You are on page 1of 13

Article

pubs.acs.org/IECR

Effects of Catalyst Activity, Particle Size and Shape, and Process


Conditions on Catalyst Effectiveness and Methane Selectivity for
Fischer−Tropsch Reaction: A Modeling Study
Miloš Mandić,† Branislav Todić,† Ljiljana Ž ivanić,†,∥ Nikola Nikačević,‡ and Dragomir B. Bukur*,†,§

Chemical Engineering Program, Texas A&M University at Qatar, P.O. Box 23874, Doha, Qatar

Faculty of Technology and Metallurgy, University of Belgrade, Karnegijeva 4, Belgrade 11000, Serbia
§
Artie McFerrin Department of Chemical Engineering, Texas A&M University, MS 3122, College Station, Texas 77843-3122, United
States

ABSTRACT: We investigate effects of catalyst activity, catalyst particle


shape (sphere, slab, and hollow cylinder), size (i.e., diffusion length),
catalyst distribution (uniform vs eggshell type distribution for a spherical
particle), and process conditions (temperature, pressure, syngas
composition, and conversion level) on catalyst effectiveness factor and
methane selectivity inside the catalyst pellet. In numerical simulations we
utilize kinetic parameters for CO consumption rate and CH4 formation
rate determined from experiments with a highly active Co/Re/γ-Al2O3
catalyst. It is found that the use of small spherical particles (0.2−0.5 mm)
or eggshell distribution for larger spherical particles with catalyst layer
thickness less than approximately 0.13 mm is needed to avoid negative
impact of diffusional limitations on CH4 selectivity under typical Fischer−
Tropsch synthesis operating conditions. For monolith reactors with wash-coated catalyst, diffusional limitations can be avoided
by using a catalyst layer thickness less than 0.11 mm at base case conditions (473 K, 25 bar, and H2/CO molar ratio of 2).

1. INTRODUCTION and (c) micro- and milli-fixed bed reactors in which small catalyst
Fischer−Tropsch synthesis (FTS) is a heterogeneous reaction particles are loaded into the packed bed.3 These types of reactors
used to convert synthesis gas into a range of hydrocarbon in general minimize negative effects of intraparticle diffusional
products. This reaction is a key step in the gas-to-liquid (GTL) limitations on activity and selectivity but are characterized by low
process in which natural gas is converted into liquid fuels and reactor productivity due to low amount of catalyst in the
value-added chemicals. Low-temperature FTS is conducted reactor.11,12 The latter can be compensated by using highly active
commercially in two types of reactors: slurry bubble column FTS catalysts such as those developed by Oxford Catalysts
reactors and multitubular fixed bed reactors.1−3 The overall goal Group8,13 and milli-fixed bed reactors loaded with small particles.
of fixed bed reactor design is to maximize productivity and The interplay between chemical reaction and intraparticle
selectivity to desired products (low methane and high selectivity diffusion for FTS applications (Co- and Fe-based catalysts) has
to liquid hydrocarbons−C5+ hydrocarbons) while minimizing been studied in the literature either on a single-particle scale or as
pressure drop and costs. To achieve these objectives, a judicious a part of modeling of fixed bed reactors and has received
choice for the catalyst particle size and shape, as well as process increased attention in recent years. In a majority of previous
conditions, is required. To avoid high pressure drop, the use of studies a simplified kinetics (first-order in hydrogen or nth-order
relatively large particles (1−3 mm) is required,1,2 whereas to in carbon monoxide) was utilized to calculate catalyst
achieve high productivity in a given reactor volume, one has to effectiveness factor as a function of Thiele modulus, and
use very active catalysts. These requirements lead to intraparticle selectivity inside the particle was not considered.14−22 The first
diffusional limitations resulting in lower reaction rate (decrease quantitative analysis of chemical reaction with diffusion problem
in catalyst effectiveness) and decrease in selectivity to liquid for FTS reaction was published by Dixit and Tavlarides23 who
hydrocarbons.4−8 Slurry reactors utilize small catalyst particles used a form of Langmuir−Hinshelwood (LH) kinetics derived
(less than 100 μm), and in this case, the intraparticle diffusional from experimental data with 0.5%Ru/γ-Al2O3 catalyst for FTS.24
limitations are not expected under normal operating conditions. This form of rate equation was later utilized by a number of
Microreactors for FTS have recently received a great deal of
attention because of increased heat removal at moderate pressure Received: January 5, 2017
drop.9,10 Several types of reactors have been considered for this Revised: February 20, 2017
purpose, including (a) reactors with microstructured catalysts Accepted: February 23, 2017
(e.g., monoliths and foams), (b) coated microchannel reactors, Published: February 23, 2017

© XXXX American Chemical Society A DOI: 10.1021/acs.iecr.7b00053


Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX
Industrial & Engineering Chemistry Research Article

research groups and became known as Yates and Satterfield ASF distribution. In particular, the actual CH4 selectivity is
kinetics.25 Dixit and Tavlarides23 conducted a parametric study significantly higher than that predicted from the ideal ASF
(in terms of dimensionless parameters) to investigate the distribution; selectivity of C2 hydrocarbons is lower than
relationship between the effectiveness factor (η) and Thiele predicted value; and α varies with chain length.36−38 Thus,
modulus (ϕ) and found that the η versus ϕ curve passes through predicted values of C5+ selectivity using the proposed approach
a maximum and can reach values exceeding 1. Also, they reported would not be very accurate. Also, the authors have used kinetic
that for some values of model parameters there exists a narrow parameters for syngas consumption from Yates and Satterfield25
range of ϕ values over which multiple steady-state solutions are experimental data with a catalyst different than the catalyst used
possible. Selectivity aspects were not considered in this study. to obtain the expression for variation of α with temperature and
Researchers at Exxon26−28 have introduced selectivity aspects syngas composition. In general, one should use kinetic
into mathematical modeling. Their model included continuity parameters for activity and product distribution obtained from
equations for both fluid phase (plug flow reactor) and the catalyst an experimental study with the same catalyst.
particle. Reaction rates for CO consumption and CH4 formation In this study we investigate effects of particle shape (sphere,
were described by LH kinetics, and a separate model was used for slab, and hollow cylinder), size (i.e., diffusion length), catalyst
modeling hydrocarbon selectivity based on 1-olefin readsorption distribution (uniform vs eggshell type distribution for a spherical
concept. They found that selectivity of liquid-phase hydro- particle), and process conditions (temperature, pressure, syngas
carbons (C5+ hydrocarbons) increases initially with increase in composition, and conversion level) on catalyst effectiveness
diffusional limitations due to diffusion-enhanced 1-olefin read- factor and methane selectivity inside a particle. In numerical
sorption resulting in formation of high molecular weight simulations we utilize kinetic parameters for CO consumption
hydrocarbons and then begins to decrease because of diffusional rate and CH4 formation rate determined from experiments with a
limitations by reactants, resulting in high H2/CO ratio inside the highly active Co/Re/γ-Al2O3 catalyst, which is more representa-
pellet which favors chain termination reactions that favor lower tive of current Co-based catalysts used in industrial applications
molecular weight hydrocarbons. To overcome these problems than the catalyst used by Yates and Satterfield.25 Models of
and satisfy pressure drop requirements, Iglesia et al.4 utilized an diffusion−reaction interaction are helpful in determining an
eggshell catalyst distribution where the catalyst is deposited only optimal catalyst layer thickness that maximizes catalyst
in a thin layer near the outer surface of a relatively large particle. A effectiveness and improves FTS product selectivity and provide
method of preparation of eggshell Co-based FTS catalysts was guidance for catalyst design and choice of process conditions.
described, and its beneficial effects on product selectivity were
illustrated with experimental data. 2. MODEL DESCRIPTION
This type of analysis was carried one step further by Wang et The overall stoichiometry of FTS on cobalt catalyst was
al.5 who utilized a comprehensive kinetic model for the catalyst described by the following equation:
pellet in which both the overall syngas consumption and the
hydrocarbon product distribution are unified. This model was CO + (13/6)H 2 = H 2O + (1/6)C6H14 (1)
developed from kinetic studies with Fe-based FTS catalyst29
which utilized the olefin readsorption concept. The advantages of which is representative of an average molecular weight of
eggshell catalyst distribution on product distribution have been hydrocarbons produced and H2/CO consumption ratio for
investigated through numerical simulations. experiments with 0.48%Re−25%Co/Al2O3 catalyst.31 In general,
A similar approach was used recently in analyzing performance H2/CO consumption ratio varies with product selectivity and lies
of a bench scale fixed bed reactor with Co-based FTS catalyst between 2 (production of very long chain molecules, α = 1) and 3
(15%Co/γ-Al2O3).30 A comprehensive kinetic model of Todic et (production of CH4 from syngas).
al.31 was used in numerical simulations, which included a It is assumed that the catalyst particle is filled with high
continuity equation for the catalyst particle. molecular weight hydrocarbons (wax) and that external mass-
An alternative, and numerically significantly less demanding, and heat-transfer resistances are negligible, so that concen-
approach was used by Vervloet et al.7 to study the effect of trations of species and temperature at the particle surface are the
process conditions on catalyst effectiveness and C5+ productivity same as those in the bulk. The reactants and products are
and provide guidance for optimal reactor operation. The dissolved in wax at the surface and throughout the particle. For
interplay between reaction and diffusion in a single catalyst steady-state conditions, assuming Fick’s law of diffusion and
particle was analyzed using the LH rate expression of Yates and isothermal pellet, the general reaction−diffusion continuity
Satterfield25 for CO consumption rate and a variable chain equation is given as
growth parameter (α) dependent on temperature and syngas
1 d ⎛ g dyi ⎞
composition (H2/CO ratio). The same approach (the use of · ⎜x · ⎟ + υi ·(ϕ′i )2 ·R̅ CO = 0
Yates and Satterfield25 kinetic parameters and variable α x g dx ⎝ dx ⎠ (2)
parameters from Vervloet et al.7) was used by Becker et al.6,32
as well as Gardezi and Joseph.33 There are some issues associated where x is the dimensionless distance and superscript g stands for
with the approach used by Vervloet et al.7 Variation of α with geometry (g = 2, 1, and 0 for sphere, cylinder and slab,
temperature and H2/CO ratio was determined from exper- respectively); yi (Csi /CsCO) is dimensionless concentration (i =
imental data of De Deugd34 for methane selectivity. Local α CO, H2, H2O, and n-hexane), νi stoichiometric coefficient (νi =
values were calculated from reactant concentrations in the −1, −13/6, 1, and 1/6 for CO, H2, H2O, and n-hexane,
particle, and average value of α was obtained by numerical respectively), R̅ CO the dimensionless CO consumption rate
integration. Selectivities of C1−C4 and C5+ hydrocarbons were (R C O /R Cs O ), and ϕ′ i the dimensionless parameter
s
then calculated assuming the ideal Anderson−Schulz−Flory ρp ·(−R CO)
(ϕ′i = Lg ). This parameter includes catalyst density
(ASF) distribution.28,35,36 However, it is well-known that actual
s
De ,i ·CCO
product distribution on Co FT catalysts deviates from the ideal (ρp), geometric characteristic length (Lg), surface concentration
B DOI: 10.1021/acs.iecr.7b00053
Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX
Industrial & Engineering Chemistry Research Article

of CO (CsCO) in the liquid medium, and effective diffusivity dyi


x = 0 (sphere or slab): =0
coefficient of species i (De,i). dx (2b)
The assumption of uniform temperature distribution in a
pellet (isothermal conditions) was verified by numerical solution x = ri /ro (hollow cylinder): yi = C is/CCO
s
(2c)
of mass and energy balance equations under a conservative set of
conditions to maximize temperature gradients inside the pellet: dyi
relatively large spherical particle (dp = 4 mm); low value of x = 1 − δ /R (eggshell distribution): =0
dx (2d)
effective thermal conductivity, keff = 0.1 W/(m·K); and high
value of heat of reaction, −173 kJ/mol. It was found that the 2.1. Kinetics. CO consumption rate per unit mass of catalyst
maximum temperature difference in the particle is 0.15 K, which is described by LH kinetics:23,25
has no impact on concentration profiles, effective diffusivity, and
k·PCO·PH2
CH4 selectivity. Vervloet et al.7 reached the same conclusion by ( −R CO) =
using relevant dimensionless groups for estimation of the (1 + a ·PCO)2 (3)
maximum ΔT in a pellet.
The second-order differential equation (eq 2) was solved for Numerical values of kinetic rate constant (k) and adsorption
several particle shapes (sphere with uniform catalyst distribution constant (a) have been reported by Yates and Satterfield25 at two
and eggshell catalyst distribution, flat plate, and hollow cylinder), temperatures (220 and 240 °C) from regression of experimental
as shown in Figure 1. The geometric characteristic length (Lg) for data with 21.4% Co/3.9% Mg/SiO2 catalyst, which will be
referred to as YS catalyst. Maretto and Krishna39 estimated
activation energy and adsorption enthalpy of k and a for YS
catalyst. Temperature dependence for these two parameters is
given by the following equations:
⎡ ⎛ 1 1 ⎞⎤
k(T ) = 8.8533 × 10−3· exp⎢4494.41⎜ − ⎟⎥
⎣ ⎝ 493.15 T ⎠⎦
mol/(s kgcat bar 2) (4)

⎡ ⎛ 1 1 ⎞⎤
a(T ) = 2.226· exp⎢ − 8236⎜ − ⎟⎥1/bar
⎣ ⎝ 493.15 T ⎠⎦ (5)

It should be noted that Yates and Satterfield25 and Maretto and


Krishna39 obtained numerical values of k and a from data for
syngas (H2 + CO) consumption rate rather than CO
consumption rate. It appears that several research groups used
eq 4 as the rate constant for CO consumption rate,6,7,33 without
correction for stoichiometry, in modeling the reaction with
diffusion in a single particle.
Figure 1. Representation of particle shapes and dimensions: (a) We have estimated parameters k and a in eq 3 from
spherical particle with uniform catalyst distribution, (b) spherical experimental data with a catalyst synthesized at Center for
particle with eggshell catalyst distribution, (c) hollow cylinder particle, Applied Energy Research (CAER) at the University of Kentucky
and (d) slab (flat plate). having composition 0.48%Re−25%Co/Al2O3. Experimental
data used to estimate kinetic parameters can be found in Table
each particle shape is as follows: radius (R) for a sphere, outer 1 of Todic et al.37 Numerical values of kinetic parameters used in
radius (ro) for a hollow cylinder, and thickness (L) for a slab (flat simulations for both CAER and YS catalysts are summarized in
plate). Dimensionless distance x, in eq 2, for different pellet Table 1. Numerical value of k in eq 4 for YS catalyst was divided
shapes is x = r/R (sphere), x = r/ro (hollow cylinder), or x = z/L by (1 + 13/6) to the obtain rate constant for CO consumption. It
(slab). is worth noting that CAER catalyst is an order of magnitude more
Dirichlet boundary conditions are used for surfaces exposed to active under typical FTS operating conditions compared to the
the surrounding fluid (x = 1), and Neumann (zero flux) YS catalyst,25 and as such is more representative of modern Co
conditions are used at x = 0 (sphere with uniform catalyst catalysts used in industrial applications.
distribution and slab) or at x = 1 − δ/R (eggshell distribution in a The key aspect of FTS catalyst selectivity is methane
spherical particle): selectivity. It is the most undesirable product, and reduction in
the amount of methane is of utmost importance. Methane
x = 1 (outer surface of the particle, all four geometries): yi = C is/CCO
s
formation rate was calculated using a recently proposed
(2a) expression:40

Table 1. Kinetic Parameters for Rate of CO Disappearance According to Equation 3

Yates and Satterfield (YS) CAER


catalyst 21.4% Co/3.9% Mg/SiO2 25% Co/0.48% Re/Al2O3
k (mol/(s·kg·bar2)) 26·exp(−4494.4/T) 4.2 × 105·exp(−8742/T)
a (1/bar) 1.24 × 10−7·exp(8236/T) 6.45 × 10−2·exp(1295.3/T)

C DOI: 10.1021/acs.iecr.7b00053
Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX
Industrial & Engineering Chemistry Research Article

aM bM Table 2. Reaction Conditions and Parameters Used in


R CH4 = kM PCOP H2 /(1 + mM ·PH2O/PH2) (6)
Numerical Simulations
kM = AM e−EM / R gT (6a) description symbol value ref
temperature T 473−493 K
where numerical values of kinetic parameters were estimated
pressure P 20−30 bar
from the same set of experimental data as those in eq 3 (Table 1
syngas ratio in the bulk H2/CO 1.4−2.0
of Todic et al.37). Estimated values of numerical parameters in eq
catalyst porosity ε 0.5
6 are aM = −0.99, bM = 1.28, mM = 0.58, AM = 2.25 × 1011 (mol/
catalyst tortuosity τ 2
(kg·s·bar0.29)), and EM = 140 (kJ/mol).
catalyst particle density ρp 1200 kg m−3
Substituting numerical values for constants aM, bM, and mM catalyst particle diameter dp 0.085−4 mm
into eq 6, we obtain the following expression for CH4 formation catalyst layer thickness (slab) L 0.035−0.300 mm
rate: eggshell layer thickness δ 0.075−0.300 mm
kM(T) ·(PH2 /PCO)0.99 ·PH2 0.29 CO diffusivity in wax Dwax,CO (1.24−1.44) × 10−8 m2 s−1 50
R CH4 = (mol/(kg·s)) H2 diffusivity in wax Dwax,H2 (3.13−3.62) × 10−8 m2 s−1 50
1 + 0.58·(PH2O/PH2) (7) H2O diffusivity in wax Dwax,H2O (2.10−2.42) × 10−8 m2 s−1 50
Equation 7 predicts that CH4 formation rate increases with C6H14 diffusivity in wax Dwax,C6H14 (5.12−5.98) × 10−9 m2 s−1 50
temperature, partial pressure of H2, and H2/CO ratio and is Henry’s coefficient for CO HCO 345.1−356.6 bar 53
inhibited by partial pressure of water. The inhibiting effect of Henry’s coefficient for H2 HH2 442.3−472.6 bar 53
water on CH4 selectivity is well-established from studies with Henry’s coefficient for H2O HH2O 41.9−49.3 bar 53
both indigenous and externally added water,40−47 whereas the Henry’s coefficient for C6H14 HC6H14 9.88−12.3 bar 53
underlying causes for this behavior are still not completely
density of wax (n-octacosane) ρw 678.1−691.4 kg m−3 50
understood.41,48,49
2.2. Physical Properties. We consider the case in which the
catalyst particle is completely filled with hydrocarbon wax, the For normal reactions (rate decreases as conversion increases),
physical properties of which are assumed to be those of n- the effectiveness factor is ≤1, and is a measure of utilization of
octacosane (C28H58). Diffusivities of species in wax were catalyst volume for reaction. However, in our case, the CO
estimated from Akgerman’s correlation.50,51 This correlation reaction rate (eq 3) passes through a maximum before it starts
has been used previously,6,21 and its predictions are similar to decreasing with decrease in reactant concentrations, and it is
those obtained using correlations of Wang et al.5 and possible to have η > 1 for some conditions.
Makrodimitri et al.52 Solution of eq 2 provides partial pressures needed to evaluate
The effective diffusivity of the component was calculated from methane formation rate (eq 7) and calculate local methane
selectivity at any point inside the catalyst as
ε
De,i = Dwax,i R CH4
τ (8) SCH4 = ·100%
( −R CO) (11)
where ε is particle porosity and τ is tortuosity factor.
Henry’s law constants for reactants and products in n- which is valid for Co catalysts because of their low water gas shift
octacosane were calculated from Marano and Holder’s activity, i.e., low CO2 formation.
correlation.53 This correlation is based on experimental data Then, average CH4 selectivity can be obtained by integration
with fluids (different types of FTS waxes and high molecular of local reaction rates as follows:
weight hydrocarbons) and process conditions (temperature and V V
1 1
pressure) that are typical for FTS (Marano54). Liquid-phase SCH
̅ 4=
V
· ∫0 ρp R CH4 dV / ·
V
∫0 ρp ( −R CO)dV
concentration of a species i is related to partial pressure of i via
V
Henry’s law constant ∫0 R CH4 dV
=
Pi = Hi(C i /C L) (9) V
∫0 (−R CO)dV (12)
where Ci is molar concentration of species i (mol/m3), CL the
total liquid-phase concentration (mol/m3), Pi partial pressure It should be noted that CH4 formation rate (eq 7) is not
(bar), and Hi Henry’s law constant (bar). bounded, and for large H2/CO ratios it can assume a very high
The parameter values used in model simulations are value. Also, as the CO gets depleted inside the particle, RCO
summarized in Table 2. approaches zero. If either one or both of these two possibilities
2.3. Catalyst Performance Indicators. Catalyst efficiency occur, the local CH4 selectivity would tend to infinity, whereas its
and CH4 selectivity are the two key factors describing catalyst value physically cannot exceed 100%, i.e., CH4 formation rate
performance. The efficiency is expressed in terms of effectiveness cannot exceed CO consumption rate (RCH4 ≤ (−RCO)). Thus, if
factor, which represents a ratio of average reaction rate in the at some point in the pellet it is found that RCH4 > (−RCO), the
pellet divided by rate at the surface conditions: condition RCH4 = (−RCO) is imposed in calculation of local and
V average CH4 selectivities (eqs 11 and 12) from this point onward.
∫0 (−R CO)ρp dV 2.4. Numerical Methods. Solution was obtained using
η= s
V ·( −R CO)ρp (10) Matlab’s bvp4c function, which is used to solve boundary value
problems for ordinary differential equations. The function uses
where V represents the volume of the region in which the catalyst finite difference method, which implements a colocation formula.
is located. For each component, eq 2 was written as a system of two first-
D DOI: 10.1021/acs.iecr.7b00053
Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX
Industrial & Engineering Chemistry Research Article

Figure 2. Effect of catalyst activity (CAER vs YS catalyst) on (a) CO and H2 concentration, (b) H2/CO ratio, (c) CO reaction rate, and (d) local CH4
selectivity (CAER catalyst). Conditions: dp = 2 mm, T = 473 K, P = 25 bar, XCO = 0%, and H2/CO = 2.

order differential equations, resulting in a system of eight first- representative of sizes (1−3 mm) used in industrial reactors to
order differential equations with conditions specified at two avoid a high pressure drop.1,2,4 Results are shown for conditions
different points. Spacing between grid points was not fixed near the reactor inlet where the CO conversion is very small (XCO
because of the nature of bvp4c algorithm, because the method = 0%). These conditions are the least favorable for achieving low
iteratively adjusts spacing in order to achieve convergence methane selectivity because of the absence of water at the particle
tolerance. Initial guesses used depended on expected profiles. surface.
For conditions which would result in small values of ϕ′i (low Steep gradients are observed for reactant liquid-phase
diffusional limitations), boundary conditions on the surface were concentrations in the case of CAER catalyst, whereas
taken as initial guesses for all points, while for larger ϕ′i concentrations decrease slowly for YS catalyst. For CAER
conditions, guesses were chosen by trial and error. catalyst, the concentration of CO approaches zero at
Local CH4 selectivity was calculated algebraically using approximately 0.2 mm from the particle surface (Figure 2a)
previously calculated concentration profiles of CO, H2, and because of lower rate of diffusion of CO compared to H2. This in
H2O. Average CH4 selectivity and effectiveness factor were turn causes very rapid increase in H2/CO ratio inside the particle
calculated using trapezoidal rule (Matalab’s function trapz). This (Figure 2b). We show results only up to H2/CO = 20, but this
method is accurate enough because of the high density of mesh ratio tends to infinity deeper inside the particle. On the other
over which the integration is done. hand, H2/CO ratio for the YS catalyst increases slightly from 2 at
the surface of the particle to ∼2.4 at the center.
CO consumption rate for the CAER catalyst increases rapidly
3. RESULTS AND DISCUSSION
in the region near the surface to reach a maximum value at
Even though the governing differential equations were solved approximately 0.11 mm from the surface and then decreases to
numerically in their dimensionless form, we present results in essentially zero value at about 0.27 mm from the surface (Figure
terms of dimensional variables to facilitate better physical 2c). This behavior is caused by change in effective reaction order
understanding and relate them to previous studies and industrial from negative at high CO concentration to positive at low CO
applications. concentrations (see eq 3). For YS catalyst, the CO consumption
3.1. Effect of Catalyst Activity. Results from simulations rate increases slightly from the surface to the center of the
with CAER catalyst (high-activity catalyst) and YS catalyst (low- particle, which is consistent with moderate decrease in CO
activity catalyst) are shown in Figure 2 for conditions typical of concentration shown in Figure 1a. It is important to note the
industrial practice (473 K, 25 bar, H2/CO ratio of 2). Simulations difference in scale for the two ordinate axis. Consumption rate of
are for a spherical particle 2 mm in diameter, which is CO is much lower with the YS catalyst relative to the CAER
E DOI: 10.1021/acs.iecr.7b00053
Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX
Industrial & Engineering Chemistry Research Article

Figure 3. Effect of particle size on (a) CO concentration, (b) H2/CO ratio, (c) CO reaction rate, and (d) local CH4 selectivity. Conditions: T = 473 K, P
= 25 bar, XCO = 0%, and H2/CO = 2.

catalyst. The ratio of reaction rates (CAER/YS) at the catalyst were calculated assuming reaction stoichiometry described by eq
surface for these process conditions is 36. Observed differences in 1. Calculated values of the effectiveness factor were 0.46 (CAER)
concentration profiles and H2/CO ratios with the two catalysts and 1.06 (YS), whereas CH4 selectivity was 13.8%. Initial (at the
are caused by respective differences in the catalyst activity. surface) reaction rates are higher when XCO = 0%, but the impact
Intraparticle diffusional limitations are much stronger with the on effectiveness factor is small, whereas CH4 selectivity is
more active catalyst. significantly lower when XCO = 50%. The latter is due to the fact
Local CH4 selectivity (Figure 2d) increases exponentially, that water is present at the outer surface and it inhibits CH4
mirroring trends in H2/CO ratio (Figure 2b), and reaches 100% formation rate (see eq 7). However, the calculated value of CH4
at approximately 0.16 mm from the surface. This is in accordance selectivity is too high for industrial applications.
with the rate equation for CH4 formation (eq 7), which predicts The above results show that the use of YS rate equation with
increase in rate with increase in H2/CO ratio, and the definition original parameter values obtained from experiments with Co/
of local selectivity, which has the rate of CO disappearance in the Mg/SiO2 catalyst25 underestimates the extent of diffusional
denominator (eq 11). The calculated value of local CH4 resistance in the case of more active Co FTS catalysts. Some
selectivity tends to infinity with increasing distance from the researchers7,8 have simply multiplied YS rate by a constant value
surface of the particle, but physically selectivity cannot exceed (up to 10) to simulate performance of more active catalysts, but
100%; this is shown as a broken horizontal line in Figure 2d. We this approach does not account correctly for variation of rate with
do not have rate parameters for CH4 formation for the YS temperature and pressure. The ratio of reaction rates (CAER/YS
catalyst; thus, there are no results for local CH4 selectivity for this catalyst) at the surface is a function of process conditions, and it
catalyst. varied between 14.9 (at 493 K, 20 bar, and XCO = 50%) and 37 (at
The catalyst effectiveness values are 0.52 (CAER) and 1.05 473 K, 30 bar, and XCO = 0%). For each catalyst, one needs to
(YS), whereas the average CH4 selectivity, calculated from eq 12, determine the correct values of activation energy and enthalpy of
for CAER catalyst is 19.4%. The latter clearly indicates a negative adsorption experimentally, as well as reaction orders with respect
impact of intraparticle diffusional resistance on the catalyst to reactants. In the remaining sections of this paper we present
performance (high CH4 and consequently low C5+ selectivity). results using kinetic parameter values for CAER catalyst.
Simulations were performed for these process conditions and 3.2. Effect of Particle Size. Although the use of small
particle size for external conditions corresponding to location in a particles (less than 1 mm) is not feasible for a large-scale
reactor where CO conversion is 50%. Liquid-phase concen- industrial fixed bed reactors because of excessive pressure drop,
trations of species (reactants and products) at the catalyst surface this is not as significant constraint for microchannel structures
F DOI: 10.1021/acs.iecr.7b00053
Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX
Industrial & Engineering Chemistry Research Article

Figure 4. Effect of catalyst shape and catalyst distribution on (a) CO concentration, (b) H2/CO ratio, (c) CO reaction rate, and (d) local CH4 selectivity.
Conditions: L = δ = ro − ri = 0.15 mm, T = 473 K, P = 25 bar, XCO = 0%, and H2/CO = 2.

which utilize shorter lengths. The use of small particles decreases For a given set of conditions (473 K, 25 bar, H2/CO = 2, and
intraparticle transport resistances, which is essential for XCO = 0%) the use of spherical particle of 0.4 mm in diameter
maintaining low CH4 selectivity. Results of simulations (CAER would be an optimal choice because it results in high
catalyst) with small spherical particles (0.2, 0.4, and 0.6 mm) and effectiveness factor and low CH4 selectivity. Compared to the
comparison with a 2 mm spherical particle are shown in Figure 3. 0.2 mm particle, increased diffusional resistance is beneficial in
As expected, the CO concentration profile becomes steeper as that it gives a slightly higher effectiveness factor without an
the particle size increases (Figure 3a). H2/CO ratio (Figure 3b) adverse effect on CH4 selectivity. The use of 0.4 mm particles in
for smaller particles (0.2 and 0.4 mm in diameter) is close to the microchannel structure may be an upper limit for channels up to
bulk ratio of 2, whereas it increases rapidly for larger particles (0.6 about 5 mm, whereas a smaller particle size would be preferable
and 2 mm). CO consumption rate increases gradually from the for smaller channels. The use of larger particle 0.6 mm in
surface to the center for the smaller particles (Figure 3c) because diameter results in even higher effectiveness factor, but CH4
of reduction in rate inhibition by CO. For larger particles, CO selectivity is higher than with the two smaller particles, which
consumption rate passes through a maximum and then makes it less suitable for commercial applications.
3.3. Effects of Particle Shape and Catalyst Distribution.
decreases. For dp = 0.6 mm, the reaction rate at the center of
One way to minimize negative impact of intraparticle diffusional
the pellet is about one-third of the rate at the surface, whereas for
limitations on hydrocarbon selectivity (high CH4 and low C5+
the largest particle (dp = 2 mm), the CO rate approaches zero at selectivity) while maintaining pressure drop requirements is to
approximately 0.27 mm from the surface, and a large fraction of use eggshell catalyst distribution where cobalt is located near the
the particle volume is not utilized for reaction. Local methane outer pellet surface.4,33,55−57
selectivity increases with distance from the surface, but the Reduction in diffusion length can also be accomplished using
increase is small for two small particles (0.2 and 0.4 mm) and wall coated monolith reactors for FTS reaction, and this has
very rapid for the larger particles (0.6 and 2 mm). In the latter received considerable attention in recent years.12,58−62 This
case, local selectivity reaches 100% at 0.23 mm for 0.6 mm geometry can be approximated by that of an insulated slab (flat
particle and at 0.16 mm for 2 mm particle (Figure 3d). plate). Another way to minimize diffusion path of reactants is to
Effectiveness factor values are 1.01 (0.2 mm), 1.04 (0.4 mm), utilize hollow cylinder particles of small thickness.
1.12 (0.6 mm), and 0.52 (2 mm), whereas the corresponding Figure 4 shows results of simulations with these three
average values of CH4 selectivity are 5.6, 5.7, 7.6, and 19.4%, geometric shapes (slab/plate, hollow cylinder, and sphere with
respectively. eggshell distribution) having the same catalyst layer thickness of
G DOI: 10.1021/acs.iecr.7b00053
Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX
Industrial & Engineering Chemistry Research Article

Table 3. Effect of Process Conditions on Effectiveness Factor and Average Methane Selectivity of a Spherical Catalyst Particle
effect of process conditions P (bar) T (K) H2/CO (−) XCO (%) η (−) S̅CH4 (%) ϕCOa (−)
dp = 0.4 mm effect of pressure 20 473 2 0 1.042 5.71 0.56
30 473 2 0 1.041 5.71 0.48
effect of temperature 25 473 2 0 1.042 5.67 0.52
25 493 2 0 1.112 12.7 0.76
effect of syngas ratio 25 473 2 0 1.042 5.67 0.52
25 473 1.7 0 1.016 5.39 0.46
25 473 1.4 0 1.001 5.13 0.39
effect of CO conversion 25 473 2 0 1.042 5.67 0.52
25 473 2 25 1.036 5.05 0.53
25 473 2 50 1.020 4.08 0.55
dp = 2 mm effect of pressure 20 473 2 0 0.458 19.8 2.81
30 473 2 0 0.563 18.9 2.40
effect of temperature 25 473 2 0 0.516 19.4 2.58
25 493 2 0 0.355 28.3 3.82
effect of syngas ratio 25 473 2 0 0.516 19.4 2.58
25 473 1.7 0 0.552 16.4 2.28
25 473 1.4 0 0.586 11.8 1.97
effect of CO conversion 25 473 2 0 0.516 19.4 2.58
25 473 2 50 0.455 13.8 2.75
a
ϕCO defined by eq 13.

0.15 mm (L = δ = ro − ri = 0.15 mm in Figure 1, and ro = R = 1 eggshell thickness should be less than about 0.13 mm (see
mm). CO concentration declines monotonically with distance section 3.5).
from the surface for slab and eggshell catalysts, whereas it passes Several research groups have reported that low CH4 selectivity,
through a minimum at approximately 0.075 mm for a hollow comparable to that of powder catalysts, can be obtained with
cylinder (Figure 4a). The observed trend for the latter is the monolith reactors having layer thickness less than 0.05 mm.12,58
consequence of the fact that a hollow cylinder has two surfaces This is much more conservative than the estimated value for
exposed to the surrounding fluid, whereas the slab and the CAER catalyst at base case conditions. Iglesia et al.4 used eggshell
particle with the eggshell catalyst distribution have only one catalyst distribution (thickness 0.1−0.2 mm) with spherical
surface exposed to the fluid. The steepest gradient is observed particles 2.2 mm in diameter at 473 K, 20 bar, H2/CO = 2.1, and
with the slab layer, followed by the eggshell layer. As a result of CO conversion of 50−60%, whereas Fratalocchi et al.56 reported
differences in diffusion rates of H2 and CO in the particle, the H2/ that eggshell catalyst with 0.075 mm thickness (0.6 mm particle
CO ratio increases with the distance from the surface, with a diameter) has the same CH4 selectivity and slightly higher C5+
trend opposite to that of CO concentration (Figure 4b). Internal
selectivity than that of the powder catalyst (0.075−0.10 mm
H2/CO ratios for the slab and eggshell layer are appreciably
diameter) at 503 K, 25 bar, and H2/CO = 1.7 and CO
higher than those at the surface, whereas this ratio is relatively
constant for the hollow cylinder. Local CH4 selectivity follows conversions between 34 and 42%. Our estimate is in good
the same qualitative trend as the H2/CO ratio (Figure 4c). We agreement with values reported by Iglesia et al.4 under similar
see from these data that intraparticle diffusional resistance reaction conditions and with a similar particle size, whereas
increases in the following order: hollow cylinder < eggshell layer Fratalocchi et al.56 reported smaller thickness but at significantly
< slab, which is consistent with the order of increasing effective higher reaction temperature and with a smaller overall particle
diffusional path (Lc = V/S) for these three geometries. Even size.
though these three geometries have the same catalyst layer 3.4. Effect of Process Conditions. Effect of process
thickness, the corresponding effective diffusional paths are conditions (pressure, temperature, syngas ratio on the surface,
different. and CO conversion level) on effectiveness factor and methane
Effectiveness factor values are 1.216 (slab), 1.171 (eggshell), selectivity of CAER catalyst is summarized in Table 3 for two
and 1.024 (hollow cylinder). As seen above for small spherical spherical particles with diameters of 0.4 mm and 2 mm. We
particles, a moderate intraparticle diffusional resistance results in found (section 3.2) that the use of a small particle (0.4 mm)
higher effectiveness factor due to reduction in rate inhibition by results in negligible diffusional limitations at the baseline
CO (negative reaction order kinetics). However, CH4 selectivity conditions (473 K, 25 bar, H2/CO = 2, and XCO = 0%), whereas
is adversely affected by diffusional limitations because of increase pore diffusion has a strong effect on catalyst performance in the
in H2/CO ratio inside the pellet, and the average values of CH4 case of the 2 mm particle. Results shown with these two particle
selectivity are 7.90% (slab), 6.53% (eggshell), and 5.62% (hollow sizes show influence of process conditions in kinetic regime (dp =
cylinder). In this case, the hollow cylinder pellet would be the 0.4 mm) and pore diffusion regime (dp = 2 mm).
best choice from the performance perspective; however, it would Change of total pressure from 20 to 30 bar has a negligible
not be suitable for industrial applications because of low effect on η and SC̅ H4 for the small particle, whereas for 2 mm, the
mechanical strength. To avoid diffusional limitations in wall
coated monolith reactors using CAER catalyst at these increase in pressure results in a small increase in η and a small
conditions (473 K, 25 bar, H2/CO = 2, XCO = 0%), one would decrease in S̅CH4, which is due to decrease in the value of the
need to use layers with thickness less than 0.11 mm, whereas the Thiele modulus.
H DOI: 10.1021/acs.iecr.7b00053
Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX
Industrial & Engineering Chemistry Research Article

Increase in temperature from 473 to 493 K results in 7%


increase in the value of η and significant increase in methane
selectivity (from 5.67 to 12.7%) for dp = 0.4 mm, whereas in the
region of pore diffusion control (dp = 2 mm), η decreases and
S̅CH4 increases because of increase in diffusional resistance
(higher value of ϕCO).
Effect of syngas ratio changes on η is relatively small, but
decrease in the H2/CO ratio results in reduction of SC̅ H4 for both
particle sizes. The effect on S̅CH4 is more pronounced in the case
of larger particles.
Vervloet et al.7 have shown through model simulations and
optimization that space-time-yield of C5+ hydrocarbons can be
significantly improved relative to typical conditions of T = 500 K,
P = 30 bar, and H2/CO = 2 by increasing reaction temperature to
530 K and pressure to 36 bar while simultaneously decreasing
bulk H2/CO ratio to 1. Robota et al.8 showed CO, H2, and H2/
CO profiles for different bulk H2/CO ratios ranging from 0.5 to
1.8 for a 0.25 mm diameter particle at 478 K and 24 bar using the
YS rate equation with scaling factor of 10. For H2/CO bulk ratios
less than 0.9, the internal H2/CO ratio decreases slightly from the
surface to the center of the particles, whereas for bulk ratios
greater than 1.1, the ratio increases with distance from the
surface.
It is clear from eq 7 that the rate of CH4 formation will
decrease with decreasing H2/CO ratio; however, it is known that
one of the mechanisms for catalyst deactivation is carbon
deposition,63−65 and operation at low H2/CO ratios would
accelerate the rate of deactivation. Therefore, it is not clear
whether lowering of H2/CO feed ratio would be a feasible
method of controlling CH4 selectivity under diffusion limited Figure 5. Effectiveness factor as a function of Thiele modulus. (a) Slab,
conditions. sphere, eggshell and hollow cylinder particle (range of conditions: T =
473−493 K, P = 20−30 bar, CAER and YS catalyst, XCO = 0−50%, dp =
As the conversion along the reactor increases, the concen- 0.085−4 mm, L = 0.035−0.300 mm, δ = 0.075−0.300 mm, ro − ri =
tration of water outside the catalyst particle increases, and this 0.150−0.500 mm). (b) Spherical particle (CAER catalyst, range of
has significant effect on CH4 selectivity. The effect of CO conditions: T = 473−493 K, P = 20−30 bar, XCO = 0−50%, dp = 0.085−4
conversion on η is small with both particles because the value of mm).
the Thiele modulus does not change much with conversion (0−
50% range). However, significant reductions in S̅CH4 were Results in Figure 5a show η versus ϕ data for all shapes, sizes,
observed as conversion increases from 0% (reactor inlet) to 50% and process conditions investigated except for data correspond-
because of increase in water concentration, which inhibits ing to H2/CO ratios of 1.4 and 1.7. These data are excluded
methane formation rate (see eq 7). because the H2/CO ratio has effect on both boundary condition
3.5. Effectiveness Factor and Methane Selectivity as a (eq 2a) for H2, as well as on dimensionless reaction rate in the
Function of Generalized Thiele Modulus. Results from differential equation (eq 1). Changes in total pressure have no
simulations with different particle shapes (sphere, slab, hollow effect on the boundary conditions and have a small effect on
cylinder), different sizes (sphere with diameters ranging from reaction rate for the range of pressures investigated (20−30 bar);
0.085 mm to 4 mm, catalyst layers of different thicknesses thus, data for all pressures in this range are shown in Figure 5a.
ranging from 0.035 mm to 0.5 mm), form of catalyst distribution Temperature has a small effect on the boundary condition (eq
(uniform and eggshell), and process conditions (T = 473−493 K, 2a) for H2 (through variation of solubility with temperature), and
P = 20−30 bar, H2/CO = 1.4−2) are shown in Figures 5 and 6. dimensionless reaction rate is not very sensitive to temperature
Abscissa values represent generalized Thiele modulus based (473−493 K) at a constant value of ϕ. Therefore, data at both
on characteristic length for diffusion (Lc = V/S) as proposed by temperatures (473 and 493 K) are included. The largest scatter of
Aris.66 The formula for generalized Thiele modulus is points occurs in the region 0.7 < ϕ < 1, because different shapes
have different values of η at similar values of ϕ, and differences in
ρp ·( −R COs) kinetic parameters (CAER vs YS catalyst) have impact on results
ϕ = ϕCO = Lc s
De,CO·CCO in this range of ϕ values. In general, values of η for spherical
(13)
particles become greater than 1.06 in the region 0.7 < ϕ < 1,
where Lc = V/S (V = volume in which the catalyst is located; S = whereas for other shapes this takes place at higher values of ϕ
surface area exposed to surrounding fluid). (0.99 < ϕ < 1.26). For ϕ values less than about 0.6, η ≈ 1 (1−
Characteristic diffusion lengths for different pellet shapes and 1.06), which is the region of kinetic control, whereas for ϕ > 1.6,
forms of catalyst distribution as shown in Figure 1 are as follows: the effectiveness factor decreases with increasing ϕ and is less
Lc = R/3 (sphere with uniform catalyst distribution), Lc = (R/3)· than 1 for all particle shapes. This is the region where
[1 − (1 − δ/R)3] (sphere with eggshell catalyst distribution), Lc intraparticle diffusional resistance becomes increasingly impor-
= (ro − ri)/2 (hollow cylinder), and Lc = L (plate/slab). tant. For the YS catalyst, the maximum value of ϕ was 0.94
I DOI: 10.1021/acs.iecr.7b00053
Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX
Industrial & Engineering Chemistry Research Article

6b (slab, eggshell, and hollow cylinder). Methane selectivity does


not depend only on ϕ, but also on temperature, H2/CO ratio
outside the particle, and conversion level in the reactor (i.e., on
partial pressure of water at the catalyst surface); therefore, data
were grouped into several categories. Legends in Figure 6 show
which condition was changed relative to base case value, while
the remaining conditions are kept at the corresponding base case
values. For ϕ < 0.7, average methane selectivity is nearly constant
and then starts increasing rapidly with increasing value of the
Thiele modulus (Figure 6a). Decrease in the H2/CO ratio results
in lower CH4 selectivity compared to the base case (H2/CO = 2),
but the impact is small in the kinetically controlled region (ϕ <
0.6−0.7), whereas the effect is much more pronounced as ϕ
increases. Increase in CO conversion results in significant
reduction of CH4 selectivity (XCO = 50% vs XCO = 0%) even in
the kinetically controlled region. As stated earlier, this is caused
by higher partial pressure of water which inhibits methane
formation rate (see eq 7). Increase in temperature to 493 K leads
to increase in CH4 selectivity for all values of ϕ. At both
temperatures, CH4 selectivity starts increasing rapidly for ϕ > 0.7.
Changes in pressure (20 or 30 bar) have small effect in both
kinetic (ϕ = 0.48−0.56) and diffusion control (ϕ = 2.4−2.8)
regions.
Selected CH4 selectivity data for different shapes (slab and
hollow cylinder) and eggshell distribution with spherical particle
with dp = 2 mm are shown in Figure 6b for ϕ values less than 1.3.
The qualitative trends regarding the influence of H2/CO ratio,
CO conversion, temperature, and pressure are the same as in the
case of spherical particles with uniform catalyst distribution. In
this case, the kinetically controlled region extends to ϕ ≈ 0.9.
Figure 6. Methane selectivity as a function of Thiele modulus (a) for Both higher CO conversion and lower H2/CO feed ratio have
spherical particle under different conditions (shown in legend) and (b)
for different shapes and conditions (shown in legend). Base case: T =
beneficial effect on CH4 selectivity, and the combined effect (H2/
473 K, P = 25 bar, H2/CO = 2, XCO = 0%. Other conditions: Legend CO = 1.7 and XCO = 50%) results in significant reduction in CH4
shows numerical value of condition that was changed from its base case selectivity relative to the base case conditions.
value, whereas the remaining conditions are kept at the corresponding From data shown in Figure 6 we can estimate the maximum
base case values. particle size (sphere) or thickness of catalyst layer for slab and
eggshell catalyst distribution for which the intraparticle diffusion
(corresponding to dp = 4 mm) for the range of conditions limitations may be considered to have negligible effect on CH4
investigated because of low activity of this catalyst compared to selectivity at base case conditions for the CAER catalyst. We
the CAER catalyst. Most of the points for the YS catalyst are in define negligible effect on CH4 selectivity as a value of ϕ where
the region of negligible diffusional resistance (ϕ < 0.46). In the the increase in CH4 selectivity is less than 6% of the value
region 0.8 < ϕ < 1, the effectiveness factor for the YS catalyst is corresponding to minimum CH4 selectivity at the base case
higher than that for CAER catalyst (spherical geometry, H2/CO conditions and beyond which CH4 selectivity increases more
= 2). rapidly. The minimum value of CH4 selectivity for CAER catalyst
Our results are in qualitative agreement with previous studies at the base conditions for all shapes considered is 5.60%. Using
on the effect of catalyst shape on the effectiveness factor as a the maximum 6% increase in CH4 selectivity as the criterion, this
function of generalized Thiele modulus. Rester and Aris67 gives the following values for the magnitude of ϕ: 0.65 for
reported that maximum deviation between flat plate and spherical particle with uniform catalyst distribution and 0.90 for
spherical particle for a first-order reaction is 14−16% and occurs all others (slab, hollow cylinder, and eggshell distribution). In
for 1.3 < ϕ < 1.7. Knudsen et al.68 showed that deviations could terms of particle size or catalyst layer thickness, we obtain dp =
be higher (up to 27.2%) for some forms of LH kinetics in the 0.50 mm for a spherical particle with uniform catalyst
region of ϕ values between 1 and 1.5. distribution; L = 0.116 mm (slab/flat plate thickness); δ =
Figure 5b shows data for CAER catalyst and spherical particle 0.132 mm (eggshell thickness in 2 mm spherical particle); and ro
(P = 20−30 bar, T = 473−493 K, H2/CO = 2, XCO = 0−50%). As − ri = 0.232 mm (hollow cylinder with the outer radius of 1 mm).
can be seen, the data scatter is much less, in comparison to Figure It should be noted that in all these cases the catalyst effectiveness
5a, when the effects of particle shape and catalyst type are factor is greater than 1 (η = 1.08−1.12).
removed. Minor scatter is caused by differences in pressure,
temperature, and/or conversion level. The maximum value of η is 4. CONCLUSIONS
1.15 for ϕ = 0.71 (dp = 0.4 mm at 493 K, 30 bar and XCO = 0%), We have utilized a simple one-dimensional catalyst particle
whereas the maximum value of η at base case conditions (473 K, model which consists of four second-order differential equations
25 bar, XCO = 0%) is 1.12 at ϕ = 0.77 (dp = 0.6 mm). for reactants (H2 and CO) and product species (H2O and
Methane selectivity as a function of generalized Thiele C6H14) and an algebraic equation for methane formation rate in
modulus is shown in Figure 6a (spherical particle) and Figure terms of concentrations of H2, CO, and H2O inside the particle.
J DOI: 10.1021/acs.iecr.7b00053
Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX
Industrial & Engineering Chemistry Research Article

Numerical solution of this model enables one to calculate two Notes


key performance parameters: catalyst effectiveness factor and The authors declare no competing financial interest.

methane selectivity (both local and average). Kinetic parameters Deceased.


for the rate of CO consumption and rate of CH4 formation were
determined from experimental data with a state-of-the art 25% ACKNOWLEDGMENTS
Co/0.48% Re/γ-Al2O3 catalyst (CAER catalyst).
This research was made possible by a grant (NPRP Grant 7-559-
It was shown that activity of CAER catalyst is 15−37 higher
2-211) from the Qatar National Research Fund (a member of
(depending on process conditions) than that of Yates and
Qatar Foundation). The statements made herein are solely the
Satterfield catalyst. Kinetic parameters for YS catalyst were used
responsibility of the authors.


in several recent modeling studies of diffusion with chemical
reaction. The use of kinetic parameters derived from original
experimental data with YS catalyst severely underestimates the NOTATION
importance of diffusional limitations in FTS under conditions a = adsorption coefficient for CO reaction rate (bar−1)
representative of industrial practice. From numerical simulations aM = reaction order of CO in methane formation reaction
with kinetic parameters for CAER catalyst, the following AM = pre-exponential factor for CH4 reaction rate coefficient
conclusions can be made: (mol kg−1 s−1 bar−0.29)
• The use of small spherical particles (0.2−0.5 mm) or bM = reaction order of H2 in methane formation reaction
eggshell distribution with a larger spherical particle (dp = 2 Ci = concentration of species i in liquid phase (mol m−3)
mm) with catalyst layer thickness less than approximately CL = total liquid phase concentration (mol m−3)
dp = sphere particle diameter (mm)
0.13 mm is needed to avoid negative impact of diffusional
De,i = effective diffusivity coefficient of species i in particle
limitations on CH4 selectivity. The catalyst effectiveness
(m2 s−1)
factor under these conditions is slightly higher than 1
Dwax,i = diffusivity coefficient of component i in wax (m2 s−1)
because of negative reaction order with respect to CO at EM = activation energy for CH4 reaction (J mol−1)
high CO concentrations (i.e., low CO conversion inside Hi = Henry’s law constant for component i (bar)
the particle). k = reaction rate coefficient for CO consumption reaction rate
• For monolith reactors with wash-coated catalyst, diffu- (mol kg−1 s−1 bar−2)
sional limitations can be avoided by using catalyst layer kM = CH4 formation reaction rate constant (mol kg−1 s−1
thickness less than 0.12 mm at base case conditions (473 bar−0.29)
K, 25 bar, H2/CO = 2, XCO = 0%). keff = effective thermal conductivity of catalyst particle
• Methane selectivity decreases with increase in overall CO (W m−1 K−1)
conversion along the reactor because of presence of water L = slab thickness (mm)
which inhibits CH4 formation rate. Lc = characteristic diffusion length (mm)
• The use of substoichiometric H2/CO feed ratios (H2/CO Lg = geometric characteristic length (mm)
= 1.7 and 1.4 in our study) leads to improvement in lower mM = water effect coefficient in CH4 formation equation
values of CH4 selectivity (lower H2/CO ratio inside the P = total pressure (bar)
pellet). This is consistent with results reported by Vervloet Pi = partial pressure of component i (bar)
et al.7 r = variable radius within sphere and hollow cylinder (mm)
• Total pressure (20−30 bar) has a small effect on the ri = hollow cylinder inner radius (mm)
catalyst effectiveness factor and CH4 selectivity. ro = hollow cylinder outer radius (mm)
• Temperature (473−493 K) has a small effect on the R = sphere radius (mm)
effectiveness factor, but CH4 selectivity increases signifi- RCH4 = reaction rate of CH4 (mol kg−1 s−1)
cantly with increase in temperature. RCO = reaction rate of CO (mol kg−1 s−1)
• Graphs of η versus ϕ presented in the paper can provide R̅ CO = dimensionless CO reaction rate
rough estimates of η for other Co-based FT catalysts Rg = universal gas constant (J mol−1 K−1)
provided that sufficient information is available to calculate S = surface area of catalyst exposed to surrounding fluid (m2)
the value of ϕ. The biggest errors would occur in the SCH4 = local methane selectivity (%)
region 0.5 < ϕ < 1.5. SC̅ H4 = average methane selectivity (%)
T = temperature (K)
To capitalize on some of these findings in practical V = volume of catalyst loaded region (m3)
applications, one would need to develop preparation methods x = dimensionless length
which increase catalyst loading in a layer near the surface XCO = CO conversion (%)
(eggshell catalyst distribution) and develop Co catalysts that are yi = dimensionless concentration of species i in liquid phase
stable during long-term operation at low H2/CO feed ratios.


z = variable length within slab (mm)
AUTHOR INFORMATION Greek Letters
α = chain growth parameter
Corresponding Author
δ = eggshell layer thickness (mm)
*Chemical Engineering Program, Texas A&M University at ε = pellet porosity
Qatar 219S Texas A&M Engineering Building, Education City, η = effectiveness factor
PO Box 23874, Doha, Qatar. Tel.: +974 4423 0134. Fax: +30 νi = stoichiometric coefficient of species i
2310 996184. E-mail: dragomir.bukur@qatar.tamu.edu. ρp = catalyst density (kg m−3)
ORCID ρw = wax density (kg m−3)
Dragomir B. Bukur: 0000-0002-5065-4163 τ = pellet tortuosity
K DOI: 10.1021/acs.iecr.7b00053
Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX
Industrial & Engineering Chemistry Research Article

ϕ′i = dimensionless Thiele modulus based on geometric (20) Mazidi, S. K.; Sadeghi, M. T.; Marvast, M. A. Optimization of
characteristic length of a particle Fischer−Tropsch Process in a Fixed-Bed Reactor Using Non-uniform
ϕ = generalized Thiele modulus based on characteristic length Catalysts. Chem. Eng. Technol. 2013, 36, 62−72.
for diffusion (21) Brunner, K. M.; Perez, H. D.; Peguin, R. P. S.; Duncan, J. C.;
Harrison, L. D.; Bartholomew, C. H.; Hecker, W. C. Effects of Particle
Superscripts Size and Shape on the Performance of a Trickle Fixed-Bed Recycle
s = value at the surface


Reactor for Fischer−Tropsch Synthesis. Ind. Eng. Chem. Res. 2015, 54,
2902−2909.
REFERENCES (22) Post, M. F. M.; Van’t Hoog, A. C.; Minderhoud, J. K.; Sie, S. T.
(1) Steynberg, A.; Dry, M.; Davis, B.; Breman, B. Fischer−Tropsch Diffusion limitations in Fischer−Tropsch catalysts. AIChE J. 1989, 35,
reactors. Stud. Surf. Sci. Catal. 2004, 152, 64−195. 1107−1114.
(2) Rytter, E.; Tsakoumis, N. E.; Holmen, A. On the selectivity to (23) Dixit, R. S.; Tavlarides, L. L. Integral method of analysis of
higher hydrocarbons in Co-based Fischer−Tropsch synthesis. Catal. Fischer−Tropsch synthesis reactions in a catalyst pellet. Chem. Eng. Sci.
Today 2016, 261, 3−16. 1982, 37, 539−544.
(3) Todic, B.; Ordomsky, V. V.; Nikacevic, N. M.; Khodakov, A. Y.; (24) Dixit, R. S.; Tavlarides, L. L. Kinetics of the Fischer−Tropsch
Bukur, D. B. Opportunities for intensification of Fischer−Tropsch synthesis. Ind. Eng. Chem. Process Des. Dev. 1983, 22, 1−9.
synthesis through reduced formation of methane over cobalt catalysts in (25) Yates, I. C.; Satterfield, C. N. Intrinsic kinetics of the Fischer−
microreactors. Catal. Sci. Technol. 2015, 5, 1400−1411. Tropsch synthesis on a cobalt catalyst. Energy Fuels 1991, 5, 168−173.
(4) Iglesia, E.; Soled, S. L.; Baumgartner, J. E.; Reyes, S. C. Synthesis (26) Iglesia, E.; Reyes, S.; Madon, R.; Soled, S. In Advances in Catalysis
and Catalytic Properties of Eggshell Cobalt Catalysts for the Fischer− and Related Subjects; Eley, D., Pines, H., Weisz, P., Eds.; Academic Press:
Tropsch Synthesis. J. Catal. 1995, 153, 108−122. New York, 1993; Vol. 39, p 239.
(5) Wang, Y.-N.; Xu, Y.-Y.; Xiang, H.-W.; Li, Y.-W.; Zhang, B.-J. (27) Iglesia, E.; Reyes, S.; Soled, S. In Computer-Aided Design of
Modeling of catalyst pellets for Fischer−Tropsch synthesis. Ind. Eng. Catalysts and Reactors; Becker, E., Pereira, C., Eds.; Marcel Dekker: New
Chem. Res. 2001, 40, 4324−4335. York, 1993; p 199.
(6) Becker, H.; Güttel, R.; Turek, T. Experimental evaluation of (28) Iglesia, E.; Reyes, S. C.; Madon, R. J. Transport-enhanced α-olefin
catalyst layers with bimodal pore structure for Fischer−Tropsch readsorption pathways in Ru-catalyzed hydrocarbon synthesis. J. Catal.
synthesis. Catal. Today 2016, 275, 155−163. 1991, 129, 238−256.
(7) Vervloet, D.; Kapteijn, F.; Nijenhuis, J.; van Ommen, J. R. Fischer− (29) Wang, Y.-N.; Ma, W.-P.; Lu, Y.-J.; Yang, J.; Xu, Y.-Y.; Xiang, H.-
Tropsch reaction-diffusion in a cobalt catalyst particle: aspects of activity
W.; Li, Y.-W.; Zhao, Y.-L.; Zhang, B.-J. Kinetics modelling of Fischer−
and selectivity for a variable chain growth probability. Catal. Sci. Technol.
Tropsch synthesis over an industrial Fe−Cu−K catalyst. Fuel 2003, 82,
2012, 2, 1221−1233.
(8) Robota, H. J.; Richard, L. A.; Deshmukh, S.; LeViness, S.; 195−213.
Leonarduzzi, D.; Roberts, D. High Activity and Selective Fischer− (30) Ghouri, M. M.; Afzal, S.; Hussain, R.; Blank, J.; Bukur, D. B.;
Tropsch Catalysts for Use in a Microchannel Reactor. Catal. Surv. Asia Elbashir, N. O. Multi-scale modeling of fixed-bed Fischer−Tropsch
2014, 18, 177−182. reactor. Comput. Chem. Eng. 2016, 91, 38−48.
(9) Myrstad, R.; Eri, S.; Pfeifer, P.; Rytter, E.; Holmen, A. Fischer− (31) Todic, B.; Bhatelia, T.; Froment, G. F.; Ma, W.; Jacobs, G.; Davis,
Tropsch synthesis in a microstructured reactor. Catal. Today 2009, 147, B. H.; Bukur, D. B. Kinetic model of Fischer−Tropsch synthesis in a
S301−S304. slurry reactor on Co−Re/Al2O3 catalyst. Ind. Eng. Chem. Res. 2013, 52,
(10) De Deugd, R. M.; Kapteijn, F.; Moulijn, J. A. Trends in Fischer− 669−679.
Tropsch reactor technologyopportunities for structured reactors. (32) Becker, H.; Güttel, R.; Turek, T. Optimization of Catalysts for
Top. Catal. 2003, 26, 29−39. Fischer−Tropsch Synthesis by Introduction of Transport Pores. Chem.
(11) Guettel, R.; Turek, T. Comparison of different reactor types for Ing. Tech. 2014, 86, 544−549.
low temperature Fischer−Tropsch synthesis: A simulation study. Chem. (33) Gardezi, S. A.; Joseph, B. Performance Characteristics of Eggshell
Eng. Sci. 2009, 64, 955−964. Co/SiO2 Fischer−Tropsch Catalysts: A Modeling Study. Ind. Eng.
(12) Kapteijn, F.; de Deugd, R. M.; Moulijn, J. A. Fischer−Tropsch Chem. Res. 2015, 54, 8080−8092.
synthesis using monolithic catalysts. Catal. Today 2005, 105, 350−356. (34) De Deugd, R. M. Fischer−Tropsch synthesis revisited; Efficiency
(13) Wang, Y.; Van der Wiel, D. P.; Tonkovich, A. L. Y.; Gao, Y.; Baker, and Selectivity Benefits from Imposing Temporal and/or Spatial
E. G. Catalyst structure and method of fischer-tropsch synthesis. Battelle Structure in the Reactor. Ph.D. dissertation, Delft University of
Memorial Institute, U.S. Patent 6558634B1, 2003. Technology, Delft, 2004.
(14) Zimmerman, W. H.; Rossin, J. A.; Bukur, D. B. Effect of particle (35) Friedel, R.; Anderson, R. Composition of synthetic liquid fuels. I.
size on the activity of a fused iron Fischer−Tropsch catalyst. Ind. Eng. Product distribution and analysis of C5-C8 paraffin isomers from cobalt
Chem. Res. 1989, 28, 406−413. catalyst. J. Am. Chem. Soc. 1950, 72, 2307−2307.
(15) Yan, S.; Fan, L.; Zhang, Z.; Zhou, J.; Fujimoto, K. Supercritical- (36) Van Der Laan, G. P.; Beenackers, A. Kinetics and selectivity of the
phase process for selective synthesis of heavy hydrocarbons from syngas Fischer−Tropsch synthesis: a literature review. Catal. Rev.: Sci. Eng.
on cobalt catalysts. Appl. Catal., A 1998, 171, 247−254.
1999, 41, 255−318.
(16) Zhan, X.; Davis, B. H. Assessment of internal diffusion limitation
(37) Todic, B.; Ma, W.; Jacobs, G.; Davis, B. H.; Bukur, D. B. Effect of
on Fischer−Tropsch product distribution. Appl. Catal., A 2002, 236,
149−161. process conditions on the product distribution of Fischer−Tropsch
(17) Jung, A.; Kern, C.; Jess, A., Modeling of Internal Diffusion synthesis over a Re-promoted cobalt-alumina catalyst using a stirred
Limitations in a Fischer−Tropsch Catalyst. In Advances in Fischer− tank slurry reactor. J. Catal. 2014, 311, 325−338.
Tropsch Synthesis, Catalysts, and Catalysis; Davis, B. H., Occelli, M. L., (38) Bartholomew, C. H.; Farrauto, R. J. Fundamentals of Industrial
Eds.; CRC Press: Boca Raton, FL, 2009; pp 215−227. Catalytic Processes.; John Wiley & Sons: Hoboekn, NJ, 2006.
(18) Xu, B.; Fan, Y.; Zhang, Y.; Tsubaki, N. Pore diffusion simulation (39) Maretto, C.; Krishna, R. Modelling of a bubble column slurry
model of bimodal catalyst for Fischer−Tropsch synthesis. AIChE J. reactor for Fischer−Tropsch synthesis. Catal. Today 1999, 52, 279−
2005, 51, 2068−2076. 289.
(19) Hallac, B. B.; Keyvanloo, K.; Hedengren, J. D.; Hecker, W. C.; (40) Ma, W.; Jacobs, G.; Das, T. K.; Masuku, C. M.; Kang, J.; Pendyala,
Argyle, M. D. An optimized simulation model for iron-based Fischer− V. R. R.; Davis, B. H.; Klettlinger, J. L.; Yen, C. H. Fischer−Tropsch
Tropsch catalyst design: Transfer limitations as functions of operating Synthesis: Kinetics and Water Effect on Methane Formation over 25%
and design conditions. Chem. Eng. J. 2015, 263, 268−279. Co/γ-Al2O3 Catalyst. Ind. Eng. Chem. Res. 2014, 53, 2157−2166.

L DOI: 10.1021/acs.iecr.7b00053
Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX
Industrial & Engineering Chemistry Research Article

(41) Storsæter, S.; Borg, Ø.; Blekkan, E. A.; Holmen, A. Study of the (62) Almeida, L.; González, O.; Sanz, O.; Paul, A.; Centeno, M.;
effect of water on Fischer−Tropsch synthesis over supported cobalt Odriozola, J.; Montes, M. Fischer-tropsch catalyst deposition on
catalysts. J. Catal. 2005, 231, 405−419. metallic structured supports. Stud. Surf. Sci. Catal. 2007, 167, 79−84.
(42) Schulz, H.; Claeys, M.; Harms, S. Effect of water partial pressure (63) Saib, A. M.; Moodley, D. J.; Ciobîcă, I. M.; Hauman, M. M.;
on steady state Fischer−Tropsch activity and selectivity of a promoted Sigwebela, B. H.; Weststrate, C. J.; Niemantsverdriet, J. W.; van de
cobalt catalyst. Stud. Surf. Sci. Catal. 1997, 107, 193−200. Loosdrecht, J. Fundamental understanding of deactivation and
(43) Botes, F. G. Influences of water and syngas partial pressure on the regeneration of cobalt Fischer−Tropsch synthesis catalysts. Catal.
kinetics of a commercial alumina-supported cobalt Fischer− Tropsch Today 2010, 154, 271−282.
catalyst. Ind. Eng. Chem. Res. 2009, 48, 1859−1865. (64) Moodley, D.; Van de Loosdrecht, J.; Saib, A.; Overett, M.; Datye,
(44) Krishnamoorthy, S.; Tu, M.; Ojeda, M. P.; Pinna, D.; Iglesia, E. An A.; Niemantsverdriet, J. Carbon deposition as a deactivation mechanism
investigation of the effects of water on rate and selectivity for the of cobalt-based Fischer−Tropsch synthesis catalysts under realistic
Fischer−Tropsch synthesis on cobalt-based catalysts. J. Catal. 2002, conditions. Appl. Catal., A 2009, 354, 102−110.
211, 422−433. (65) Tsakoumis, N. E.; Rønning, M.; Borg, Ø.; Rytter, E.; Holmen, A.
(45) Lögdberg, S.; Boutonnet, M.; Walmsley, J. C.; Järås, S.; Holmen, Deactivation of cobalt based Fischer−Tropsch catalysts: A review. Catal.
A.; Blekkan, E. A. Effect of water on the space-time yield of different Today 2010, 154, 162−182.
supported cobalt catalysts during Fischer−Tropsch synthesis. Appl. (66) Aris, R. On shape factors for irregular particlesI: The steady
Catal., A 2011, 393, 109−121. state problem. Diffusion and reaction. Chem. Eng. Sci. 1957, 6, 262−268.
(46) Bukur, D. B.; Pan, Z.; Ma, W.; Jacobs, G.; Davis, B. H. Effect of CO (67) Rester, S.; Aris, R. Communications on the theory of diffusion and
conversion on the product distribution of a Co/Al2O3 Fischer−Tropsch reactionII the effect of shape on the effectiveness factor. Chem. Eng.
synthesis catalyst using a fixed bed reactor. Catal. Lett. 2012, 142, 1382− Sci. 1969, 24, 793−795.
1387. (68) Knudsen, C. W.; Roberts, G. W.; Satterfield, C. N. Effect of
(47) Borg, Ø.; Eri, S.; Blekkan, E. A.; Storsæter, S.; Wigum, H.; Rytter, geometry on catalyst effectiveness factor. Langmuir-Hinshelwood
E.; Holmen, A. Fischer−Tropsch synthesis over γ-alumina-supported kinetics. Ind. Eng. Chem. Fundam. 1966, 5, 325−326.
cobalt catalysts: effect of support variables. J. Catal. 2007, 248, 89−100.
(48) Bertole, C. J.; Kiss, G.; Mims, C. A. The effect of surface-active
carbon on hydrocarbon selectivity in the cobalt-catalyzed Fischer−
Tropsch synthesis. J. Catal. 2004, 223, 309−318.
(49) Bertole, C. J.; Mims, C. A.; Kiss, G. The effect of water on the
cobalt-catalyzed Fischer−Tropsch synthesis. J. Catal. 2002, 210, 84−96.
(50) Akgerman, A. Diffusivities of synthesis gas and Fischer−Tropsch
products in slurry media, DOE Final Report for DOE Grant No. AC22-
84PC70032, 1987.
(51) Erkey, C.; Rodden, J.; Akgerman, A. A correlation for predicting
diffusion coefficients in alkanes. Can. J. Chem. Eng. 1990, 68, 661−665.
(52) Makrodimitri, Z. A.; Heller, A.; Koller, T. M.; Rausch, M. H.;
Fleys, M. S.; Bos, A. R.; van der Laan, G. P.; Fröba, A. P.; Economou, I.
G. Viscosity of heavy n-alkanes and diffusion of gases therein based on
molecular dynamics simulations and empirical correlations. J. Chem.
Thermodyn. 2015, 91, 101−107.
(53) Marano, J. J.; Holder, G. D. Characterization of Fischer−Tropsch
liquids for vapor-liquid equilibria calculations. Fluid Phase Equilib. 1997,
138, 1−21.
(54) Marano, J. J. Property Correlation and Characterization of
Fischer−Tropsch Liquids for Process Modeling. Ph.D. dissertation,
University of Pittsburgh, Pittsburgh, PA, 1996.
(55) Peluso, E.; Galarraga, C.; de Lasa, H. Eggshell catalyst in Fischer−
Tropsch synthesis: Intrinsic reaction kinetics. Chem. Eng. Sci. 2001, 56,
1239−1245.
(56) Fratalocchi, L.; Visconti, C. G.; Lietti, L.; Tronconi, E.; Rossini, S.
Exploiting the effects of mass transfer to boost the performances of Co/
γ-Al2O3 eggshell catalysts for the Fischer−Tropsch synthesis. Appl.
Catal., A 2016, 512, 36−42.
(57) Gardezi, S. A.; Wolan, J. T.; Joseph, B. Effect of catalyst
preparation conditions on the performance of eggshell cobalt/SiO2
catalysts for Fischer−Tropsch synthesis. Appl. Catal., A 2012, 447−448,
151−163.
(58) Hilmen, A.-M.; Bergene, E.; Lindvåg, O.; Schanke, D.; Eri, S.;
Holmen, A. Fischer−Tropsch synthesis on monolithic catalysts of
different materials. Catal. Today 2001, 69, 227−232.
(59) Visconti, C. G.; Tronconi, E.; Lietti, L.; Groppi, G.; Forzatti, P.;
Cristiani, C.; Zennaro, R.; Rossini, S. An experimental investigation of
Fischer−Tropsch synthesis over washcoated metallic structured
supports. Appl. Catal., A 2009, 370, 93−101.
(60) Kibby, C. L.; Saxton, R. J.; Jothimurugesan, K.; Das, T. K.;
Lacheen, H. S.; Bartz, M.; Has, A. Modified fischer-tropsch monolith
catalysts and methods for preparation and use thereof. Chevron U.S.A.
Inc., U.S. Patent 20130210942A1, 2013.
(61) Liu, W.; Hu, J.; Wang, Y. Fischer−Tropsch synthesis on ceramic
monolith-structured catalysts. Catal. Today 2009, 140, 142−148.

M DOI: 10.1021/acs.iecr.7b00053
Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX

You might also like