modeling and optimization of a large-scale slurry bubble ... · modeling and optimization of a...

11
Modeling and optimization of a large-scale slurry bubble column reactor for producing 10,000 bbl/day of Fischer–Tropsch liquid hydrocarbons Laurent Sehabiague a , Romain Lemoine a , Arsam Behkish a , Yannick J. Heintz a , Mariela Sanoja a , Rachid Oukaci b , Badie I. Morsi a, * a Chemical and Petroleum Engineering Department, University of Pittsburgh, Pittsburgh, PA 15261, USA b Energy Technology Partners, 135 William Pitt Way, Pittsburgh, PA 15238, USA Abstract A user-friendly simulator based on a comprehensive computer model for slurry bubble column reactors (SBCRs) for Fischer–Tropsch (F–T) synthesis, taking into account the hydrodynamics, kinetics, heat transfer, and mass transfer was developed. The hydrodynamic and mass transfer data obtained in our laboratories under typical F–T conditions along with those available in the literature were correlated using Back Propagation Neural Network and empirical correlations with high confidence levels. The data used covered wide ranges of reactor geometry, gas distributor, and operating conditions. All reactor partial differential equations, equation parameters and boundary conditions were simultaneously solved numerically. The simulator was systematically used to predict the effects of reactor geometry (inside diameter and height) as well as superficial gas velocity and catalyst concentration on the performance of a large-scale SBCR provided with cooling pipes and operating under F–T conditions with cobalt- supported catalyst and H 2 /CO = 2. The performance of the SBCR was expressed in terms of CO conversion, liquid hydrocarbon yield, catalyst productivity, and space time yield. The simulator was also used to optimize the reactor geometry and operating conditions in order to produce 10,000 barrels/day (bbl/day) of liquid hydrocarbons. # 2008 Published by Elsevier B.V. on behalf of Taiwan Institute of Chemical Engineers. Keywords: Hydrodynamics; Mass transfer; Multiphase reactors; Bubble column reactor 1. Introduction Synthetic fuels produced via Fischer–Tropsch (F–T) reactions exhibit quite attractive properties, such as very low sulfur content and almost no aromatics, which make them good candidates as alternative and/or complementary sources of fuel. The interest in gas-to-liquid (GTL), coal-to-liquid (CTL) or biomass-to-liquid (BTL) technologies using F–T process as shown in Fig. 1 will certainly increase in the coming years with more attractive costs when compared with those of crude oil. Recently, it has been estimated that using F–T synthesis to produce fuels can become more competitive if the price of the crude oil exceeds $20–24/bbl (Dry, 2004; Vosloo, 2001; Wilhelm et al., 2001). In the F–T synthesis, the syngas (CO + H 2 ) reacts in presence of a catalyst (commonly iron or cobalt) to form a mixture of liquid hydrocarbons: CO þ 1 þ m 2n H 2 ! r FT 1 n C n H m þ H 2 O (1) The main products of the F–T synthesis are paraffins and the usage ratio H 2 /CO is typically between 2.06 and 2.16 (Dry, 2002; Steynberg and Dry, 2004). It has been known since the 1950s that slurry bubble column reactors (SBCRs) could be used to carry out F–T synthesis. In this type of reactors, the solid-phase consists of fine catalytic particles suspended in a liquid phase. The advantages of SBCRs over agitated and packed-bed reactors include (Dry, 2002; Nigam and Schumpe, 1996; Satterfield and Huff, 1982; Zimmerman and Bukur, 1990): (1) better temperature control/removal due www.elsevier.com/locate/jcice Available online at www.sciencedirect.com Journal of the Chinese Institute of Chemical Engineers 39 (2008) 169–179 Abbreviations: BCR, bubble column reactor; F–T, Fischer–Tropsch synth- esis; M-ON, multiple orifices nozzle; PfP, perforated plate; PM, perforated membrane; PoP, porous plate; R, ring; S, spider; SBCR, slurry bubble column reactor; S-ON, single orifice nozzle; SP, sintered plate; VLE, vapor–liquid equilibrium. * Corresponding author. Tel.: +1 412 624 9650; fax: +1 412 624 9639. E-mail address: [email protected] (B.I. Morsi). 0368-1653/$ – see front matter # 2008 Published by Elsevier B.V. on behalf of Taiwan Institute of Chemical Engineers. doi:10.1016/j.jcice.2007.11.003

Upload: nguyenbao

Post on 17-Apr-2018

219 views

Category:

Documents


3 download

TRANSCRIPT

www.elsevier.com/locate/jcice

Available online at www.sciencedirect.com

l Engineers 39 (2008) 169–179

Journal of the Chinese Institute of Chemica

Modeling and optimization of a large-scale slurry bubble

column reactor for producing 10,000 bbl/day of

Fischer–Tropsch liquid hydrocarbons

Laurent Sehabiague a, Romain Lemoine a, Arsam Behkish a, Yannick J. Heintz a,Mariela Sanoja a, Rachid Oukaci b, Badie I. Morsi a,*

a Chemical and Petroleum Engineering Department, University of Pittsburgh, Pittsburgh, PA 15261, USAb Energy Technology Partners, 135 William Pitt Way, Pittsburgh, PA 15238, USA

Abstract

A user-friendly simulator based on a comprehensive computer model for slurry bubble column reactors (SBCRs) for Fischer–Tropsch (F–T)

synthesis, taking into account the hydrodynamics, kinetics, heat transfer, and mass transfer was developed. The hydrodynamic and mass transfer

data obtained in our laboratories under typical F–T conditions along with those available in the literature were correlated using Back Propagation

Neural Network and empirical correlations with high confidence levels. The data used covered wide ranges of reactor geometry, gas distributor, and

operating conditions. All reactor partial differential equations, equation parameters and boundary conditions were simultaneously solved

numerically.

The simulator was systematically used to predict the effects of reactor geometry (inside diameter and height) as well as superficial gas velocity

and catalyst concentration on the performance of a large-scale SBCR provided with cooling pipes and operating under F–T conditions with cobalt-

supported catalyst and H2/CO = 2. The performance of the SBCR was expressed in terms of CO conversion, liquid hydrocarbon yield, catalyst

productivity, and space time yield. The simulator was also used to optimize the reactor geometry and operating conditions in order to produce

10,000 barrels/day (bbl/day) of liquid hydrocarbons.

# 2008 Published by Elsevier B.V. on behalf of Taiwan Institute of Chemical Engineers.

Keywords: Hydrodynamics; Mass transfer; Multiphase reactors; Bubble column reactor

1. Introduction

Synthetic fuels produced via Fischer–Tropsch (F–T)

reactions exhibit quite attractive properties, such as very low

sulfur content and almost no aromatics, which make them good

candidates as alternative and/or complementary sources of fuel.

The interest in gas-to-liquid (GTL), coal-to-liquid (CTL) or

biomass-to-liquid (BTL) technologies using F–T process as

shown in Fig. 1 will certainly increase in the coming years with

more attractive costs when compared with those of crude oil.

Recently, it has been estimated that using F–T synthesis to

Abbreviations: BCR, bubble column reactor; F–T, Fischer–Tropsch synth-

esis; M-ON, multiple orifices nozzle; PfP, perforated plate; PM, perforated

membrane; PoP, porous plate; R, ring; S, spider; SBCR, slurry bubble column

reactor; S-ON, single orifice nozzle; SP, sintered plate; VLE, vapor–liquid

equilibrium.

* Corresponding author. Tel.: +1 412 624 9650; fax: +1 412 624 9639.

E-mail address: [email protected] (B.I. Morsi).

0368-1653/$ – see front matter # 2008 Published by Elsevier B.V. on behalf of T

doi:10.1016/j.jcice.2007.11.003

produce fuels can become more competitive if the price of the

crude oil exceeds $20–24/bbl (Dry, 2004; Vosloo, 2001;

Wilhelm et al., 2001).

In the F–T synthesis, the syngas (CO + H2) reacts in

presence of a catalyst (commonly iron or cobalt) to form a

mixture of liquid hydrocarbons:

COþ�

1þ m

2n

�H2�!

rFT

�1

n

�CnHm þ H2O (1)

The main products of the F–T synthesis are paraffins and

the usage ratio H2/CO is typically between 2.06 and 2.16

(Dry, 2002; Steynberg and Dry, 2004). It has been known

since the 1950s that slurry bubble column reactors (SBCRs)

could be used to carry out F–T synthesis. In this type of

reactors, the solid-phase consists of fine catalytic particles

suspended in a liquid phase. The advantages of SBCRs over

agitated and packed-bed reactors include (Dry, 2002; Nigam

and Schumpe, 1996; Satterfield and Huff, 1982; Zimmerman

and Bukur, 1990): (1) better temperature control/removal due

aiwan Institute of Chemical Engineers.

Nomenclature

a gas–liquid interfacial area per unit slurry volume

(1/m)

a1 constant in Eq. (33)

A1 coefficient in Eq. (34) (mol/kgcatalyst Pa2 s)

A2 coefficient in Eq. (35) (1/Pa)

C reactant concentration (mol/m3)

C* equilibrium concentration (mol/m3)

CW solid concentration by weight (wt.%)

dorf. diameter of the orifices of the gas distributor (m)

dP diameter of the solid particles (m)

dS Sauter-mean bubble diameter (m)

D dispersion coefficient (m2/s)

DAB diffusivity of gas in the liquid (m2/s)

DC diameter of the column (m)

E constant in Eqs. (34) and (35) (J/mol)

F expression in Eq. (19)

g gravitational constant (m/s2)

h0 heat transfer coefficient (W/m2 K)

H0 coefficient for Henry’s law constant (Pa m3/mol)

He Henry’s law constant (Pa m3/mol)

DH heat of reaction (J/mol)

k reaction rate constant (mol/kg s Pa2)

kLa volumetric liquid-side mass transfer coefficient

(1/s)

L reactor length (m)

m constant in Eq. (1)

MW molecular weight (kg/mol)

n constant in Eq. (1)

NO number of orifices of the gas distributor

P pressure (Pa)

PS saturation vapor pressure (Pa)

r reaction rate (mol/m3 s)

rFT Fischer–Tropsch reaction rate (mol/kg catalyst s)

R universal gas constant (J/mol K)

T temperature (K)

U superficial velocity (m/s)

UTS terminal settling velocity (m/s)

XW concentration of purest component in the liquid

mixture (wt.%)

Greek symbols

a constant defined in Table 2

G gas sparger coefficient

e holdup

m viscosity (kg/m s)

j open area

r density (kg/m3)

s surface tension (N/m)

Subscripts

FT refer to the Fischer–Tropsch reaction

G refer to the gas phase

i refer to the reactant/species i

L refer to the liquid phase

large refer to the large bubble class in the churn-

turbulent flow regime

P refer to the solid particles

S refer to the solid particles suspension

SL refer to the slurry phase

small refer to the small bubble class in the churn-

turbulent flow regime

L. Sehabiague et al. / Journal of the Chinese Institute of Chemical Engineers 39 (2008) 169–179170

to large liquid volume; (2) low maintenance/operating cost

due in part to simple design and absence of moving

components; (3) low capital cost due to the small floor space

needed; (4) lower pressure drop which is equivalent to the

hydrostatic head; (5) ability of using fine catalyst particles

(<100 mm) allowing huge surface area per unit volume

and better liquid–solid mass transfer, (5) higher yield per

reactor volume; (6) higher on-line factor where the catalyst

can be added and removed continuously, allowing longer

runs without reactor shutdown; (7) higher gas–liquid

interfacial areas and volumetric mass transfer coefficients

at low energy input; and (8) greater liquid volume and longer

residence time due to reactor geometry and height to

diameter ratio (L/D). Despite these advantages, SBCRs

have not been applied in a commercial-scale until 1993

(Dry, 2002) due to the following challenges (Nigam and

Schumpe, 1996): (1) considerable back-mixing and bubble–

bubble interactions in the churn-turbulent flow regime; (2)

catalyst attrition (particularly with iron) and deactivation due

to high shear created near the gas distributor; (3) important

side products due to high liquid to solid ratio; (4) difficult

catalyst/liquid separation, particularly for highly viscous

slurries containing fine particles; and (5) lack of knowledge

on the hydrodynamic and mass/heat transfer characteristics

in these reactors under typical industrial conditions.

Recently, correlations and a calculation algorithm (Behkish

et al., 2005, 2006; Lemoine, 2005; Lemoine et al., 2005) have

been developed to predict the hydrodynamics and mass

transfer parameters in bubble column reactors (BCRs) and

SBCRs based on a large experimental database covering

typical F–T conditions, however, those correlations and

algorithms have not been yet included in a model to simulate

the F–T synthesis in a large-scale SBCR. Therefore, the

purpose of this study is to build a comprehensive mathema-

tical model for the SBCRs using the correlations and

algorithm previously developed (Behkish et al., 2006;

Lemoine, 2005; Lemoine et al., 2005) and to use this model

for the design of a commercial-size F–T SBCR to produce

10,000 barrels/day (bbl/day) of synthetic liquid hydrocarbons.

The model will be also used to optimize the performances of

such a commercial reactor.

2. Simulation of the Fischer–Tropsch synthesis

In SBCRs, the gas-phase is conventionally sparged through

the slurry at the bottom of the reactor through a specially

designed distributor, leading to different flow regimes and

Fig. 1. GTL, CTL and BTL processes.

L. Sehabiague et al. / Journal of the Chinese Institute of Chemical Engineers 39 (2008) 169–179 171

complex hydrodynamic as well as mass transfer behaviors.

There are no flow regime maps available in the literature for

SBCRs, however, several flow regime maps were proposed to

delineate the hydrodynamic flow regimes in BCRs, including

the one by Oshinowo and Charles (1974), which identifies six

different flow regimes in an upward flow; and that by

Deckwer et al. (1980) based on the reactor diameter and gas

velocity for air/water system. In BCRs operating with

superficial gas velocities �0.05 m/s, the bubbly or homo-

genous flow regime prevails, which is characterized by a

homogeneous gas bubbles distribution, weak interactions

among gas bubbles, and almost constant gas bubbles

residence time. In this regime, the gas injection point was

reported to have a strong impact on the gas bubbles

formation, whereas the reactor diameter was not as important

(Magaud et al., 2001; Sarrafi et al., 1999). In small BCRs

with internal diameters �0.15 m, increasing the superficial

gas velocity could lead to the formation of large gas bubbles

in the reactor, which is designated as a slug flow regime. In

this regime, the wall effect (Lee and Tsui, 1999; Saxena and

Chen, 1994) is important and has a strong impact on the

hydrodynamic and mass transfer parameters. In large-scale

BCRs, however, increasing the superficial gas velocity leads

the reactor to operate in the heterogeneous or churn-turbulent

flow regime. In this regime, large and fast-rising gas bubbles

induce strong circulations and create back-mixing or re-

circulation zones wherein the small gas bubbles are entrained

(Shah et al., 1982; de Swart et al., 1996; Vermeer and

Krishna, 1981). Visual observations and photographic

methods revealed the coexistence of distinctly two-class

gas bubbles (small and large) in BCRs (Behkish et al., 2007;

Grund et al., 1992; Krishna et al., 2000; Lemoine et al.,

2004; Vermeer and Krishna, 1981) operating in the churn-

turbulent flow regime.

In this study, the commercial-scale SBCR for F–T synthesis

was simulated using an axial dispersion model in conjunction

with the two-class gas bubbles model (de Swart, 1996; de Swart

et al., 1996; de Swart and Krishna, 2002; Lemoine et al., 2004;

Rados et al., 2003). The model was built with the following

features:

- T

he SBCR is operated in the churn-turbulent flow regime.

- T

he gas is sparged at the bottom of the reactor through a

multiple orifices gas distributor.

- T

he catalyst used is a cobalt-supported catalyst.

- T

he slurry is moving upward.

- T

he total heat of reaction is removed using saturated water

flowing in a bundle of cooling tubes.

The model was also based on the following key assump-

tions:

- T

he mass transfer resistance in the bulk gas-phase and liquid–

solid interface are negligible when compared with that in the

liquid film.

- T

he slurry superficial velocity is constant.

- T

he slurry temperature is constant.

- T

he catalyst suspension behavior follows the sedimenta-

tion-dispersion model (Behkish, 2004; Kato et al., 1972;

Kojima et al., 1986; O’Dowd et al., 1987; Smith and

Ruether, 1985).

- T

he reactor is operating in a steady-state.

stitute of Chemical Engineers 39 (2008) 169–179

In the next sections the different equations and correlations

used in the simulator are given.

L. Sehabiague et al. / Journal of the Chinese In172

2.1. Mass and energy balances

The mass balances of each component in the liquid and gas

phases over a differential element of the reactor are:

@

@z

�ð1� eGÞDL

@Ci;L

@z

�� @ðUSLCi;LÞ

@z

þ kLai;largeðC�i;large � Ci;LÞ þ kLai;smallðC�i;small � Ci;LÞþ ð1� eGÞri ¼ 0 (2)

@

@z

�eG;smallDG;small

@Ci;G;small

@z

�� @ðUG;smallCi;G;smallÞ

@z

� kLai;smallðC�i;small � Ci;LÞ ¼ 0 (3)

@

@z

�eG;largeDG;large

@Ci;G;large

@z

�� @ðUG;largeCi;G;largeÞ

@z

� kLai;largeðC�i;large � Ci;LÞ ¼ 0 (4)

The catalyst particles mass balance is:

@

@z

�ð1� eGÞDS

@CS

@z

�þ @

@z½ðð1� eGÞUP � USLÞCS� ¼ 0 (5)

The heat balance is derived as:

@

@z

�rSLC pSLð1� eGÞDL

@T

@z

�� @ðrSLC pSLUSLTÞ

@z

� UHeatðT � TcoolÞ þ ð1� eGÞð�DHFTrFTÞ ¼ 0 (6)

The boundary conditions for the gas, liquid and slurry at the

inlet (bottom) of the reactor are Danckwerts’ type:

USLCi;L � ð1� eGÞDL

@Ci;L

@z¼ 0 (7)

UG;smallCi;G;small � eG;smallDG;small

@Ci;G;small

@z¼ UG;smallCi;G;inlet

(8)

UG;largeCi;G;large � eG;largeDG;large

@Ci;G;large

@z¼ UG;largeCi;G;inlet

(9)

ð1� eGÞDS

@CS

@zþ ðð1� eGÞUP � USLÞCS þ USLCaverage ¼ 0

(10)

USLrSLT � rSLð1� eGÞDL

@T ¼ ðUSLrSLTÞ0 (11)

@z

The boundary conditions at the outlet of the reactor are:

@Ci;L

@z¼ 0 (12)

@Ci;G;small

@z¼ 0 (13)

@Ci;G;large

@z¼ 0 (14)

@T

@z¼ 0 (15)

The change of gas velocity due to gas consumption was

estimated using the total gas-phase mass balance at each level

in the reactor. The gas velocity of the small bubbles was taken

from de Swart (1996) and the gas velocity of the large bubbles

was calculated from:

UG;large ¼ UG � UG;small (16)

The pressure profile was obtained from (Deckwer, 1992):

@P

@zþ ½ð1� eGÞrSL þ eGrG�g ¼ 0 (17)

2.2. Hydrodynamics and mass transfer correlations

The knowledge of the hydrodynamics and mass transfer

characteristics of both bubble classes is essential (de Swart and

Krishna, 2002; Gupta et al., 2001; Shah et al., 1985) for

modeling SBCRs. The available literature correlations sum-

marized by Behkish et al. (2006) and Lemoine et al. (2005)

along with those proposed by Krishna et al. (1997), Koide et al.

(1984) and Fukuma et al. (1987) could be used to predict the

hydrodynamics and mass transfer parameters for gases in F–T

liquids. These correlations considered the effect of several

variables, such as liquid/solid properties, operating conditions,

and reactor diameter, on these parameters. In our research

laboratories, Behkish et al. (2005, 2006) and Lemoine et al.

(2005) have recently developed Back Propagation Neural

Network (BPNN) correlations as well as empirical correlations

for predicting the gas holdup, Sauter mean bubble diameter and

volumetric mass transfer coefficients in BCRs and SBCRs. As

can be seen in Table 1, the empirical correlations by these

authors take into account not only the liquid/solid properties,

operating conditions, and reactor diameter similar to available

literature correlations, but also the gas distributor type, number/

size of nozzles, catalyst loading, and foamability of the F–T

products, which have been shown to dramatically impact the

hydrodynamics and mass transfer characteristics in SBCRs.

Thus, due to their versatility and proven applicability, the

correlations given in Table 1 were used in the computer model

developed in this study in order to predict the effects of reactor

geometry, superficial gas velocity and catalyst concentration on

the performance of a large-scale SBCR; and to optimize the

reactor geometry as well as operating conditions to produce

10,000 bbl/day of liquid hydrocarbons.

Table 1

Gas holdup, Sauter mean bubble diameter and mass transfer coefficient correlations used in the model (Lemoine et al., 2005; Behkish et al., 2006)

Total gas holdup (Behkish et al., 2006)

eG ¼ 4:94� 10�3 r0:42L r0:18

G

m0:17L s0:27

L

� �U0:55

G

PT

PT � PS

� �0:20 DC

DC þ 1

� ��0:12

G 0:05 expð�2:23CV � 0:16ðrPdPÞ � 0:24XWÞ(18)

Gas holdup of large gas bubbles (Behkish et al., 2006)

eG;large ¼ e0:84G 1� 3:04� 10�6 r0:97

L

m0:16L

e4:50XW�4:49CV

� �¼ e0:84

G ðFÞ(19)

Gas holdup of small gas bubbles (Behkish et al., 2006)

If eG is > ðFÞ25=4 eG;small ¼ eG � eG;large

If eG is > ðFÞ25=4small gas bubbles do not exist

(20)

Total Sauter mean bubble diameter (Lemoine et al., 2005)

dS ¼ 37:19� m0:08L s1:22

L r0:02G T1:66

r1:52L M0:12

W-Gas

U0:14G

DC

DC þ 1

� �0:30

ð1� eGÞ1:56G�0:02e�2:29XWþ2:81CVþ2:77rPdP

(21)

Sauter mean bubble diameter of large gas bubbles (Lemoine et al., 2005)

If eG is> ðFÞ25=4 dS;large ¼ d0:96S ð1� 10�5r0:22

L m0:03L s8:60

L U0:04G e2:37

G e2:74G;largeÞ

(22)

Sauter mean bubble diameter of small gas bubbles (Lemoine et al., 2005)

If eG is > ðFÞ25=4 eG;small

dS;small

¼ eG

dS

� eG;large

dS;large

(23)

Mass transfer coefficient (Lemoine et al., 2005)

kLa

ð1� eGÞ¼ 6:14� 104 r0:26

L m0:12L

s0:52L r0:06

G

e1:21G

U0:12G d0:05

S

D0:50AB

T0:68G0:11 DC

DC þ 1

� �0:40(24)

Table 3

Upper and lower limits of the variables used in Eqs. (18)–(26) (Lemoine et al.,

2005)

Variables Minimum value Maximum value

PT (MPa) 0.1 19.8

PS (MPa) 0.0 0.7�3

L. Sehabiague et al. / Journal of the Chinese Institute of Chemical Engineers 39 (2008) 169–179 173

In Table 1, XW designates the concentration of the primary

liquid in a liquid mixture, and its value varies between 0.5 and

1. For a single-component or an organic liquid mixture,

consisting of different chain length hydrocarbons, such as F–T

products, XW equals 1. Also, G which represents the gas sparger

type is defined as:

G ¼ ðKd � NOdaOÞ (25)

The values of the coefficient Kd and the exponent a for

several distributors are given in Table 2. For perforated plates,

the exponent a depends on z, which is expressed as:

z ¼ NO

�d0

DC

�2

(26)

Table 3 presents the ranges of the conditions of applicability

of Eqs. (18)–(26). It should be noted that the correlations listed

Table 2

Values of a used in Eq. (25) (Lemoine et al., 2005)

Distributor z (%) a Kd (m�a)

PfP <0.055 0.017 1.364

PfP �0.055 and �0.3 0.303 1.364

PfP >0.3 0.293 1.364

M-ON 0.303 1.364

S-ON 0.134 1.205

R, S 0.015 1.000

BC 0.500 1.553

PoP, SP 0.650 1.533

in Table 1 are valid when the volume of internals, commonly

used in SBCRs for cooling or heating purposes, is �20% of the

reactor volume. This is because several literature findings

(Chen et al., 1999; De et al., 1999; Forret et al., 2003; Joseph,

1985; O’Dowd et al., 1987; Saxena et al., 1992; Saxena and

Chen, 1994; Shah et al., 1978; Yamashita, 1987) showed

limited or no effect of internals on the hydrodynamic and mass

transfer parameters as long as their volume fraction remains

under 20%. Also, these correlations should be valid for reactor

UG (m/s) 3.5 � 10 0.574

CV (vol.%) 0 0.36

XW (wt.%) 0.5 1

T (K) 275 538

MW-Gas (kg/kmol) 2 44

DAB (109 m2/s) 2.78 � 10�8 1.25 � 10�11

dp (10�6 m) 5 300

rP (kg/m3) 700 4000

rG (kg/m3) 0.06 223.77

rL (kg/m3) 633 1583

mL (10�3 Pa s) 0.16 398.80

sL (10�3 N/m) 8.4 75

DC (m) 0.0382 7.6200

G 0.06 143800

z (%) 0.0097 75

Table 4

Values for the kinetics constants in Eqs. (34) and (35)

Constant Value Units

A1 8.037 � 10�9 mol/kgcatalyst Pa2 s

E1 37369 J/mol

A2 1.243 � 10�12 Pa�1

E2 �68478 J/mol

Table 5

Coefficients for Henry’s law constant from Soriano (2005)

Gas A B H0

N2 �2.214 � 105 1,267 51.62

He �5.442 � 105 3,336 53.17

H2 �1.929 � 105 1,345 42.18

CO �3.651 � 105 1,873 22.87

L. Sehabiague et al. / Journal of the Chinese Institute of Chemical Engineers 39 (2008) 169–179174

height/inside diameter ratio (L/DC) varying from 4 to 20,

because the considerable data available in the literature

(Eickenbusch et al., 1995; Guy et al., 1986; Moustiri et al.,

2001; Pino et al., 1992; Shah et al., 1982) used to develop these

correlations cover such an L/DC range. It should be mentioned

also that Lemoine et al. (2005) have argued that the use of Eq.

(24) to calculate kLa for small and large gas bubbles using their

corresponding gas holdup and Sauter mean bubble diameter is a

more accurate approach than the ones used by de Swart and

Krishna (2002) and Grund et al. (1992); because it underscores

the fact that the mass transfer behavior of SBCRs is controlled

by the gas-liquid interfacial area (Behkish et al., 2002; Inga and

Morsi, 1997; Lemoine et al., 2004).

All the correlations listed in Table 1 were used in the

simulator according to the algorithm developed by Lemoine

et al. (2005) with the exception of the mass transfer coefficients

of small gas bubbles which was obtained from:

kLai;small ¼ kLai � kLai;large (27)

2.3. Heat transfer

The heat transfer coefficient in the slurry-side (h0) was taken

from Cho et al. (2002):

h0 ¼ 11710U0:445G ðmSL � 103Þ�0:060ðP� 10�6Þ0:176

(28)

2.4. Dispersion coefficients

Baird and Rice (1975) correlated the liquid-phase dispersion

coefficient in bubble columns (DL) using a large number of

experimental data reported by various authors as:

DL ¼ 0:35D4=3C ðgUGÞ0:33

(29)

Since the small gas bubbles are easily entrained by the liquid

recirculation in the reactor under given operating conditions, it

was assumed (de Swart, 1996) that they have the same

dispersion as the liquid phase:

DG;small ¼ DL (30)

For the large gas bubbles, on the other hand, the following

dispersion coefficient (Field and Davidson, 1980; Deckwer and

Schumpe, 1993) was used:

DG;large ¼ 5:64� 10�3

�UG

eG;large

�3:56

D4=3C (31)

The particles settling velocity and solid dispersion

coefficient were estimated using the data/correlations given

by several authors (Kato et al., 1972; Kojima et al., 1986;

O’Dowd et al., 1987; Smith and Ruether, 1985).

The terminal particles settling velocity was calculated from

Ramachandran and Chaudhari (1983):

UTS ¼ðrS � rLÞgD2

C

18mL

(32)

2.5. Kinetics

The F–T reaction rate using cobalt catalyst was taken from

Satterfield and Yates (1991):

rFT ¼ kFT

HeH2CL;H2

HeCOCL;CO

ð1þ a1HeCOCL;COÞ2(33)

with

kFT ¼ A1e�E1=RT (34)

a1 ¼ A2e�E2=RT (35)

The constant in Eqs. (34) and (35) were estimated using the

data from Satterfield and Yates (1991) and are given in Table 4.

2.6. Gas solubilities

The gas solubilities in the liquid phase were determined

from the work by Soriano (2005) who measured the solubilities

of H2, CO, N2 and He in Sasol wax in a small agitated autoclave

under Fischer–Tropsch conditions. The Henry’s law constant

was expressed as a function of temperature as:

He ¼ H0 e½Að1=TÞ2þBð1=TÞ� (36)

The coefficients of Eq. (36) are listed in Table 5.

The solubilities of other gases (H2O, CO2, CH4, etc., . . .)were estimated from Marano and Holder (1997).

2.7. Flash calculations

The multi-component vapor–liquid equilibria (VLE) model

developed by Marano and Holder (1997), along with the

assumption that the liquid products follow a classical Shultz-

Flory distribution (Satterfield and Huff, 1982) with a growth

probability factor of 0.92 (Jager and Espinoza, 1995), was used

to perform equilibrium calculations in order to estimate the

final composition of the gaseous and liquid hydrocarbon

products.

Table 6

Operating variables used in the simulator for F–T SBCR

System H2/CO(2/1)-Al2O3 supported cobalt catalyst

Reactor and sparger geometry L (m) 30–50

DC (m) 6–8

NO 1033–1367

dorf. (m) 0.065–0.075

Sparger type M-ON

Cooling pipes Number of cooling pipes 275–489

Cooling pipes outside diameter (m) 0.114

Operating variables T (K) 500

P (MPa) 3

UG (m/s) 0.15–0.45

USL (m/s) 0.01

dP (mm) 42

rP (kg/m3) 3218

CW (wt.%) 25–50

Fig. 2. Effect of reactor geometry on the performances of the F–T SBCR (UG = 0.3 m/s, CS = 37.5 wt.%).

L. Sehabiague et al. / Journal of the Chinese Institute of Chemical Engineers 39 (2008) 169–179 175

L. Sehabiague et al. / Journal of the Chinese Institute of Chemical Engineers 39 (2008) 169–179176

3. Results and discussion

All the partial differential equations of the material and

energy balances for all components given above along with

equation parameters and boundary conditions were incorpo-

rated in a user-friendly simulator and simultaneously solved

numerically using Athena Visual Studio Software version 11.0

with FORTRAN compiler g95. The simulator was then used to

study the effects of different operating variables, including

reactor inside diameter, reactor height, superficial gas velocity,

and catalyst concentration on the performance of a SBCR

operating in the churn-turbulent flow regime. The ranges of all

operating variables in addition to sparger type/geometry,

catalyst particle size/density, slurry superficial velocity, and

cooling pipes number/size used in the simulator are given in

Table 6.

The performance of the reactor was evaluated using the CO

conversion, the liquid hydrocarbon yield, space time yield

(STY) and catalyst productivity. The STY is defined similar to

de Swart (1996) as the mass of liquid hydrocarbons produced

(kg) per reactor volume (m3) per unit time (hour). The catalyst

productivity was also defined as the mass of liquid

Fig. 3. Effect of superficial gas velocity and catalyst concentration

hydrocarbons produced (kg) per mass of catalyst in the reactor

(kg) per unit time (hour).

The effects of reactor inside diameter (ID) and height (L) on

the performances of an F–T SBCR operating at constant

superficial gas velocity (UG = 0.3 m/s) and catalyst concentra-

tion (CS = 37.5 wt.%) are shown in Fig. 2. As can be seen in

Fig. 2(a), CO conversion appears to increase by more than 50%

with increasing reactor height from 30 to 50 m. This behavior

agrees with that by de Swart (1996) who reported an increase of

CO conversion with increasing the height of a 7.5 m ID SBCR

from 10 to 30 m. Fig. 2(a) also shows that CO conversion

slightly decreases with increasing reactor ID. The effect of

reactor ID and height on the hydrocarbon yield, depicted in

Fig. 2(b), can be directly related to their effects on CO

conversion. The increase of hydrocarbon yield with increasing

reactor ID and height was expected since larger diameter and

height leads to larger gas throughput and consequently greater

hydrocarbon production.

Fig. 2(c) and (d) show the effects of reactor height and ID on

the space time yield (STY) and the catalyst productivity. As can

be seen in these figures, at constant reactor length, both STY

and catalyst productivity appear to slightly decrease with

on the performances of the F–T SBCR (DC = 7 m, L = 30 m).

Table 7

Geometry and operating conditions of the optimized F–T SBCR to produce a

10,000 bbl/day of liquid hydrocarbons

Units Value

Column inside diameter (DC) m 7

Column length (L) m 30

Superficial gas velocity (UG) m/s 0.255

Catalyst loading (CW) wt.% 37.5

CO conversion % 60.6%

Yield bbl/day 10,808

Space time yield (STY) kgHC=m3reactor h 55.04

Catalyst productivity kgHC/kgcatalyst h 0.225

L. Sehabiague et al. / Journal of the Chinese Institute of Chemical Engineers 39 (2008) 169–179 177

increasing reactor ID; whereas at constant reactor diameter,

both STY and catalyst productivity appear to increase with

reactor length by about 5% and then level off after a reactor

height of 40 m. This can be related to the fact that the overall F–

T reaction rate and accordingly the volumetric hydrocarbons

production rate is expected to decrease due to the predictable

decrease of the syngas velocity, CO and H2 partial pressures,

and catalyst concentrations with the reactor height. Thus, Fig. 2

leads one to conclude that in order to produce about 10,000 bbl/

day of liquid hydrocarbons, a SBCR with an internal diameter

of 7 m and a height of 30 m, could be used.

The effects of the superficial gas velocity and catalyst

concentration on the performance of a SBCR with the geometry

given above (ID = 7 m and 30 m height) are illustrated in Fig. 3.

As can be seen in Fig. 3(a), the CO conversion decreases with

increasing superficial gas velocity suggesting a small increase

of the mass transfer rate compared to the decrease of the

residence time of the gaseous reactants (CO and H2) which

becomes too short for converting more CO. Fig. 3(a) also shows

that increasing catalyst concentration, on the other hand,

increases the CO conversion. For low gas velocity (�0.15 m/s)

CO conversion appears to level off at high catalyst loading. This

can be attributed to the fact that increasing catalyst concentra-

tion above 40 wt.% (or 13 vol.%) decreases the mass transfer

coefficients (kLa) and consequently the mass transfer rate,

which drives the SBCR to operate in a mass transfer-controlled

regime, leading to a maximum of CO conversion (�86%). This

effect of catalyst concentration on the CO conversion agrees

with the findings by Inga and Morsi (1996) who simulated a

4.8 m ID and 12 m height F–T SBCR using iron catalyst and

reported that the syngas conversion reached a maximum and

then decreased with increasing catalyst loading. It should be

pointed out that few studies (Krishna and Sie, 2000; van der

Laan et al., 1999) found a plateau of conversion at lower

superficial gas velocities when increasing solid concentration;

and other authors (de Swart, 1996; Fernandez, 2006; Maretto

and Krishna, 1999), who did not find a maximum of conversion

when increasing catalyst concentration up to 35 vol.%. This

could be related to the difference in the catalyst size, density

and activity as well as the absence of catalyst concentration in

the mass transfer coefficients correlations used by these

authors. Fig. 3(a) also shows that increasing gas velocity will

shift the region of mass transfer controlled regime towards

higher catalyst concentrations. Fig. 3(b) illustrates the effects

on superficial gas velocity and catalyst concentration on the

liquid hydrocarbon yield, and as can be observed at constant

superficial gas velocity, the liquid hydrocarbon yield increases

with increasing catalyst concentration, which is similar to the

behavior of the CO conversion. At low catalyst concentration

(<35 wt.%), the hydrocarbon yield decreases with increasing

the superficial gas velocity, whereas, for catalyst concentrations

greater than 35 wt.%, the liquid hydrocarbons yield appears to

increase with superficial gas velocity, reach a maximum and

then decrease.

Fig. 3(c) shows the effects of catalyst concentration and

superficial gas velocity on the space time yield (STY); and as

expected the STY follows the same trend of the yield because

the reactor volume under these conditions is constant. Fig. 3(d)

illustrates the effects of catalyst concentration and superficial

gas velocity on the catalyst productivity; and as can be observed

the catalyst productivity reaches a maximum at catalyst

concentration of 25 wt.% and superficial gas velocities between

0.2 and 0.25 m/s; and generally decreases with increasing

catalyst concentration up to 40 wt.%. At catalyst concentrations

>40 wt.%, however, the catalyst productivity appears to

increase with increasing the superficial gas velocity. At

superficial gas velocities >0.25 m/s, the catalyst productivity

increases with catalyst concentrations, levels off, and then

decreases. From these four plots, Fig. 3 clearly shows that

maximum STY can be achieved using operating conditions

(high catalyst loading and high superficial gas velocities) which

are fundamentally different from those needed to obtain

maximum catalyst productivity (low catalyst loading and

average superficial gas velocities).

Thus, Fig. 3 leads to conclude that it is possible to produce a

10,000 bbl/day of liquid hydrocarbons using different operating

conditions; however, since the conditions for maximum STY

would require greater capital and operating costs than those for

maximum catalyst productivity, the conditions corresponding

to the maximum productivity should be used.

Table 7 summarizes the geometry and operating conditions

of the optimized F–T SBCR with the ancillaries given in

Table 6 in order to produce a 10,000 bbl/day of liquid

hydrocarbons. Table 7 also shows the performance character-

istic of this SBCR.

4. Conclusions

The simulator predictions for F–T SBCR indicated the

following.

At constant superficial gas velocity (0.3 m/s) and catalyst

concentration (37.5 wt.%), the CO conversion and liquid

hydrocarbon yield increased with increasing reactor height.

Also, the CO conversion slightly decreased with increasing

reactor ID, whereas the hydrocarbon yield increased with

increasing reactor ID. Both the space time yield and reactor

productivity increased with reactor length by about 5% and then

leveled off after a reactor height of 40 m. Also, both space-

time-yield and reactor productivity appeared to slightly

decrease with increasing reactor ID.

L. Sehabiague et al. / Journal of the Chinese Institute of Chemical Engineers 39 (2008) 169–179178

At constant reactor inside diameter (7 m) and height (30 m),

the CO conversion decreased with increasing the superficial gas

velocity and increased with catalyst concentration. Increasing the

catalyst concentration above 40 wt.% drove the SBCR to operate

in a mass transfer-controlled regime. Also, the liquid hydro-

carbon yield and space time yield increased with increasing

catalyst concentration at constant superficial gas velocity. At low

catalyst concentration (<35 wt.%), both the hydrocarbon yield

and space time yield decreased with increasing the superficial gas

velocity, whereas, for catalyst concentrations >35 wt.%, the

liquid hydrocarbons yield increased with superficial gas velocity,

reached a maximum and then decreased. The catalyst

productivity reached a maximum at the lowest catalyst

concentration of 25 wt.% and superficial gas velocities between

0.2 and 0.25 m/s, and generally decreased with increasing

catalyst concentration up to 40 wt.%. At catalyst concentrations

greater than 40 wt.%, however, the catalyst productivity

increased with superficial gas velocity. At superficial gas

velocities >0.25 m/s, the catalyst productivity increased with

catalyst concentrations, levels off, and then decreased. Thus, the

maximum space time yield occurred at the highest superficial gas

velocity and catalyst concentration, whereas the maximum

catalyst productivity occurred at a medium superficial gas

velocity and the lowest catalyst concentration.

In order to produce 10,000 bbl/day of F–T liquid hydro-

carbons, a SBCR (7 m ID and 30 m height), operating with

37.5 wt.% of cobalt-supported catalyst at 0.255 m/s superficial

gas velocity, 3 MPa, and 500 K, is required. The SBCR should

be provided with (375) cooling pipes representing 10% of the

total reactor volume, and a multi-nozzle distributor with (1200)

nozzles. The reactor would produce 10,808 bbl/day of liquid

hydrocarbons at CO conversion of 60.6% with space time yield

of 55:04 kgHC=m3reactor h and catalyst productivity of

0.225 kgHC/kgcatalyst h.

References

Baird, M. H. I. and R. G. Rice, ‘‘Axial Dispersion in Large Unbaffled

Columns,’’ Chem. Eng. J., 9 (2), 171 (1975).

Behkish, A., ‘‘Hydrodynamic and Mass Transfer Parameters in Large-Scale

Slurry Bubble Column Reactors,’’ Ph.D. Thesis, University of Pittsburgh,

Pittsburgh, U.S.A. (2004).

Behkish, A., R. Lemoine, R. Oukaci, and B. I. Morsi, ‘‘Novel Correlations for

Gas Holdup in Large-Scale Slurry Bubble Column Reactors Operating

under Elevated Pressures and Temperatures,’’ Chem. Eng. J., 115 (3), 157

(2006).

Behkish, A., R. Lemoine, L. Sehabiague, R. Oukaci, and B. I. Morsi, ‘‘Pre-

diction of the Gas Holdup in Industrial-Scale Bubble Columns and Slurry

Bubble Column Reactors Using Back-Propagation Neural Networks,’’ Int.

J. Chem. Reactor Eng., 3 (A53) (2005).

Behkish, A., R. Lemoine, L. Sehabiague, R. Oukaci, and B. I. Morsi, ‘‘Gas

Holdup and Bubble Size Behavior in a Large-Scale Slurry Bubble Column

Reactor Operating with an Organic Liquid under Elevated Pressures and

Temperatures,’’ Chem. Eng. J., 128 (2–3), 69 (2007).

Behkish, A., Z. Men, J. R. Inga, and B. I. Morsi, ‘‘Mass Transfer Characteristics

in a Large-Scale Slurry Bubble Column Reactor with Organic Liquid

Mixtures,’’ Chem. Eng. Sci., 57 (16), 3307 (2002).

Chen, J., F. Li, S. Degaleesan, P. Gupta, M. H. Al-Dahhan, M. P. Dudukovic, and

B. A. Toseland, ‘‘Fluid Dynamic Parameters in Bubble Columns with

Internals,’’ Chem. Eng. Sci., 54 (13–14), 2187 (1999).

Cho, Y. J., J. W. Kwang, K. Yong, and D. K. Sang, ‘‘Dynamic Characteristics of

Heat Transfer Coefficient in Pressurized Bubble Columns with Viscous

Liquid Medium,’’ Chem. Eng. Process, 41 (8), 699 (2002).

De, S. K., S. Ghosh, R. K. Parichha, and P. De, ‘‘Gas Hold-Up in Two-Phase

System with Internals,’’ Indian Chem. Eng., 41 (2), 112 (1999).

de Swart, J. W. A., ‘‘Scale-Up of a Fischer-Tropsch Slurry Reactor,’’ Ph.D.

Thesis, University of Amsterdam, Amsterdam, Netherlands (1996).

de Swart, J. W. A. and R. Krishna, ‘‘Simulation of the Transient and Steady

State Behavior of a Bubble Column Slurry Reactor for Fisher-Tropsch

Synthesis,’’ Chem. Eng. Process, 41 (1), 35 (2002).

de Swart, J. W. A., R. E. van Vliet, and R. Krishna, ‘‘Size, Structure and

Dynamics of ‘‘Large’’ Bubbles in a Two-Dimensional Slurry Bubble

Column,’’ Chem. Eng. Sci., 51 (20), 4619 (1996).

Deckwer, W. -D. , Bubble Column Reactors, John Wiley & Sons, New York,

U.S.A. (1992).

Deckwer, W.-D. Y. Louisi, M. Ralek, and A. Zaidi, ‘‘Hydrodynamic Properties

of the Fischer-Tropsch Slurry Process,’’ Ind. Eng. Chem. Process Des. Dev.,

19 (4), 699 (1980).

Deckwer, W.-D. and A. Schumpe, ‘‘Improved Tools for Bubble Column

Reactor Design and Scale-Up,’’ Chem. Eng. Sci., 48 (5), 889 (1993).

Dry, M. E., ‘‘The Fischer-Tropsch Process: 1950–2000,’’ Catal. Today, 71 (3–

4), 227 (2002).

Dry, M. E., ‘‘Present and Future Applications of the Fischer-Tropsch Process,’’

Appl. Catal. A: Gen., 276 (1–2), 1 (2004).

Eickenbusch, H., P. O. Brunn, and A. Schumpe, ‘‘Mass Transfer into Viscous

Pseudoplastic Liquid in Large-Diameter Bubble Columns,’’ Chem. Eng.

Process, 34 (5), 479 (1995).

Fernandez, F. A. N., ‘‘Modeling and Product Grade Optimization of Fischer-

Tropsch Synthesis in a Slurry Reactor,’’ Ind. Eng. Chem. Res., 45 (3), 1047

(2006).

Field, R. W. and J. F. Davidson, ‘‘Axial Dispersion in Bubble Columns,’’ Trans.

Inst. Chem. Eng., 58 (4), 228 (1980).

Forret, A., Schweitzer, J.-M. T. Gauthier, R. Krishna, and D. Schweich, ‘‘Liquid

Dispersion in Large Diameter Bubble Columns, with and Without

Internals,’’ Can. J. Chem. Eng., 81 (3–4), 360 (2003).

Fukuma, M., K. Muroyama, and A. Yasunishi, ‘‘Properties of Bubble Swarn in

a Slurry Bubble Column,’’ J. Chem. Eng. Jpn., 20 (1), 28 (1987).

Grund, G., A. Schumpe, and W. D. Deckwer, ‘‘Gas-Liquid Mass Transfer in a

Bubble Column with Organic Liquids,’’ Chem. Eng. Sci., 47 (13–14), 3509

(1992).

Gupta, P., M. H. Al-Dahhan, M. P. Dudukovic, and B. A. Toseland, ‘‘Compar-

ison of Single- and Two-Bubble Class Gas-Liquid Recirculation Models-

Application to Pilot-Plant Radioactive Tracer Studies during Methanol

Synthesis,’’ Chem. Eng. Sci., 56 (3), 1117 (2001).

Guy, C., P. J. Carreau, and J. Paris, ‘‘Mixing Characteristics and Gas Holdup of

a Bubble Column,’’ Can. J. Chem. Eng., 64 (1), 23 (1986).

Inga, J. R. and B. I. Morsi, ‘‘A Novel Approach for the Assessment of the Rate-

Limiting Step in Fischer-Tropsch Slurry Process,’’ Energy Fuels, 10 (3),

566 (1996).

Inga, J. R. and B. I. Morsi, ‘‘Effect of Catalyst Loading on Gas/Liquid Mass

Transfer in a Slurry Reactor: A Statistical Experimental Approach,’’ Can. J.

Chem. Eng., 75 (5), 872 (1997).

Jager, B. and R. Espinoza, ‘‘Advances in Low Temperature Fischer-Tropsch

Synthesis,’’ Catal. Today, 23 (1), 17 (1995).

Joseph, S., ‘‘Hydrodynamic and Mass Transfer Characteristics of a

Bubble Column,’’ Ph.D. Thesis, University of Pittsburgh, Pittsburgh,

U.S.A. (1985).

Kato, Y., A. Nishiwaki, T. Fukuda, and S. Tanaka, ‘‘The Behavior of Suspended

Solid Particles and Liquid in Bubble Columns,’’ J. Chem. Eng. Jpn., 5 (2),

112 (1972).

Koide, K., A. Takazawa, M. Komura, and H. Matsunaga, ‘‘Gas Holdup and

Volumetric Liquid-Phase Mass Transfer Coefficient in Solid-Suspended

Bubble Columns,’’ J. Chem. Eng. Jpn., 17 (5), 459 (1984).

Kojima, H., H. Anjyo, and Y. Mochizuki, ‘‘Axial Mixing in Bubble Column

with Suspended Solid Particles,’’ J. Chem. Eng. Jpn., 19 (3), 232 (1986).

Krishna, R., J. W. A. de Swart, J. Ellenberg, G. B. Martina, and C. Maretto,

‘‘Gas Holdup in Slurry Bubble Columns: Effect of Column Diameter and

Slurry Concentrations,’’ AIChE J., 43, 311 (1997).

L. Sehabiague et al. / Journal of the Chinese Institute of Chemical Engineers 39 (2008) 169–179 179

Krishna, R. and S. T. Sie, ‘‘Design and Scale-Up of the Fischer-Tropsch

Bubble Column Slurry Reactor,’’ Fuel Process. Technol., 64 (1),

73 (2000).

Krishna, R., M. I. Urseanu, and A. J. Dreher, ‘‘Gas Holdup in Bubble Columns:

Influence of Alcohol Addition Versus Operation at Elevated Pressures,’’

Chem. Eng. Process, 39 (4), 371 (2000).

Lee, S.-Y. and Y. P. Tsui, ‘‘Succeed at Gas/Liquid Contacting,’’ Chem. Eng.

Progr., 95 (7), 23 (1999).

Lemoine, R., A. Behkish, and B. I. Morsi, ‘‘Hydrodynamic and Mass Transfer

Characteristics in Organic Liquid Mixtures in a Large-Scale Bubble Col-

umn Reactor for the Toluene Oxidation Process,’’ Ind. Eng. Chem. Process

Des. Dev., 43 (19), 6195 (2004).

Lemoine, R., A. Behkish, L. Sehabiague, Y. Heintz, R. Oukaci, and B. I. Morsi,

‘‘An Algorithm for Predicting the Hydrodynamic and Mass Transfer Para-

meters in Slurry Bubble Column Reactors for Fischer-Tropsch Synthesis,’’

In: Proceedings of the Twenty Second Annual International Pittsburgh Coal

Conference, Pittsburgh, PA, September 12–16 (2005).

Lemoine, R., ‘‘Hydrodynamics, Mass Transfer and Modeling of the Liquid-

Phase Toluene Oxidation Process,’’ Ph.D. Thesis, University of Pittsburgh,

Pittsburgh, U.S.A. (2005).

Magaud, F., M. Souhar, G. Wild, and N. Boisson, ‘‘Experimental Study

of Bubble Column Hydrodynamics,’’ Chem. Eng. Sci., 56 (15), 4597

(2001).

Marano, J. J. and G. D. Holder, ‘‘Characterization of Fischer-Tropsch Liquids

for Vapor-Liquid Equilibria Calculations,’’ Fluid Phase Equilib., 138 (1–2),

1 (1997).

Maretto, C. and R. Krishna, ‘‘Modelling of a Bubble Column Slurry Reactor for

Fischer-Tropsch Synthesis,’’ Catal. Today, 52 (2–3), 279 (1999).

Moustiri, S., G. Hebrard, S. S. Thakre, and M. Roustan, ‘‘A Unified Correlation

for Predicting Liquid Axial Dispersion Coefficient in Bubble Columns,’’

Chem. Eng. Sci., 56 (3), 1041 (2001).

Nigam, K. D. P. and A. Schumpe, Three-Phase Sparged Reactors, Gordon and

Breach Publishers, Amsterdam, The Netherlands (1996).

O’Dowd, W., D. N. Smith, J. A. Ruether, and S. C. Saxena, ‘‘Gas and Solids

Behavior in a Baffled and Unbaffled Slurry Bubble Column,’’ AIChE J., 33(12), 1959 (1987).

Oshinowo, T. and M. Charles, ‘‘Vertical Two-Phase Flow. Part I. Flow Pattern

Correlations,’’ Can. J. Chem. Eng., 52 (1), 25 (1974).

Pino, L. Z., R. B. Solari, S. Siquier, L. Antonio Estevez, and M. M. Yepez,

‘‘Effect of Operating Conditions on Gas Holdup in Slurry Bubble Columns

with a Foaming Liquid,’’ Chem. Eng. Commun., 117 (1), 367 (1992).

Rados, N., M. H. Al-Dahhan, and M. P. Dudukovic, ‘‘Modeling of the Fischer–

Tropsch Synthesis in Slurry Bubble Column Reactors,’’ Catal. Today, 79–80, 211 (2003).

Ramachandran, P. A. and R. V. Chaudhari, Three Phase Catalytic Reactors,

Gordon and Breach, New York, NY, U.S.A. (1983).

Sarrafi, A., M. Jamialahmadi, H. Muller-Steinhagen, and J. M. Smith, ‘‘Gas

Holdup in Homogeneous and Heterogeneous Gas-Liquid Bubble Column

Reactors,’’ Can. J. Chem. Eng., 77 (1), 11 (1999).

Satterfield, C. N. and G. A. Huff, ‘‘Product Distribution from Iron Catalyst in

Fischer-Tropsch Slurry Reactors,’’ Ind. Eng. Chem. Process Des. Dev., 21(3), 465 (1982).

Satterfield, C. N. and I. C. Yates, ‘‘Intrinsic Kinetics of the Fischer-Tropsch

Synthesis on a Cobalt Catalyst,’’ Energy Fuels, 5 (1), 168 (1991).

Saxena, S. C., N. S. Rao, and P. R. Thimmapuram, ‘‘Gas Phase Holdup in Slurry

Bubble Column for Two- and Three-Phase Systems,’’ Chem. Eng. J., 49 (3),

151 (1992).

Saxena, S. C. and Z. D. Chen, ‘‘Hydrodynamics and Heat Transfer of Baffled and

Unbaffled Slurry Bubble Columns,’’ Rev. Chem. Eng., 10 (3–4), 195 (1994).

Shah, Y. T., S. Joseph, D. N. Smith, and J. A. Ruether, ‘‘Two-Bubble Class

Model for Churn Turbulent Bubble-Column Reactor,’’ Ind. Eng. Chem.

Process Des. Dev., 24 (4), 1096 (1985).

Shah, Y. T., B. G. Kelkar, S. P. Godbole, and W. -D. Deckwer ‘‘Design

Parameters Estimations for Bubble Column Reactors,’’ AIChE J., 28 (3),

353 (1982).

Shah, Y. T., C. A. Ratway, and H. G. McIlvried, ‘‘Back-Mixing Characteristics

of a Bubble Column with Vertically Suspended Tubes,’’ Trans. Inst. Chem.

Eng., 56 (2), 107 (1978).

Smith, D. N. and J. A. Ruether, ‘‘Dispersed Solid Dynamics in a Slurry Bubble

Column,’’ Chem. Eng. Sci., 40 (5), 741 (1985).

Soriano, J. P., ‘‘Mass Transfer Characteristics in an Agitated Slurry Reactor

Operating Under Fischer-Tropsch Conditions,’’ M.S. Thesis, University of

Pittsburgh, Pittsburgh, U.S.A. (2005).

Steynberg, A. and M. Dry, Fischer-Tropsch Technology, Elsevier Science,

U.S.A. (2004).

van der Laan, G. P., A. A. C. M. Beenackers, and R. Krishna, ‘‘Multicomponent

Reaction Engineering Model for Fe-Catalyzed Fischer-Tropsch Synthesis in

Commercial Scale Slurry Bubble Column Reactors,’’ Chem. Eng. Sci., 54(21), 5013 (1999).

Vermeer, D. and R. Krishna, ‘‘Hydrodynamics and Mass Transfer in Bubble

Columns in Operating in the Churn-Turbulent Regime,’’ Ind. Eng. Chem.

Process Des. Dev., 20 (3), 475 (1981).

Vosloo, A. C., ‘‘Fischer-Tropsch: A Futuristic View,’’ Fuel Process. Technol.,

71 (1–3), 149 (2001).

Wilhelm, D. J., D. R. Simbeck, A. D. Karp, and R. L. Dickenson, ‘‘Syngas

Production for Gas-to-Liquids Applications: Technologies, Issues and

Outlook,’’ Fuel Process. Technol., 71 (1–3), 139 (2001).

Yamashita, F., ‘‘Effects of Vertical Pipe and Rod Internals on Gas Holdup in

Bubble Columns,’’ J. Chem. Eng. Jpn., 20 (2), 204 (1987).

Zimmerman, W. H. and D. B. Bukur, ‘‘Reaction Kinetics over Iron Catalysts

Used for the Fischer-Tropsch Synthesis,’’ Can. J. Chem. Eng., 68 (2), 292

(1990).