number n (n g 2) (17) Huff, G. A.; Satterfield, C. N. Liquid accumulation in a
RA ) chain growth factor without olefin readsorption Fischer-Tropsch fixed bed reactor. Ind. Eng. Chem., Process Des.
Ri,j ) stoichiometric coefficient for the ith component in the Dev. 1985, 24, 986.
(18) Wojciechowski, B. W. The kinetics of the Fischer-Tropsch
jth reaction
synthesis. Catal. Rev.-Sci. Eng. 1988, 30, 629.
βn ) readsorption factor of 1-olefin with carbon number n (19) (a) Novak, S.; Madon, R. J.; Suhl, H. Models of hydrocarbon
(n g 2) product distributions in Fischer-Tropsch synthesis. J. Chem.
Phys. 1981, 74 (11), 6083. (b) Novak, S.; Madon, R. J.; Suhl, H.
Superscripts and Subscripts Secondary effects in the Fischer-Tropsch synthesis. J. Catal.
cal ) calculated value 1982, 77, 141.
exp ) experimental value (20) Iglesia, E.; Soled, S. L.; Baumgartner, J. E.; Reyes, S. C.
i ) index indicating reactions Synthesis and catalytic properties of eggshell cobalt catalysts for
the Fischer-Tropsch synthesis. J. Catal. 1995, 153, 108.
j ) index indicating components
(21) Pichler, H.; Schultz, H.; Elstner, M. Gesetzmassigheiten
M ) methane bei der synthese von kohlenwasserstoffen aus kohlenoxide und
n ) number of carbon atoms wasserstoff (Some laws of the synthesis of hydrocarbons from
* ) equilibrium state carbon oxide and hydrogen). Brennst. Chem. 1967, 48, 78.
(22) Satterfield, C. N.; Huff, G. A. Carbon number distribution
Literature Cited of Fischer-Tropsch products formed on an iron catalyst in a slurry
reactor. J. Catal. 1982, 73, 187.
(1) Schulz, H. Short history and present trends of Fischer- (23) König, L.; Gaube, J. Fischer-Tropsch-synthesis, recent
Tropsch synthesis. Appl. Catal. A 1999, 186, 3. investigations and developments. Chem.-Ing.-Tech. 1983, 55, 14.
(2) (a) Dry, M. E. The Fischer-Tropsch process: 1950-2000. (24) Satterfield, C. N.; Huff, G. A.; Longwell, J. P. Product
Catal. Today 2002, 71, 227. (b) Dry, M. E. Fischer-Tropsch distribution from iron catalysts in Fischer-Tropsch slurry reac-
reactions and the environment. Appl. Catal. A 1999, 189 (2), 185. tors. Ind. Eng. Chem., Process Des. Dev. 1982, 21 (3), 465.
(3) Dry, M. E. High quality diesel via the Fischer-Tropsch (25) Egiebor, N. O.; Cooper, W. C.; Wojciechowski, B. W. Carbon
processsa review. J. Chem. Technol. Biotechnol. 2001, 77, 43. number distribution of Fischer-Tropsch CO hydrogenation prod-
(4) (a) van der Laan, G. P.; Beenackers, A. A. C. M. Intrinsic ucts from precipitated iron catalysts. Can. J. Chem. Eng. 1985,
kinetics of the gas-solid Fischer-Tropsch and water gas shift 63, 826.
reactions over a precipitated iron catalyst. Appl. Catal. A 2000, (26) Huff, G. A.; Satterfield, C. N. Evidence for two chain
193, 39. (b) van der Laan, G. P. Kinetics, selectivity and scale up growth probabilities on iron catalysts in Fischer-Tropsch syn-
of the Fischer-Tropsch synthesis. Ph.D. Dissertation, University thesis. J. Catal. 1984, 85, 370.
of Groningen, Groningen, Germany, 1999. (27) Jordan, D. S.; Bell, A. T. Catalytic reactions of carbon
(5) Liu, Z.-T.; Li, Y.-W.; Zhou, J.-L.; Zhang, B.-J. Intrinsic monoxide. J. Phys. Chem. 1986, 90, 4797.
kinetics of Fischer-Tropsch synthesis over an Fe-Cu-K catalyst. (28) Iglesia, E.; Reyes, S. C.; Madon, R. J. Selectivity control
J. Chem. Soc., Faraday Trans. 1995, 91 (18), 3255. and catalyst design in the Fischer-Tropsch synthesis. sites,
(6) (a) Zimmerman, W. H.; Bukur, D. B. Reaction kinetics over pellets, and reactors. In Advances in catalysis and related subjects;
iron catalysts used for Fischer-Tropsch synthesis. Can. J. Chem. Eley, D. D., Weisz, P. B., Pines, H., Eds.; Academic Press: New
Eng. 1990, 68, 292. (b) Zimmerman, W. H. Kinetic modeling of York, 1993; Vol. 39, p 221.
the Fischer-Tropsch synthesis. Ph.D. Thesis, TAMU, 1990. (29) Kuipers, E. W.; Vinkenburg, I. H.; Oosterbeek, H. Chain
(7) Anderson, R. B. Catalysis for the Fischer-Tropsch synthesis; length dependence of R-olefin readsorption in Fischer-Tropsch
Van Nostrand Reinhold: New York, 1956; Vol. 4, p 29. synthesis. J. Catal. 1995, 152 (1), 137.
(8) Dry, M. E. Advances in Fischer-Tropsch chemistry. Ind. (30) Iglesia, E.; Reyes, S. C.; Madon, R. J. Transport enhanced
Eng. Chem., Prod. Res. Dev. 1976, 15, 282. R-olefin readsorption pathways in Ru-catalyzed hydrocarbon
(9) (a) Sarup, B.; Wojciechowski, B. W. Studies of the Fischer- synthesis. J. Catal. 1991, 129, 238.
Tropsch Synthesis on a Cobalt Catalyst. I. Evaluation of Product (31) Zimmerman, W. H.; Bukur, D.; Ledakowicz, S. Kinetic
Distribution Parameters from Experimental Data. Can. J. Chem. Model of Fischer-Tropsch Synthesis Selectivity in the Slurry
Eng. 1988, 66, 831. (b) Sarup, B.; Wojciechowski, B. W. Studies of Phase. Chem. Eng. Sci. 1992, 47, 2707.
the Fischer-Tropsch Synthesis on a Cobalt Catalyst. II. Kinetics (32) Kuipers, E. W.; Scheper, C.; Wilson, J. H.; Vinkenburg,
of Carbon Monoxide Conversion to Methane and to Higher I. H.; Oosterbeek, H. Non-ASF Product Distributions Due to
Hydrocarbons. Can. J. Chem. Eng. 1989, 67, 62. (c) Sarup, B.; Secondary Reactions during Fischer-Tropsch Synthesis. J. Catal.
Wojciechowski, B. W. Studies of the Fischer-Tropsch Synthesis 1996, 158 (1), 288.
on a Cobalt Catalyst. III. Mechanistic Formulation of the Kinetics (33) Schulz, H.; Claeys, M. Kinetic modelling of Fischer-
of Selectivity for Higher Hydrocarbon Formation. Can. J. Chem. Tropsch product distribution. Appl. Catal. A 1999, 186 (1-2), 91.
Eng. 1989, 67, 620. (34) Overett, M. J.; Hill, R. O.; Moss, J. R. Organometallic
(10) van der Laan, G. P.; Beenackers, A. A. C. M. Kinetics and chemistry and surface science: mechanistic models for the
selectivity of the Fischer-Tropsch synthesis: A literature review. Fischer-Tropsch synthesis, Coord. Chem. Rev. 2000, 206-207,
Catal. Rev.-Sci. Eng. 1999, 41 (3-4), 255 and the literature herein. 581.
(11) Lox, E. S.; Froment, G. F. Kinetic of the Fischer-Tropsch (35) Ponec, V. Some aspects of the mechanism of methanation
reaction on a precipitated promoted iron catalyst. 1. experimental and Fischer-Tropsch synthesis. Catal. Rev.-Sci. Eng. 1978, 18 (1),
procedure and results. Ind. Eng. Chem. Res. 1993, 32, 61. 151.
5090 Ind. Eng. Chem. Res., Vol. 42, No. 21, 2003

(36) Brady, R.; Pettit, R. Reactions of diazomethane on transi- (51) Ponec, V.; Van Barneveld, W. A. The role of chemisorption
tion-metal surfaces and their relationship to the mechanism of in Fischer Tropsch synthesis. Ind. Eng. Chem., Process Des. Des.
the Fischer-Tropsch reaction. J. Am. Chem. Soc. 1980, 102, 6181. 1979, 4, 268.
(37) Martinez, J. M.; Adams, H.; Bailey, N. A.; Maitlis, P. M. (52) Li, S.; Meitzner, G. D.; Iglesia, E. Structure and site
The Coupling of Vinyl and µ-Methylene Ligands: A New View of evolution of iron oxide catalyst precursors during the Fischer-
the Mechanism of the Fischer-Tropsch Polymerisation Reaction. Tropsch synthesis. J. Phys. Chem. B 2001, 105, 5743.
J. Chem. Soc., Chem. Commun. 1989, 5, 286. (53) (a) Jiang, M.; Koizumi, N.; Yamada, M. Adsorption proper-
(38) Joyner, R. W. The mechanism of chain growth in the ties of iron-manganese catalysts investigated by in-situ diffuse
Fischer-Tropsch hydrocarbon synthesis. Catal. Lett. 1988, 1, 307. reflectance FTIR spectroscopy. J. Phys. Chem. B 2000, 104, 7636.
(39) (a) Bai, L.; Xiang, H.-W.; Li, Y.-W.; Han, Y.-Z.; Zhong, B. (b) Bian, G.; Oonuki, A.; Kobayashi, Y.; Koizumi, N.; Yamada, M.
Slurry phase Fischer-Tropsch synthesis over manganese-pro- Syngas adsorption on precipitated iron catalysts reduced by H2,
moted iron ultrafine particle catalyst. Fuel 2002, 81, 1577. (b) syngas, or CO and on those used for high-pressure FT synthesis
Zhong, B.; Wang, Q.; Peng, S. Y. Chinese Patent ZL95106156.9, by in-situ diffuse reflectance FTIR spectroscopy. Appl. Catal. A
1995. 2001, 219, 13.
(40) Itoh, H.; Nagano, S. Catalytic properties and crystalline (54) Dry, M. E. Practical and theoretical aspects of the catalytic
structures of manganese-promoted iron ultrafine particles for Fischer-Tropsch process. Appl. Catal. A 1996, 138, 319.
liquid-phase hydrogenation of carbon monoxide. Appl. Catal. A (55) Huff, G. A.; Satterfield, C. N. Intrinsic kinetics of the
1993, 96, 125. Fischer-Tropsch synthesis on a reduced fused-magnetite catalyst.
(41) Li, X. G.; Zhong, B. Fischer-Trpoch synthesis on Fe-Mn Ind. Eng. Chem., Process Des. Dev. 1984, 23, 696.
ultrafine catalysts. Catal. Lett. 1994, 23, 245. (56) Bell, A. T. Catalytic synthesis of hydrocarbons over group
(42) (a) Maiti, G. C.; Malessa, R.; Baerns, M. Iron/manganese VIII metals. A discussion of the reaction mechanism. Catal. Rev.-
oxide catalysts for Fischer-Tropsch synthesis. Part I: structural Sci. Eng. 1981, 23 (1-2), 203.
and textural changes by calcination, reduction, and synthesis. (57) Turner, M. L.; Marsih, N.; Mann, B. E.; Quyoum, R.; Long,
Appl. Catal. 1983, 5, 151. (b) Maiti, G. C.; Malessa, R.; Löchner, H. C.; Maitlis, P. M. Investigations by 13C NMR spectroscopy of
U.; Papp, H.; Baerns, M. Iron/manganese oxide catalysts for ethene-initiated catalytic CO hydrogenation. J. Am. Chem. Soc.
Fischer-Tropsch synthesis. Part II: crystal phase composition, 2002, 124, 10456.
activity and selectivity. Appl. Catal. 1985, 16, 215. (c) Löchner, (58) Vannice, M. A. Synthesis of hydrocarbons from CO and
U.; Papp, H.; Baerns, M. Iron/manganese oxide catalysts for H2. Catal. Rev.-Sci. Eng. 1976, 14 (2), 153.
Fischer-Tropsch synthesis. Part III: Phase changes in iron/
(59) Hanlon, R. T.; Satterfield, C. N. Reactions of selected
manganese oxide Fischer-Tropsch catalysts during start-up and
1-olefins and ethanol added during the Fischer-Tropsch synthesis.
synthesis process. Appl. Catal. 1986, 23, 339. (d) Malessa, R.;
Energy Fuels 1988, 2, 196.
Baerns, M. Iron/manganese oxide catalysts for Fischer-Tropsch
(60) Newsome, D. S. The water-gas shift reaction. Catal. Rev.-
synthesis. 4: activity and selectivity. Ind. Eng. Chem. Res. 1988,
Sci. Eng. 1980, 21, 275.
27, 279. (e) Grzybek, T.; Papp, H.; Baerns, M. Iron/manganese
oxide catalysts for Fischer-Tropsch synthesis. Part V: XPS (61) Rethwisch, D. G.; Dumesic, J. A. Adsorptive and catalytic
surface characterization of calcined and reduced samples. Appl. properties of supported metal oxides. III. Water gas shift over
Catal. 1987, 29, 335. (f) Grzybek, T.; Papp, H.; Baerns, M. Iron/ supported iron and zinc oxides. J. Catal. 1986, 101, 35.
manganese oxide catalysts for Fischer-Tropsch synthesis. Part (62) Ovesen, C. V.; Stoltze, P.; Norskov, J. K.; Campbell, C. T.
VI: surface characterization during start-up period and pseudo- A Kinetic Model of the Water-Gas Shift Reaction. J. Catal. 1992,
steady-state synthesis. Appl. Catal. 1987, 29, 351. 134, 445.
(43) (a) Burkur, D. B.; Lang, X.; Ding, Y. Pretreatment effect (63) Ovesen, C. V.; Clausen, B. S.; Hammershøi, B.; Askgård,
studies with a precipitated iron Fischer-Tropsch catalyst in a T.; Chorkendorff, I.; Nørskov, J. K.; Rasmussen, P. B.; Stoltze, P.;
slurry reactor. Appl. Catal. A 1999, 186, 255. (b) Mansker, L. D.; Taylor, P. A. Kinetics of the water-gas shift reaction at high
Jin, Y.; Burkur, D. B.; Datye, A. K. Characterization of slurry pressures. J. Catal. 1996, 158, 170.
phase iron catalysts for Fischer-Tropsch synthesis. Appl. Catal. (64) Graaf, G. H.; Winkelman, J. G. M.; Stamhuis, E. J.;
A 1999, 186, 277. Beenackers, A. A. C. M.; Kinetics of the three phase methanol
(44) (a) Jin, Y.; Datye, A. K. Phase transformations in iron synthesis. Chem. Eng. Sci. 1988, 43, 2161.
Fischer-Tropsch catalysts during temperature programmed re- (65) Gear, G. W. Numerical initial value problem in ordinary
duction. J. Catal. 2000, 196, 8. (b) Sault, A. G. An Auger electron differential equations; Prentice Hall: Englewood Cliffs, NJ, 1971.
spectroscopy study of the activation of iron Fischer-Tropsch (66) Park, T. Y.; Froment, G. F. A hybrid genetic algorithm for
catalysts. Part I: Hydrogen activation. J. Catal. 1993, 140, 121. the estimation of the parameters in detailed kinetics models.
(c) Sault, A. G.; Datye, A. K. An Auger electron spectroscopy study Comput. Chem. Eng. 1998, 22, S103.
of the activation of iron Fischer Tropsch catalysts: Part II: Carbon (67) Han, R. F.; Zhang, Y. K.; Wang, Y. N.; Xu, Y. Y.; LI, Y. W.
monoxide activation. J. Catal. 1993, 140, 136. Optimization of parameters of FTS kinetic model by genetic
(45) Itoh, H.; Nagano, S.; Takeda, K.; Kikuchi, E. Catalytic algorithm. J. Fuel Chem. Technol. (China) 2001, 29 (4), 371.
properties and crystalline structures of manganese-promoted (68) Ji, Y.-Y.; Xiang, H.-W.; Yang, J.-L.; Xu, Y.-Y.; Li, Y.-W.;
iron ultrafine particles for liquid-phase hydrogenation of carbon Zhong, B. Effect of reaction conditions on the product distribution
monoxide. Appl. Catal. A 1993, 186, 125. during Fischer-Tropsch synthesis over an industrial Fe-Mn
(46) (a) Liu, Z.-P.; Hu, P. A new insight into Fischer-Tropsch catalyst. Appl. Catal. A 2001, 214 (1), 77.
synthesis. J. Am. Chem. Soc. 2002, 124, 11568. (b) Ciobı̂cã, I. M. (69) Dictor, R. A.; Bell, A. T. Fischer-Tropsch synthesis over
The molecular basis of the Fischer-Tropsch reaction. Ph.D. Thesis, reduced and unreduced iron oxide catalysts. J. Catal. 1986, 97,
Technical University Eindhoven, Eindhoven, The Netherlands, 121.
2002. (70) Feimer, J. L.; Silveston, P. L.; Hudgins, R. R. Steady-state
(47) Lox, E. S.; Marin, G. B.; de Graeve, E.; Bussier, P. study of the Fischer-Tropsch reaction. Ind. Eng. Chem., Prod. Res.
Characterization of promoted precipitated iron catalyst for Dev. 1981, 20, 609.
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 Received for review February 14, 2003
iron catalyst for Fischer-Tropsch synthesis by in-situ laser Raman Revised manuscript received June 12, 2003
spectroscopy. J. Catal. 1985, 95, 325. Accepted June 13, 2003
(50) Adesina, A. A. Hydrocarbon synthesis via Fischer-Tropsch
reaction: travails and triumphs. Appl. Catal. A 1996, 138, 345. IE030135O

You might also like