Sei sulla pagina 1di 14

A mathematical modeling of catalytic milli-fixed bed

reactor for Fischer–Tropsch synthesis: Influence of tube


diameter on Fischer Tropsch selectivity and thermal
behavior
Giovanni Chabot, Richard Guilet, Patrick Cognet, Christophe Gourdon

To cite this version:


Giovanni Chabot, Richard Guilet, Patrick Cognet, Christophe Gourdon. A mathematical modeling of
catalytic milli-fixed bed reactor for Fischer–Tropsch synthesis: Influence of tube diameter on Fischer
Tropsch selectivity and thermal behavior. Chemical Engineering Science, Elsevier, 2015, 127, pp.72-83.
�10.1016/j.ces.2015.01.015�. �hal-01917338�

HAL Id: hal-01917338


https://hal.archives-ouvertes.fr/hal-01917338
Submitted on 9 Nov 2018

HAL is a multi-disciplinary open access L’archive ouverte pluridisciplinaire HAL, est


archive for the deposit and dissemination of sci- destinée au dépôt et à la diffusion de documents
entific research documents, whether they are pub- scientifiques de niveau recherche, publiés ou non,
lished or not. The documents may come from émanant des établissements d’enseignement et de
teaching and research institutions in France or recherche français ou étrangers, des laboratoires
abroad, or from public or private research centers. publics ou privés.
Open Archive Toulouse Archive Ouverte

OATAO is an open access repository that collects the work of Toulouse


researchers and makes it freely available over the web where possible

This is an author’s version published in: http://oatao.univ-toulouse.fr/20596

Official URL:
https://doi.org/10.1016/j.ces.2015.01.015

To cite this version:


Chabot, Giovanni and Guilet, Richard and Cognet, Patrick
and Gourdon, Christophe A mathematical modeling of
catalytic milli-fixed bed reactor for Fischer–Tropsch synthesis:
Influence of tube diameter on Fischer Tropsch selectivity and
thermal behavior. (2015) Chemical Engineering Science, 127.
72-83. ISSN 0009-2509

Any correspondence concerning this service should be sent


to the repository administrator: tech-oatao@listes-diff.inp-toulouse.fr
A mathematical modeling of catalytic milli-fixed bed reactor
for Fischer–Tropsch synthesis: Influence of tube diameter
on Fischer Tropsch selectivity and thermal behavior
Giovanni Chabot a, Richard Guilet b,n, Patrick Cognet a, Christophe Gourdon a
a
Université de Toulouse, INPT, ENSIACET, Laboratoire de Génie Chimique (LGC UMR 5503), 4 Allée Emile Monso, F-31432 Toulouse, France
b
Université de Toulouse, UPS, Laboratoire de Génie Chimique (LGC UMR 5503), 4 Allée Emile Monso, F-31432 Toulouse, France

H I G H L I G H T S

! A two-dimensional pseudo-homogeneous model in packed bed reactors was built.


! Influence of tube reactor diameter for Fischer–Tropsch synthesis was investigated.
! Increasing tube diameter leads to hot spot within the reactor.
! Liquid hydrocarbons selectivity decreases if isothermal behavior is not fulfilled.
! High productivities in liquid hydrocarbons are obtained at millimeter scale.

a b s t r a c t

A two-dimensional pseudo-homogeneous model has been developed to investigate the influence of tube size
on the thermal behavior and performance of packed fixed bed reactor for the low temperature Fischer–
Tropsch (FT) synthesis over alumina supported cobalt. Velocity, temperature and composition fields are
determined by solving the fundamental transport equations in porous media. Special attention was paid to
the variation of transport properties with temperature and composition of the gas mixture. High dependency
Keywords: of the thermal behavior on the thermal conductivity of the gas mixture is highlighted, whereas viscosity and
Fischer–Tropsch synthesis
heat capacity of gas mixture have very little influence. Moreover for the considered catalyst, simulation results
Fixed bed reactor
have displayed high heat removal for the millimetric scale with a tube inner diameter below 2.75 mm for an
Milli-reactor "1 "1
Intensification extended range of weight hourly space velocity (20–600 gsyngas min kgcat: , T¼ 493 K and P¼ 20 bar). With a
Reactor modeling millimetric reactor, high CO conversion ðX CO 4 90%Þ is obtained for values of space velocity between 30 and
"1
COMSOL multiphysics 120 gsyngas min " 1 kgcat: . For higher diameter tube than 3.11 mm, thermal runaway occurs and even worse, no
convergence achieved due to the very low heat transfer global coefficient and the weak surface/volume ratio,
leading to a significant decrease of liquid fuels selectivity and an increase of light hydrocarbon (C1 to C4)
selectivity up to 14%. To conclude, results from scale-up study with the millimetric scale are outstanding,
"1 "3
more than 2900 kg h mcat: of C5 þ could be produce after numbering-up 3033 tubes of 10 centimeters in
length whereas conventional units (multitubular fixed bed reactors or slurry phase reactors) do not exceed
"1 "3
400 kg h mcat: .

1. Introduction instabilities in oil producing regions. Thus, there is nowadays a strong


need of alternative sources to supply the world demand for liquid
Over the last few years, a major concern has arisen regarding fuels. For several years, first generation biofuels have been produced
the decreasing global oil reserves and the associated increasing crude from the fermentation of sugars or starches to provide bioethanol
oil price both driven by a strong world demand and by political for gasoline engine or else from the transesterification of veget-
able oil to produce biodiesel. However this alternative to fossil
fuels has become very controversial and is not indeed devoid of
n
Corresponding author. Tel.: þ 33 5 34 32 37 18. major drawbacks involving a new demand for lands that had been
E-mail address: richard.guilet@iut-tlse3.fr (R. Guilet). previously devoted to agricultural production and feeding people.
Nevertheless other agricultural feedstock like woodchips and residual (573–623 K, HTFT) and the low temperature FT processes (473–523 K,
of non-food parts of cereal crop, can be valorized and be integrated in LTFT). HTFT based on iron catalysts yields essentially C1–C15 hydro-
a second-generation Biomass to Liquid process (BTL process) to carbons in circulating fluidized-bed reactors while LTFT processes
synthesize liquid biofuels via the Fischer–Tropsch synthesis (FTS). lead mainly to linear long chain hydrocarbons (waxes and paraffins).
This old process, which was developed in the 1920s, kindles a Our study only deals with the LTFT processes for which two main
renewal of interest for providing a method to produce liquid technologies of reactor exist: the slurry-phase reactor and the multi-
hydrocarbons from synthesis gas, the so-called syngas, which is a tubular fixed-bed reactor. Table 1, not intended to be exhaustive, gives
mixture of CO and H2. the main advantages and drawbacks for both of them.
Historically the syngas was obtained from the gasification of Although several drawbacks are identified, MFBR are widely
coal (CTL process), but nowadays most of syngas is produced from used in FT process. These reactors are indeed capable of being
the steam reforming or the partial oxidation of natural gas (GTL easily scaled-up when going from a single tube to multitubular
process). Moreover for a few years the gasification of biomass (BTL reactor to the extent that a good inlet distribution can be ensured.
process) has been developed as a sustainable route. The formers Moreover, many preliminary studies are carried out in packed-bed
CTL and GTL processes are based on fossil resources whereas the reactors at lab-scale such as catalysts developments, kinetic mea-
latter technology BTL appears to be the most environmentally surements and deactivation studies. Nevertheless, within years of
friendly technology owing to the great abundance of second- development and improvement, some of these drawbacks have
generation feedstocks which do not compete with the production been reduced especially those regarding heat removal concerns.
of food crop. However, one of the main limitations of MFBR lies in high pressure
More generally, the conversion of syngas via FT synthesis drops that can arise when decreasing the catalyst particles dia-
provides a wide product spectrum consisting of methane and a meter. No such limitation occurs in slurry phase reactors allowing
complex multi-component mixture predominantly of linear hydro- consequently the use of particles as small as possible and higher
carbons (n-alkanes, n-alkenes) and oxygenates. Formation of paraf- effectiveness factors of the catalyst. Actually, particles diameters
finic and olefinic compounds are respectively illustrated by the range from about 100 μm when using a slurry reactor to about the
chemical equations (Eqs. (1) and (2)) whereas carbon dioxide is millimeter for a MFBR. The use of smaller catalyst pellets in MFBR
produced by the equilibrium reaction Water–Gas-Shift (Eq. (3)). can obviously increase the effectiveness factor by reducing the
Actually, the reaction mechanism featuring the chain growth diffusion limitations but the catalyst activity rises thereby and heat
follows a polymerization-like scheme based on sequential additions removal might be difficult to process using conventional reactors.
of the monomer CH2. The product range can be described by Alternatively, the catalyst and/or the inlet reactor feed need to be
Anderson–Schultz–Flory (ASF) distribution, characterized by the highly diluted to prevent hot spot formation. On the other hand,
chain growth probability parameter α which can be considered as process intensification that has emerged for the recent years has
independent on the chain length n (Sadeqzadeh et al., 2012): demonstrated the ability of new heat-exchanger reactors to operate
nCOþ ð2n þ 1ÞH2 -Cn H2n þ 2 þ nH2 O ð1Þ in a new way with enhanced mass and heat transfers mainly due to
a large surface to volume ratio. Actually, channel size mostly leads
nCOþ ð2nÞH2 -Cn H2n þ nH2 O ð2Þ to laminar flow but the reduction of the diameter to millimeter
allows avoiding unexpected transfer limitations. Milli-structured
COþH2 O⇌CO2 þ H2 ð3Þ reactors allow for a strong increase of heat transport in radial
direction and thus isothermal operation, even for highly exothermic
reactions, becomes feasible.
2. Background and objectives Thus, recent works have shown that microchannel reactors and
milli-fixed bed reactors were an alternative to conventional fixed-
2.1. Background bed reactors, i.e. based on centimeter-scale tubes assembly, to
operate FT synthesis. By greatly reducing the size and cost of
FT synthesis is known for its highly exothermicity ðΔH R Þ ¼ chemical processing hardware, micro and/or milli-channel process
"1
"165 kJ molCO Þ, therefore one of the most important concerns for technology holds the potential to enable cost effective production of
the development of commercial FT reactors is their necessary heat synthetic fuels in smaller scale facilities. In particular, Velocys has
removal capability that has a major impact on the products developed a microchannel FT reactor. They announced a production
selectivity: a temperature increase has the effect of rising methane from 28 to 40 barrels per day/full reactor with productivity several
production as well as results in catalyst deactivation associated to times higher than conventional process developed by Sasol or Shell.
sintering and coking. Four main types of commercial FT reactors Cao et al. (2009) have performed experiments in microchannel
are commonly implemented in industrial processes: the fluidized-bed reactor system with active cooling to maintain isothermal condi-
reactor, the multitubular fixed-bed reactor (MFBR), the slurry phase tions in the catalyst bed. The catalytic portion of the reactor is a
reactor (SPR) and the circulating fluidized-bed reactor. Two operating microchannel slot sandwiched within two separated oil-heating
processes have been developed: the high-temperature FT processes channels. Chambrey et al. (2011) experimentally compared with

Table 1
Comparison between two commercial FT reactors operating the LTFT process (Dry, 2002).

Advantages Drawbacks

MFBR Turbulent flow is ensured by the short distance between catalyst particles and High-pressure operation and pressure drop causes high gas compression costs
especially by a high velocity of syngas
Separation of liquid products from the catalyst phase is easy to process Replacement of the catalyst requires special care
Plug flow conditions are obtained Catalyst sintering can occur due to high operating temperature
When tubes diameter increases, difficult heat removal and no temperature
control can lead to hot-spot formation
SPR Slurry phase is well mixed and tends to isothermal conditions Separation of the wax product from the catalyst is difficult
High productivity per mass of catalyst (lower charge than for fixed bed reactor) Well mixed reactor behavior tends to decrease the reactions rates
centimetric and millimetric tubular fixed bed reactors under exactly approaches to describe FT kinetics arise: (1) kinetic model based
the same operating conditions (CoPt/Al2O3, 220 1C, 20 bar, H2/CO2). on mechanistic proposal consisting in a sequence of elementary
The authors found an increase from 40 to 60% of conversion reactions involving adsorbents and/or intermediates; (2) empirical
indicating that the millimetric fixed bed represented a real gain in expressions of general power-law kinetics; and (3) semi-empirical
term of productivity. The reasons of this improvement are unclear: kinetics based on FT mechanisms. The third approach is generally
absence of internal temperature gradient, role of the hydrody- based on Langmuir–Hinselwood–Hougen–Watson (LHHW) model
namics of both phases as the expected increase of the superficial equations. On the basis of carbide or enolic mechanisms, Sarup
velocity for smaller reactor diameter. Many other works report and and Wojciechowski (1989) defined with regards to LHHW equa-
highlight the interest in producing biofuels in intensified reactors. tions a kinetic model for the FT synthesis rate regarding CO and H2,
Knochen et al. (2010) concluded that a milli-structured fixed-bed consumption which the general form is given by
reactor was an interesting option for small FTS plants since high
kFT P aCO P bH2
volume specific productivity can be obtained at acceptable pressure r FT ¼ " #2 ð4Þ
drops. Moreover, Myrstad et al. (2009) showed the ability of 1 þ i ki P cCO P dHi2
P i

structured reactor to operate FT synthesis with high active catalyst


at severe conditions without thermal runaway and subsequent high A review of kinetics equation for iron-based catalyst is given by
deactivation of the catalyst. Catalyst grafted on stainless steel walls Huff and Satterfield (1984) and Zimmerman and Bukur (1990).
of a microchannel reactor substrate demonstrated high activity Many investigators also worked on kinetics expression for cobalt-
under FT reaction. Activity and selectivity were higher under based catalyst (Anderson et al., 1951; Outi et al., 1981; Sarup and
microchannel conditions than in a classical fixed-bed catalytic Wojciechowski, 1989; Yates and Satterfield, 1991). The latter
reactor (Guillou, 2005). In addition to these experimental studies, simplified the equations of Sarup and Wojciechowski down to only
some works based on numerical simulations (Atwood and Bennett, two parameters: the reaction constant kFT and the adsorption
1979; Bub et al., 1980; Jess et al., 1999; Wang et al., 2003; Wu et al., constant k1:
2010) have concluded that FT synthesis carried out in intensified kFT C CO C H2
systems is enhanced considerably due to mass and heat transfer r FT ¼ 2
ð5Þ
ð1 þ k1 C CO Þ
increase. Actually a few mathematical modeling studies on fixed bed
FT reactors have been reported in literature. An oversimplified 1-D Regarding the catalyst, iron and cobalt on various oxide supports
heterogeneous plug flow model was used to describe a centimetric have been widely used for FT synthesis. A short description of their
reactor as early as 1979 (Atwood and Bennett, 1979). Although respective advantages and drawbacks is given by Khodakov et al.
intraparticle diffusion was considered with the assumption that the (2007). Although cobalt catalysts are more expensive than iron
kinetics was first order in CO, no model of product distribution was ones, cobalt catalysts exhibit a greater resistance to attrition, a
developed. Now, several FT reactors 1-D or 2-D models have been higher productivity and a lower sensitivity to water. Moreover, a
developed based on various assumptions. A comparison between product distribution model has been developed by Philippe et al.
one-dimensional and two-dimensional model suggested that the (2009) at high temperature over cobalt-based catalyst. Only paraf-
2-D model leads to more accurate prediction when thermal runaway fin production according to Eq. (1) is considered. Indeed, for sake of
occurs (Jess and Kern, 2009). Some authors (Jess and Kern, 2009, simplicity in the modeling, olefins (Eq. (2) and oxygenates produc-
2012; Lee and Chung, 2012) reported that hot spots and radial tions are not examined and the formation of carbon dioxide by
temperature gradients occurred using a 4.6-centimeter tube dia- Water Gas Shift reaction (Eq. (3)) is neglected, cause of the weak
meter with particle diameter close to 3 mm. Although numerous activity of cobalt based catalyst for the WGS reaction (Newsome,
parameters were studied in these works such as inlet temperature, 1980). All kinetic parameters and adsorption constant k1 follow
cooling temperature, particle diameter or molar ratio H2/CO, no Arrhenius law and activation energies are taken from Philippe et al.
parametric study concerning the tube diameter has been performed (2009). The monoxide carbon consumption rate is
in the millimetric range. r CO ¼ " r FT ð6Þ
with
2.2. Objectives $ %
"Ea
a ' exp ' C CO ' C H2
The main objective of this work is to study how the diameter RT
r FT ¼$ %2 ð7Þ
affects the reactor behavior in terms of conversion, selectivities,
$ %
" ΔH b
temperature gradients and production of the various hydrocarbon 1 þ b ' exp ' C CO
RT
fractions. The model system is based on double-shell packed-bed
reactor with various diameter. A computational 2-D model coupled From the stoichiometry of Eq. (1), water formation rate becomes
with mass and heat transport and hydrodynamic equations is used r H2 O ¼ " r CO ¼ þ r FT ð8Þ
to predict how the FT reaction occurs. The hot-spot formation and
Production rates of methane r C 1 and ethane r C 2 are written following
pressure drop within the reactor were investigated according to the
two specifics laws with two temperature dependent parameters via
diameter with regards to parameters such as weight hourly space
Arrhenius law:
velocity (WHSV) and shell cooling temperature. Finally, the poten- $ %
tial of milli-structured is evaluated in a short scale-up study. " Ed
r C 1 ¼ d ' exp ' r FT ð9Þ
RT
$ %
3. Kinetics of Fischer–Tropsch synthesis "Ee
r C 2 ¼ e ' exp ' r FT ð10Þ
RT
The major challenge in kinetic modeling is to describe the FT
For higher linear hydrocarbons ðn 4 2Þ each hydrocarbon production
reaction with regards to the complexity of the reactions mechan-
rate is determined using a recursive kinetic model based on
isms and the large number of species involved. Some models are
Anderson–Schultz–Flory theory. A constant chain growth probability
based only on CO and/or H2 conversion and do not consider the
(α ¼0.9) is used. Thus, hydrocarbons production rate was given as
selectivity although those are both greatly influenced by tempera-
ture gradients. According to Lee and Chung (2012), three different rCn ¼ α ' rCn " 1 ð11Þ
Moreover, the stoichiometry of reaction (1) leads to the consump- was chosen where gas and solid were only considered. Waxes
tion rate of hydrogen as follows: formation was not included in the hydrodynamic model.
N 2. A laminar flow occurs and Reynolds number was calculated
X
"r H2 ¼ ð2i þ 1Þ ' r C i ð12Þ from spherical pellets mean diameter.
i 3. The effective diffusion of gaseous reactant through the liquid-
And also filled phase pores of the catalyst was not considered as a
limitation so that intraparticle mass transport particle limita-
N
X tion was neglected. Thus, effectiveness factor was considered to
"r CO ¼ 1 ' r C1 þ 2 ' r C2 þ3 ' r C3 þ⋯ þ N ' r C N ¼ i ' rCi ð13Þ
i¼1
be close to one.
4. The production of oxygenates and WGS reactions were
Therefore the consumption of hydrogen can be expressed: neglected due to their low occurrence when using FT synthesis
N
X cobalt-based catalyst.
"r H2 ¼ 2 ' r CO þ rCi ð14Þ 5. All gas streams were supposed to be ideal, thus the ideal gas
i¼1
law was applied to evaluate the density of gas phase and
Kinetics parameters and activation energy used in the current study concentration of gases species.
are listed in Table 2 and were taken from Sadeqzadeh et al. (2012) at 6. The tubular packed-bed reactor runs at steady state.
steady state. In this study, hydrocarbon chains were assumed to 7. The bed porosity was assumed to be constant throughout the
grow up to not more than 50. packing of the bed.

Physical properties of gas stream were considered as function of


4. Mathematical modeling of packed bed reactor temperature, pressure and composition. The viscosities μg;i , ther-
mal conductivities λg;i , specific heat capacities cp;i and enthalpies
A 2D axisymmetric geometry was built and a fully-coupled of formation of hydrocarbons ΔH f ;i of pure gas compounds were
physics model with concentrated medium mass transport, heat estimated by polynomial expressions given by Yaws (1999). The
transfer in fluids and Brinkman equations for the velocity profile viscosity of gas mixtures μg;m was calculated using Wilke's method
and continuity equation are used to simulate transports phenomena as described by Poling et al. (2000) with the modification of
and formations of hydrocarbons. A mixture of CO and H2 enters the Brokaw (1969). Regarding thermal conductivity of gas mixtures
packed-bed reactor with a mass flow, F, an inlet temperature, T0 and λg;m the Wassiljewa equation (Poling et al., 2000) with Brokaw's
a hydrogen to carbon monoxide molar ratio, θH2 =CO . A double-shell approach (Brokaw, 1969) was used.
ensures the cooling and/or heating of the system where a thermal The physical catalyst properties, 25%Co/0.1% Pt/Al2O3, are ass-
fluid flows all around the external reactor tube (Fig. 1). The ociated to the chosen kinetic model described above in Section 3.
hydrocarbon products are mainly under gas phase but formed The main characteristics taken from Sadeqzadeh et al. (2012) are
liquid waxes go through the reactor down to the outlet. Thus, the summarized in Table 3.
flow through the porous medium, i.e. the catalytic packed bed of
solid particles, is a two-phase flow with an increasing liquid flow as
the hydrocarbons chains grow. Table 3
Some assumptions were made to undertake the modeling of Characteristics of the catalyst CoPt/Al2O3 used by Sadeqzadeh et al. (2012).
the reactor. As far as possible the relevance of these assumptions is
studied in this paper. Particle diameter, dp Density, ρcat Cobalt content Void fraction of bed ϵ
ðμmÞ (kg m " 3) (%) (–)

1. Under the operating conditions, liquid phase flow was neglected 90 2030 25 0.4
(Philippe et al., 2009), so that a pseudo-homogeneous model

Coolant Catalyst
out
pellet

Feeding Radial heat FT


CO & H2 removal products
T0

z=0 z z + ∆z

Coolant
in

Fig. 1. Schematic axisymmetric tubular packed bed FT reactor.

Table 2
List of kinetic parameters used in this study (Sadeqzadeh et al., 2012).

aa Ea ba ΔHb da Ed ea Ee α
"1
ðm6 kgcat mol
"1
s " 1Þ (kJ mol " 1) (m3 mol " 1) (kJ mol " 1) (–) (kJ mol " 1) (–) (kJ mol " 1) (–)

7.17 ( 107 100 44.93 20 3.80 ( 107 81 2.01 ( 103 49 0.9

a
a, b, d and e are the pre-exponential factors in the Arrhenius laws.
4.1. Momentum and continuity equations first-order rate was written for the consumption rate of carbon
monoxide assuming a low conversion of CO:
In order to describe flow in porous media and compute fluid 0
" r CO ¼ ka C H2g ð23Þ
velocity profile, the Brinkman equation (Eq. (15)), derivated form
of Navier–Stokes equation in porous media, was adopted. Con- with
sidering the above mechanism (Eq. (1)), the total conversion of CO
0 C COg
towards n-alkane Cn involves the consumption of 2n moles of gas ka ¼ ka " #2 ð24Þ
compound in the total molar flux. Thus, the FT synthesis leads to a 1 þ kb C COg
decreasing of the volumetric gas flowrate, which also modifies the
density, thus a compressible flow formulation of the continuity The effectiveness factor is also defined as function of Thiele
equation (Eq. (16)) was used with the steady-state approximation modulus ϕ:
coupled with the Brinkman equation: tanh ϕ
' ( η¼ ð25Þ
μg;m ! μg;m ! ! ! ! ϕ
∇ ' " pI þ ∇u " u " βF J u J u ¼ 0 ð15Þ
ϵ K where the Thiele's modulus ϕ was modified so as to take into
"
!
# consideration hydrogen diffusion through the pore, filled with
∇ ' ρm;f u ¼ 0 ð16Þ liquid wax, inside the pellet:
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
Considering a laminar flow regime (Rep o 1 Nield and Bejan, 2013) 0
dp ρcat ka C H2g
with very low velocity, even if laminar flow though packed bed is ϕ¼ ð26Þ
6 Deff ;H2 ;l C H2l
fulfilled for Rep =ð1 " ϵÞ o 10, the Brinkman viscous terms are
predominant compared to the inertial term represented by the The Henry's law gave value of hydrogen concentration in liquid
quadratic velocity term (Chukwudozie et al., 2012; Marpu, 1995). wax:
The porous media permeability K was determined using Carman– P H2 RT
Kozeny: C H2l ¼ ¼ CH ð27Þ
HeH2 HeH2 2g
ϵ3 2 Only a fraction of porous media is permeable to dissolved hydro-
K¼ dp ð17Þ
180ð1 " ϵÞ2 gen and the path of diffusion through the porous particles filled
with liquid hydrocarbons, is random and tortuous (Table 4). Thus,
4.2. Transport species equations to estimate effective diffusion coefficient of dissolved hydrogen
through the catalyst pores, the tortuosity (τp) and the porosity (ϵp)
The local mass balance at steady state for each species i was of particle were taken into account in the estimation:
given by Eq. (18). Axial and radial dispersion, convection and ϵp
Deff ;H2 ;l ¼ D ð28Þ
source terms were considered: τp m;H2 ;l
! "
!
#
∇ ' ji þ ρm;f u ' ∇ wi ¼ ρcat M i ð1 " ϵÞηr i ð18Þ
! 4.3. Heat transport
The mass flux density of specie i, noted ji , caused by axial and
radial dispersion is modelled by a Fick's law analogy: In a 2D axial and radial dispersion plug flow model, heat
! transport is ensured by convection and thermal dispersion. The
ji ¼ " ρm;f De ∇wi ð19Þ
fluid and the solid were assimilated to a homogeneous body – the
where De is a diagonal matrix lumping the effective axial Dea and bed – though heat is transferred by either axial or/and radial
radial Der dispersion coefficient. For low value of the Reynolds thermal dispersion. Radiative transport is neglected in our case.
number ðRep r1Þ which is our case, the two dispersion coefficients Based on the same assumptions for transport species, the heat
are approximately the same (Delgado, 2006) and also equal to the balance was written as
average molecular diffusion coefficient Dm i (Eq. (20)) (Delgado, N
ρm;f cpg;m !
X
2006) calculated from binary diffusion constant Di;j .
+ ,
u ' ∇T ¼ ∇ ' λe ∇T þ ΔHf ;i r i ρcat ð1 " ϵÞ ð29Þ
1 " wi i¼1
Dm
i ¼P xk ð20Þ
Several theoretical and experimental studies have been performed
kai
Di;k on the determination of the effective thermal conductivity λe which is
The diffusion coefficients for binary gas systems Di;j were esti- a matrix lumping the radial thermal conductivity λer and the axial
mated with empirical correlations proposed by Fuller and thermal conductivity λea (Özgümüş et al., 2013). Predictive correlations
Giddings (1965), Fuller et al. (1966, 1969). The local mass balance for the axial or the radial thermal conductivity λei =λg;m (the subscript i
for each compound in the gas phase was rewritten according to refers to a for axial or r for radial thermal conductivity) are mainly
" # ! " # expressed in a linear form with respect to particle based Reynold (Rep)
ρm;f !u ' ∇ wi " ∇ ' ρm;f Dmi ∇wi ¼ ρcat M i ð1 " ϵÞηr i ð21Þ or Peclet number ðPep ¼ Rep PrÞ such as the following equation:

The kinetic model developed above gave the source terms corre- λei λbed
¼ þ K Rep Pr ð30Þ
sponding to the intrinsic rate of formation for each compound. The λg;m λg;m
effectiveness factor was calculated as the ratio between effective
CO consumption and the consumption at the pellet surface.
r COeffective Table 4
η¼ ð22Þ Parameters used for the Thiele modulus evaluation.
r COsurface
HeH2 (Pa m3 mol " 1) Dm;H2 ;l (m2 s " 1) ϵp =τp (–)
The investigation of competition between pore diffusion and
reaction is only based on CO consumption rate due to recursive 20,000 4.0 ( 10 "8
0.3
kinetics of hydrocarbons forming reactions. Therefore a pseudo-
The value for the radial effective heat conductivity λer can be well 4.5. Computational methods
described by the correlation of Bauer and Schlunder (1978a,b):
Commercial CFD software COMSOL Multiphysics version 4.3b
λer λbed h Rep Pr
¼ þ i ð31Þ installed on a workstation with Intel Xeon 8 cores at 2.66 Ghz with
λg;m λg;m 7 2 " +1 " 2dp =dR ,2
16 GB RAM was used to solve the coupled equations for the
modeling of the FT packed bed reactor. A free triangular mesh
with λbed the effective thermal conductivity when no fluid flows,
depending on the tube diameter is used. Given the constant
which is in fact the static contribution of the effective thermal
number of dependent variables, degrees of freedom (DOFs) only
conductivity. Thus, the ratio λbed =λg;m is constant and is usually in
depends of the grid fineness or the number of nodes. Values of
the range 2–10. As suggested by Jess and Kern (2012) the ratio was
DOFs were about 250,000 depending on tube diameter. A grid
taken equal to 4 according to Borkink and Westerterp (1992). Axial
refinement was also performed in the vicinity of the inlet of the
effective heat coefficient λea was calculated by the following correla-
reactor where hot spot potentially occurs to ensure reasonable
tion (Springer, 2010):
accuracy of solutions. Parameters used for the simulation are
λea λbed Rep Pr summarized in Table 5.
¼ þ ð32Þ
λg;m λg;m 2 Transport properties such as dynamic viscosity, thermal conduc-
tivity and specific heat of gases were evaluated as function of the
bulk gas phase composition and temperature by using external
4.4. Boundary conditions MATLAB function. Simulations were carried out in order to examine
the variations of the transport properties. All transport properties
The flow in the reactor is driven by pressure drop, so that the change along the reactor due to changes in composition and
outlet pressure is equal to the constant value P out . No slip condition temperature of gas stream. Nevertheless, heat capacity and viscosity
is applied, thus the velocity uwall in the wall vicinity tends to zero. variations do not exceed 30% of the inlet value whereas thermal
Following boundary conditions were applied at inlet and outlet: conductivity decreases of a factor of 3.4. Thus, other simulations were
! carried out using constant transport properties. Only thermal con-
At r ¼ 0 : u ðr; 0Þ ¼ u0
/!
ð33Þ
ductivity ðλg;m Þ exhibited a significant impact on the thermal
behavior of milli-fixed bed reactor. Indeed simulation, where thermal
Tðr; 0Þ ¼ T 0 ð34Þ
conductivity was kept to be constant (λg;m ¼ 0:134 W m " 1 K " 1 ,
wi ðr; 0Þ ¼ wi;0 ð35Þ
θH2 =CO ¼ 2, T w ¼ 493:15 K), led to a difference of more than 20 K
on the maximal temperature reached within the bed. This maximum
Pðr; LÞ ¼ P out ð36Þ varies from 503.4 K when the thermal conductivity is kept constant
to 524.5 K when it is not. Other simulations have highlighted a
Symmetric boundary conditions are variation of less than 1 K between the cases when heat capacity and
∂wi ∂T viscosity of gases mixture are kept constant or not along the reactor.
8 z; r ¼ 0 : ¼ ¼0 ð37Þ Thus, in this study, the viscosity and the heat capacity of gas mixture
∂r ∂r
were kept constant, calculated at the inlet conditions, whereas the
Wall boundary conditions are change in thermal conductivity was considered in order to achieve a
! ! satisfactory degree of accuracy in the model output.
8 z; r ¼ R : u ðR; zÞ ¼ 0 ð38Þ

∂wi
¼0 ð39Þ 5. Results and discussion
∂r

∂T 5.1. Intraparticle concentration profile


λer ¼ U overall ( ðT " T w Þ ð40Þ
∂r
with It is of interest to investigate the interaction between diffusion and
reaction in a Fischer–Tropsch catalyst pellet. The mathematical model
1 dR 1 e 1
¼ þ þ þ ð41Þ previously introduced was based on effectiveness factor defined as
U overall 8λer αw;int λwall αw;ext
Eq. (22). A strong dependence of the effectiveness factor was found
The term U overall represents the overall thermal transmittance in the
vicinity of the wall. The wall heat transfer coefficient αw;int is Table 5
Data and operating conditions used to model the
calculated from the following correlation of Martin and Nilles
packed fixed bed reactor for Fischer Tropsch
(1993): synthesis.
$ %
α dp dp λbed
Nuw ¼ w;int ¼ 1:3 þ 5 þ0:19 Re3=4
p Pr1=3 ð42Þ Parameters Value
λg;m dR λg;m
P out 20 bar
Heat transfer by pure diffusion through the wall was neglected T0 293.15 K
by considering the thin material thickness and the high thermal WHSV "1
20–600 gsyngas min " 1 kgcat:
conductivity of the stainless steel wall. The external heat transfer θH2 =CO 2 (–)
limitation was also neglected by assuming high velocity of the α 0.9 (–)
cooling fluid in the shell. Thus, boundary condition (Eq. (40)) at ϵ 0.4 (–)
the wall cooling was rewritten as Tw 488.15–503.15 K
μg;m a 2.4 ( 10 " 5 Pa s
∂T λg;H2 a 250 W m " 1 K " 1
8 z; r ¼ R : λer ¼ αw;int ( ðT " T w Þ ð43Þ
∂r λg;CO a 0.0376 W m " 1 K " 1
λg;m a 0:134 W m " 1 K " 1
The global heat released by the Fischer–Tropsch synthesis was
computed from the enthalpy of formation (Yaws, 1999) of each a
Estimated values for the inlet at T ¼ 493.15 K
hydrocarbon depending on bulk temperature. and molar ratio θH2 =CO ¼ 2 for mixture.
1.10 diameters, i.e. 1″, 3=4″, 5=8″, 1=2″, 3=8″, 1=4″, 1=8″ and 1=16″ with
1.00 standard thicknesses. The first results in temperature contours and
0.90 hot spot formation showed that no convergence was achieved for
0.80 tube diameters higher than 1=4″ (6.35 mm). Actually for those
[-]

0.70
diameters, temperature within the catalytic bed increases drastically
with no possible control of the reaction heat release and the thermal
Effectiveness factor

0.60
runaway leads to undefined solutions of the simulation. This was
0.50
observed whatever the chosen WHSV. Actually, this is explained by
0.40
the activity of the catalyst associated to high effectiveness factor (η).
0.30 T = 493.15 K Thus, the results reported here focus on diameters ranging from
0.20 T = 533.15 K 0.88 mm to 3.11 mm as internal diamater (dR). It was decided to
0.10 T = 563.15 K investigate the influence of the diameter without any dilution of the
0.00 catalyst that would have allow the use of large tubes.
"1
1 10 100 1000 10000
Space velocity ðWHSV ¼ 20 gsyngas min " 1 kgcat: Þ, cooling tem-
Particle diameter d p [ m]
perature (T w ¼ 493:15 K and molar ratio hydrogen to carbon mon-
Fig. 2. Variation of effectiveness factor with pellet size for three different oxide ðθH2 =CO ¼ 2Þ were chosen as constant parameters. The two-
temperatures. dimensional temperature contours are depicted for two different
tubes diameters (Fig. 3). Fig. 3 highlights the formation of hot spots
Table 6 as the tube diameter increases. For the lowest given diameter,
Volume and molar fraction of compounds under FT operating conditions, i.e. 0.88 mm (Fig. 3a), isothermal contour arises (in yellow color) and
dR ¼ 1:75 mm, WHSV ¼ 70 gsyngas min
"1 "1
kgcat: . the maximum reached temperature does not exceed 494.3 K. As
the diameter rises, maximum temperature increases up to 621.4 K,
Compounds Phase Production (493.15 K, 20 bar) Volume Molar i.e. temperature changes from yellow to red in color (Fig. 3b), hot
P v ðL h
"1 "1
kgcat: Þ fraction (%) fraction
spot appears and gets closer to inlet of the reactor. In all cases near
(%)
the inlet region, the flow is heated up by the heat transfer fluid.
H2O Gas 401.64 88.95 86.85 Indeed heat transport by diffusion (black arrows) operates from the
H2 Gas 2.08 0.46 0.45 wall, that is a hot region near the inlet, towards the centerline and
CO Gas 20.01 4.43 4.33 the coolest regions. Particularly when the hot spot arises, the heat is
C1–C6 Gas 26.02 5.76 5.63
C7–C50 Liquid 1.77 0.39 2.74
removed by the coolant fluid outside of the packed bed. In the
vicinity of the hottest zone a part of the heat released is transported
by thermal diffusion boosting the temperature increases of the
according to the pellet size (Fig. 2) whatever the operating tempera- reactor bulk flow in the coolest zone. For this reason the diffusive
ture. However it can be seen that, for very small catalyst particle heat flux is substantially opposed to the convective heat flux (red
ðdp ¼ 90–100 μmÞ, the CO-based effectiveness factor is close to one. arrows).
With increasing the particle diameter size and/or boosting the The axial temperature profile is illustrated in Fig. 4 for different
operating temperature, the effectiveness factor decreases to a rela- pipe sizes reported in metric value.
tively low value involving severe limitation of diffusion inside the It can be observed that better heat removal and a flat tem-
pores. Moreover, the results show that, for pellets size of about perature profile within the bed can be obtained with lowest tube
2–4 mm, the diffusion within the catalyst pore cannot be neglected. diameter. Between 0.88 mm and 2.00 mm inner tube diameter, the
Thus, the assumption that the effectiveness factor is equal to one is system removes efficiently the heat released during the synthesis;
verified as far as in the present model the catalyst particle diameter is no thermal runaway is observed within the bed. Only a slight
90 μm. increase up to 6 K over the cooling temperature ðT w Þ is observed
when the diameters reaches 2.00 mm. Above 3.10 mm and as
5.2. Gas–liquid hydrocarbon phase ratio under typical operating mentionned previously, thermal runaway occurs and temperature
conditions increasing is so drastic that no convergence of the simulation is
possible. The thermal behavior reported in Fig. 4 is in agreement
As indicated previously, reactor simulation was chosen to be with the two-dimensional centimeter models developed by Jess
carried out by solving the species transport equations and the Brink- and Kern (2012) and Lee and Chung (2012) in centimeter tubes.
man equations for gaseous compounds only (T w ¼ 493:15 K and They showed how the performance of the reactor and the maximal
P out ¼ 20 bar). Indeed under these operating conditions, thermody- temperature reached are highly sensitive with tube diameters and
namics equilibrium calculations (National Institute of Standards and catalyst particle size.
Technology (NIST), 2013) revealed that hydrocarbons from C1 to C6 The maximal temperature T max within the bulk is correlated to
were in gaseous phase whereas compounds from C7 to C50 where in the tube diameter of the FT reactor (Fig. 5). As shown in Fig. 5, space
"1 "1
liquid phase. Simulation results reveal that the pseudo homogeneous velocities in the range of 20–80 gsyngas min kgcat: seems to have
model is relevant regarding gaseous versus liquid flow ratio (Table 6). no effect on the maximal temperature. As the diameter decreases,
ASF distribution was used to describe hydrocarbons production over the maximal temperature induced by the heat released by the FT
C6 and results highlight that 99% of the volumetric flow representing synthesis decreases towards the setpoint temperature T w . Actually,
more than 96% molar of total compounds are in the gas phase. for the studied diameters, the CO conversion does not change
significantly in the given conditions of WHSV as the residence time
5.3. Influence of tube diameter on temperature profile and selectivity is high enough. So, the total heat released rises with WHSV due to
at constant WHSV the total gas flow increasing. For smaller diameters, overall heat
transfer coefficients are so high that no difference in T max is observed
The effect in changing the tube diameter toward millimetric scale as WHSV increases. At a given WHSV, with increasing tube diameter,
with regard to the behavior of the reactor is investigated. The first the heat removal from the core to the wall cooling decreases. Indeed
purpose was to carry out the study according to eight diameters of when diameter increases, overall heat transfer coefficient gets lower
millimetric to centrimetric size using standard stainless steel tube and the ability of the reactor to remove reaction heat diminishes.
Fig. 3. Temperature contours, diffusive heat flux (black arrow) and convective heat flux (red arrow) within the fixed bed reactor under operating condition T w ¼ 493:15 K,
"1
WHSV ¼ 20 gsyngas min " 1 kgcat: for different tube diameters. (For interpretation of the references to color in this figure caption, the reader is referred to the web version of
this article.)
3500
593.15

overall
583.15
3.12 mm 3000
573.15
3.10 mm
Centerline temperature T [K]

563.15 3.05 mm 2500


553.15 2.75 mm

[W m2 K-1]
2.50 mm 2000
543.15
2.00 mm
533.15
1.75 mm 1500
523.15 0.88 mm
513.15
1000
503.15
493.15 500
483.15
0 0.01 0.02 0.03 0.04 0.05 0
Axial length z [m] 0.00 0.01 0.01 0.02 0.02 0.03 0.03 0.04 0.04 0.05 0.05
Axial length z [m]
Fig. 4. Influence of the tube diameter on the temperature (r ¼0) along axial
"1 "1
direction of a cooled tubular fixed bed reactor ðWHSV ¼ 20 gsyngas min kgcat: ). Fig. 7. Overall heat transfer coefficient along axial length for different tube
"1
diameters ðT w ¼ 493:15 K; WHSV ¼ 20 gsyngas min " 1 kgcat: Þ.
533.15
70 g syngas/[min.kg_cat] 100.0% 14.4%
528.15
60 g syngas/[min.kg_cat]

CO conversion and C5+ selectivity [%]


50 g syngas/[min.kg_cat] 98.0%
523.15 14.2%
40 g syngas/[min.kg_cat]
96.0%
30 g syngas/[min.kg_cat]

C1 - C4 Selectivity [%]
518.15 CO conversion [%] 14.0%
Tmax [K]

20 g syngas/[min.kg_cat] 94.0% C5+ selectivity [%]


513.15 C1-C4 selectivity [%]
92.0% 13.8%
508.15
90.0%
503.15 13.6%
88.0%
498.15 13.4%
86.0%
493.15
0.50 0.70 0.90 1.10 1.30 1.50 1.70 1.90 2.10 2.30 2.50 2.70 2.90 3.10 84.0% 13.2%
0.80 1.00 1.20 1.40 1.60 1.80 2.00 2.20 2.40 2.60 2.80 3.00
Inner diameter of reactor d R [mm]
Inner diameter of reactor d R [mm]
Fig. 5. Maximal temperature versus tube diameter for different space velocity
Fig. 8. Effect of tube diameter on CO conversion with T w ¼ 493:15 K, P ¼20 bar and
ðT w ¼ 493:15 KÞ. "1
WHSV ¼ 20 gsyngas min " 1 kgcat: .

3.51E-03
Heat transfer resistance Uoverall-1 [m2 K 1 W-1]

Overall thermal resistance 0.20


3.01E-03
0.18
Bed thermal resistance
Productivity [kg h -1 kg cat. -1]

0.88 mm
0.16
2.51E-03 Internal wall thermal resistance 1.75 mm
0.14 2.00 mm
2.01E-03 0.12 2.50 mm
0.10 2.75 mm
1.51E-03 3.05 mm
0.08
0.06 Increasing diameter 3.10 mm
1.01E-03
3.11 mm
0.04
5.10E-04 0.02
0.00
1.00E-05
as

le

ax
l
e

se
lin

en
ht

d
LP
lg

W
ie

id

0.80 1.00 1.20 1.40 1.60 1.80 2.00 2.20 2.40 2.60 2.80 3.00 3.20
ap

os
o

D
e

as

M
er
Fu

N
G

Inner diameter of reactor dR [mm]


Fig. 9. Effect of tube diameter on productivity of common hydrocarbon cut (fuel
Fig. 6. Heat transfer resistance versus tube diameter reactor ðT w ¼ 493:15 K; gas, liquid petroleum gas (LPG), gasoline, naphtha, jet fuel, diesel, middle distillate,
"1
WHSV ¼ 20 gsyngas min " 1 kgcat: Þ. "1
wax) with T w ¼ 493:15 K, P ¼20 bar and WHSV ¼ 20 gsyngas min " 1 kgcat: .

Heat transfer resistance of the bed ðdR =8λer Þ rises with the tube external heat transfer resistance of the shell side is negligible due the
diameter whereas heat transfer resistance of the internal reactor high flow velocity of the coolant. Therefore, the overall heat transfer
ð1=αw;int Þ wall remains nearly constant (Fig. 6). Moreover as coefficient can be written as
expected the heat transfer resistance of the wall by conduction is
%"1
negligible and closed to 1 ( 10 " 4 m2 W " 1 K " 1. This ability to
$
1 dR
U overall ¼ þ ð44Þ
remove heat released changes so much that not only hot-spot αw;int 8λer
appears but also as WHSV rises heat removal becomes too low to
avoid temperature runaway. For centimeter tube diameters the calculated overall heat transfer
Moreover it is particularly interesting to estimate the overall heat coefficients usually range from around 200–400 W m " 2 K " 1
transfer coefficient U overall defined with the assumption that the (Philippe et al., 2009; Jess and Kern, 2012; Lee and Chung, 2012).
In this work, the overall heat transfer coefficients calculated are 100% 2.50
much higher (Fig. 7). The coefficient even reaches 3300 W m " 2 K " 1

CO conversion and C 5+ selectivitity [%]


90% 2.25
at the vicinity of the reactor inlet for a 0.88 mm tube diameter. This 80% 2.00

Productivity [kg h-1 kg cat.-1]


study reveals also the significant variation of the overall heat 70% 1.75
transfer along the axial direction z of the reactor. The value of the
60% 1.50
gas mixture thermal conductivity strongly associated to the thermal
50% 1.25
conductivity of hydrogen, increases in the heating zone involving
CO conversion
the rise of the overall heat transfer coefficient. As the synthesis 40%
C5+ selectivity
1.00

starts, the gas mixture conductivity of gas decreases drastically from 30% Productivity of C5+ 0.75
132 to 39 mW m " 1 K " 1 due to the consumption of hydrogen and
Productivity of C1-C2
20% Productivity of C3-C4 0.50
the decrease of gas velocity. 10% 0.25
The effect of tube diameter on the conversion of carbon
0% 0.00
monoxide and the product selectivity is given by Fig. 8. It can be 30 130 230 330 430 530
seen that C1–C4 fraction increases due to the temperature WHSV [g syngas min -1 kg cat.-1]

increases. CO conversion and C5 þ selectivity slightly decreases Fig. 11. CO conversion and hydrocarbon productivity versus space velocity WHSV
with increasing of tube diameter. for a centimeter-scale tube dR ¼ 2:75 mm, T w ¼ 493:15 K.
The productivities of hydrocarbon cuts are computed from the
recursive kinetics and are shown in Fig. 9. The productivities in
liquid transportation fuels (gasoline, Diesel, jet fuel) are found to
be slightly affected by the tube diameter while the quantity in fuel 700
gas increases drastically when the tube diameter is increased. WHSV=600 g_syngas/min/kg_cat.
600 WHSV=520 g_syngas/min/kg_cat.
5.4. Influence of WHSV on product selectivity and productivity WHSV=440 g_syngas/min/kg_cat.
Pressure drop [mbar] 500
at constant tube diameter WHSV=360 g_syngas/min/kg_cat.
WHSV=280 g_syngas/min/kg_cat.
400
WHSV=200 g_syngas/min/kg_cat.
The influence of space velocity on CO conversion, selectivities and
hydrocarbon productivities are presented for two tube diameters: 300
dR ¼ 0:88 mm (Fig. 10) and dR ¼ 2:75 mm (Fig. 11). These figures do
not exhibit sensitive differences except the C5 þ production. Regarding 200

the 0.88 millimiter tube it can be observed that, up to a WHSV of


100
"1
80 gsyngas min " 1 kgcat: CO conversion remains approximately constant
ðX CO ¼ 92%Þ whereas the C5 þ productivity, which is the total mass 0
0 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.1
flow rate of C5 þ hydrocarbons based on catalyst weight, increases
"1 "1 Axial length z [m]
and then remains slightly unchanged ðP mC5 þ ¼ 2:0 kg h kgcat: Þ.
Fig. 12. Pressure drop versus space velocity WHSV for a tube diameter
Then, as soon as the space velocity reaches values higher than
dR ¼ 2:75 mm, T w ¼ 493:15 K.
"1 "1
90 gsyngas min kgcat: , the conversion decreases and the productivity
"1 "1
remains constant ðP mC5 þ ¼ 2:0 kg h kgcat: Þ because the residence
time decreases below the time needed to convert efficiently the influence of WHSV on selectivity in C5 þ . Actually, a slight increase in
syngas into hydrocarbons. As regards the 2.75 mm tube (Fig. 11) temperature around 503 K occurs with 2.75 mm tube and overall heat
selectivity of C5 þ decreases and CO conversion remains both roughly transfer coefficient is not much dependent on WHSV. High efficiency
"1 in heat transfer allows the release of reaction energy and thus prevents
constant up to a space velocity of 100 gsyngas min " 1 kgcat: . Although
the enhancement of light hydrocarbons production.
the C5 þ production is higher than within a 0.88 mm tube, an optimal
"1
value of space velocity, around 180 gsyngas min " 1 kgcat: , is needed to 5.5. Scale-up considerations
ensure the maximal production of C5 þ . Productivity values of C5 þ are
in agreement with Sadeqzadeh et al. (2012) for millimeter-scale tubes Scale-up calculations via numbering-up of tubes were carried
and the production of light hydrocarbons increases with the rise in out on the basis of the requirement for a production expressed in
tube diameters. Moreover, both Figs. 11 and 10 highlight the weak barrels per day (bpd) of C5 þ fraction. Because of the low tem-
perature rise within millimeter-scale geometry, a 2.75 mm tube
100% 2.50
was chosen. Moreover C5 þ selectivity remains higher than 83%
90% 2.25 whatever the space velocity (Fig. 10) due to the high thermal
CO conversion and C5+ selectivity [%]

80% 2.00 performance to dissipate heat released. Pressure drops given by


Productivity [kg h-1 kg cat.-1]

70% 1.75 simulations for different space velocities as shown in Fig. 12 are
60% 1.50 strictly linear and obey to an analogous Darcy's law for porous
media owing to laminar flow regime because Reynolds number
50% 1.25
based on particle diameter (Rep) was in the range 0.04–1.
40% CO conversion 1.00
C5+ selectivity For a 2.75 mm diameter, the scale-up study consist in deter-
30% 0.75
Productivity of C5+ mining the number of tubes needed to produce one barrel
20% Productivity of C1-C2 0.50 of liquid hydrocarbons ðN tubes Þ. According to Table 7, an optimal
Productivity of C3-C4
"1 "3
10% 0.25 production of 2943 kg h mcat: is ranging around about
"1
0% 0.00 180–200 gsyngas min " 1 kgcat: whereas with conventional reactors
30 130 230 330 430 530 (fixed bed or slurry), the catalyst productivity of liquid hydro-
min-1 kg -1] "1 "3
WHSV [g syngas cat. carbons does not exceed 400 kg h mcat: (Oxford Catalyst, 2012).
Fig. 10. CO conversion and hydrocarbon productivity versus space velocity WHSV This range of value is in agreement with the microreactor used by
for a millimeter-scale tube dR ¼ 0:88 mm, T w ¼ 493:15 K. Velocys (Deshmukh et al., 2010).
Table 7 λg;m thermal conductivity of gas mixture (W m " 1 K " 1)
Results for the scale-up calculations for a 2.75 mm tube diameter, T w ¼ 493:15 K. thermal conductivity of the wall (W m " 1 K " 1)
λwall
WHSV us;0 X CO SC 5 þ PC5 þ N tubes
μg;i viscosity of gaseous specie i (Pa s)
"1 "1
ðgsyngas min " 1 kgcat: Þ (mm s ) (%) (%) ðkg h
"1 "3
mcat:
"1
Þ (tubes bpd ) μg;m viscosity of gas mixture (Pa s)
ϕ Thiele's modulus (dimensionless)
20 4.64 97.73 85.80 510.6 17484 ρf density of gas mixture (kg m " 3)
30 6.96 94.54 85.80 741.3 12 042
40 9.28 93.70 85.78 979.9 9111
ρp density of particle catalyst (kg m " 3)
100 23.19 92.08 85.73 2409.2 3706 τp tortuosity (dimensionless)
160 37.11 71.80 85.77 2939.7 3037 θH2 =CO molar ratio H2/CO (dimensionless)
180 41.75 64.65 85.80 2943.3 3033
200 46.38 58.79 85.83 2939.4 3037
240 55.66 49.94 85.86 2926.0 3051 Latin letters
320 74.21 38.82 85.89 2893.7 3085
360 83.49 35.11 85.90 2877.3 3103 !
ji mass diffusional flux for the specie i (kg m " 2 s " 1)
400 92.28 32.14 85.91 2860.8 3121
440 102.05 29.71 85.92 2844.2 3139 !
u velocity profile (m s " 1)
560 129.88 24.51 85.95 2794.8 3194
a pre-exponential factor in Arrhenius' law (m6 kgcat"1
600 139.15 23.23 85.96 2778.6 3213 "1 "1
mol s )
b pre-exponential factor in Arrhenius' law (m3 mol " 1)
Ci molar concentration of specie i (mol m " 3)
d pre-exponential factor in Arrhenius' law
6. Conclusions (dimensionless)
dp pellet diameter (m)
Simulation results based on the described two dimensional
dR inner tube diameter (m)
model have shown a high dependency of the thermal behavior
Dea effective axial dispersion coefficient (m2 s " 1)
with respect to the thermal conductivity of the gas mixture, which
Deff ;H2 ;l effective diffusion coefficient of hydrogen (m2 s " 1)
is linked to the hydrogen consumption, exhibiting a high thermal
Der effective radial dispersion coefficient (m2 s " 1)
conductivity compared with other compounds, along the reactor.
e pre-exponential factor in Arrhenius' law
Although the viscosity and the heat capacity slightly change
(dimensionless)
during synthesis, this variation does not cause the same variation
F inlet mass flow (kg s " 1)
of temperature within the bed. The simulations of different tube
HeH2 Henry's coefficient (Pa m3 mol " 1)
diameters of the fixed-bed reactor allowed to highlight the high
I identity matrix (–)
performance of millimeter-scale for FT synthesis. No exact con-
K porous media permeability (m2)
vergence was achieved for diameter tube higher than 3.11 mm due
ka
0
apparent constant rate of consumption of CO
to the thermal runaway affecting the other variables strongly
(m3 kgcat
"1 "1
s )
coupled. Therefore, in the work reported here using a kinetic
ka constant rate of consumption of CO
associated with a highly active catalyst, the critical diameter "1 "1
involving a hot spot lower than 10 K is found to be lower ðm6 kgcat mol s " 1Þ
2.50 mm. No significant temperature rise occurs below this dia- kb adsorption constant (m3 mol " 1)
kFT "1 "1
meter range preventing the increase of light hydrocarbons selec- kinetic constant of FT synthesis ðm6 kgcat mol s " 1Þ
tivity. Millichannel catalytic reactor provides effective heat L length of reactor (m)
removal for this exothermic synthesis, leading to an isothermal m mass (kg)
temperature profile which induces low methane selectivity and N chain number of hydrocarbons (dimensionless)
production of long chain hydrocarbons. The campaign of simula- N tubes number of tubes needed to produce one barrel of
tions also allowed to find out an optimal space velocity to liquid hydrocarbons ðC5 þ Þ per day (tubes bpd " 1)
maximize the production rate of C5 þ . Results obtained from Pi partial pressure of specie i (Pa)
scale-up study open a promising way to produce more hydro- PC5 þ mass production of (kg h " 1 mcat.
"3
)
carbons with compact module unit where temperature becomes a P out oulet pressure (Pa)
controllable parameter. r radial length (m)
ri rate of consumption of specie i (mol s " 1 kgcat
"1
)
r COeffective global reaction rate of CO consumption i
Nomenclature
(mol s " 1 kgcat
"1
)
Greek letters r COsurface intrinsic reaction rate of CO consumption i
(mol s " 1 kgcat
"1
)
α chain growth probability (dimensionless) Rep Reynolds number (dimensionless)
αw;int wall heat transfer coefficient (W m " 2 K " 1) S selectivity (dimensionless)
βF Forchheimer coefficient (m " 1) T bulk temperature (K)
ΔH f ;i enthalpy of formation of specie i (J mol " 1) T0 inlet temperature (K)
ΔH R "1
global heat of reaction of FT synthesis (J molCO ) Tw cooling temperature(K)
ϵ bed void fraction (dimensionless) uz axial velocity profile (m s " 1)
ϵp catalyst permeability (dimensionless) U overall overall thermal transmittance (W m " 2 K " 1)
η effectiveness factor (dimensionless) us;0 inlet velocity (m s " 1)
λbed bed thermal conductivity (W m " 1 K " 1) uwall velocity near the wall vicinity (m s " 1)
λea effective axial thermal conductivity (W m " 1 K " 1) wi mass fraction of specie i (dimensionless)
"1
λer effective radial thermal conductivity (W m " 1 K " 1) WHSV weight hourly space velocity ðgsyngas min " 1 kgcat: Þ
λg;i thermal conductivity of gaseous specie i (W m " 1 K " 1) X CO CO conversion (dimensionless)
z axial direction (m) Jess, A., Kern, C., 2012. Influence of particle size and single-tube diameter on
cpg;m heat capacity of the specie i (J kg " 1 K " 1) thermal behavior of Fischer–Tropsch reactors. Part I. Particle size variation for
constant tube size and vice versa. Chem. Eng. Technol. 35, 369–378.
Jess, A., Popp, R., Hedden, K., 1999. Fischer–Tropsch-synthesis with nitrogen-rich
syngas: fundamentals and reactor design aspects. Appl. Catal. A: General 186,
321–342.
Khodakov, A.Y., Chu, W., Fongarland, P., 2007. Advances in the development of
novel cobalt Fischer–Tropsch catalysts for synthesis of long-chain hydrocarbons
Acknowledgments and clean fuels. Chem. Rev. 107, 1692–1744.
Knochen, J., Güttel, R., Knobloch, C., Turek, T., 2010. Fischer–Tropsch synthesis in
The authors are grateful to Pr. Fongarland Pascal, University milli-structured fixed-bed reactors: experimental study and scale-up consid-
erations. Chem. Eng. Process.: Process Intensif. 49, 958–964.
Claude-Bernard Lyon 1, IRCELyon, for the helpful discussions Lee, T.S., Chung, J.N., 2012. Mathematical modeling and numerical simulation of a
concerning Fischer–Tropsch synthesis. Fischer–Tropsch packed bed reactor and its thermal management for liquid
hydrocarbon fuel production using biomass syngas. Energy Fuels 26,
1363–1379.
References Marpu, D.R., 1995. Forchheimer and Brinkman extended Darcy flow model on
natural convection in a vertical cylindrical porous annulus. Acta Mech. 109,
41–48.
Anderson, R.B., Friedel, R.A., Storch, H.H., 1951. Fischer–Tropsch reaction mechan- Martin, H., Nilles, M., 1993. Radiale wärmeleitung in durchströmten schüttungs-
ism involving stepwise growth of carbon chain. J. Chem. Phys. 19, 313–319. rohren. Chem. Ing. Tech. 65, 1468–1477.
Atwood, H.E., Bennett, C.O., 1979. Kinetics of the Fischer–Tropsch reaction over Myrstad, R., Eri, S., Pfeifer, P., Rytter, E., Holmen, A., 2009. Fischer–Tropsch synthesis
iron. Ind. Eng. Chem. Process Des. Dev. 18, 163–170. in a microstructured reactor. Catal. Today 147 (Supplement), S301–S304.
Bauer, R., Schlunder, E., 1978a. Effective radial thermal conductivity of packing in National Institute of Standards and Technology (NIST), 2013. Available on
gas flow. Part I. convective transport coefficient. Int. Chem. Eng. 18, 181–188. 〈http://webbook.nist.gov/chemistry/fluid/〉. Consulted September 9, 2013.
Bauer, R., Schlunder, E., 1978b. Effective radial thermal conductivity of packing in Newsome, D.S., 1980. The water-gas shift reaction. Catal. Rev. 21, 275–318.
gas flow. Part II. Thermal conductivity of the packing fraction without gas flow. Nield, D.A., Bejan, A., 2013. Convection in Porous Media. Springer, New York.
Int. Chem. Eng 18, 181–188. Outi, A., Rautavuoma, I., van der Baan, H.S., 1981. Kinetics and mechanism of the
Borkink, J.G.H., Westerterp, K.R., 1992. Influence of tube and particle diameter on Fischer Tropsch hydrocarbon synthesis on a cobalt on alumina catalyst. Appl.
heat transport in packed beds. AIChE J 38, 703–715. Catal. 1, 247–272.
Brokaw, R.S., 1969. Predicting transport properties of dilute gases. Ind. Eng. Chem. Oxford Catalyst, 2012. Available on 〈http://www.velocys.com〉, Consulted December
Process Des. Dev. 8, 240–253. 16, 2013.
Bub, G., Baerns, M., Büssemeier, B., Frohning, C., 1980. Prediction of the perfor- Özgümüş, T., Mobedi, M., Özkol, U., Nakayama, A., 2013. Thermal dispersion in
mance of catalytic fixed bed reactors for Fischer–Tropsch synthesis. Chem. Eng. porous media—a review on the experimental studies for packed beds. Appl.
Sci. 35, 348–355. Mech. Rev. 65, 031001.
Cao, C., Hu, J., Li, S., Wilcox, W., Wang, Y., 2009. Intensified Fischer–Tropsch Philippe, R., Lacroix, M., Dreibine, L., Pham-Huu, C., Edouard, D., Savin, S., Luck, F.,
synthesis process with microchannel catalytic reactors. Catal. Today 140, Schweich, D., 2009. Effect of structure and thermal properties of a Fischer–
149–156. Tropsch catalyst in a fixed bed. Catal. Today 147 (Supplement), S305–S312.
Chambrey, S., Fongarland, P., Karaca, H., Piché, S., Griboval-Constant, A., Schweich, D., Poling, B., Prausnitz, J., Connell, J.O., 2000. The Properties of Gases and Liquids.
Luck, F., Savin, S., Khodakov, A.Y., 2011. Fischer–Tropsch synthesis in milli-fixed McGraw Hill, New York.
bed reactor: comparison with centimetric fixed bed and slurry stirred tank Sadeqzadeh, M., Hong, J., Fongarland, P., Curulla-Ferré, D., Luck, F., Bousquet, J.,
reactors. Catal. Today 171, 201–206. Schweich, D., Khodakov, A.Y., 2012. Mechanistic modeling of cobalt based
Chukwudozie, C.P., Tyagi, M., Sears, S.O., White, C.D., 2012. Prediction of non-Darcy catalyst sintering in a fixed bed reactor under different conditions of Fischer–
coefficients for inertial flows through the castlegate sandstone using image- Tropsch synthesis. Ind. Eng. Chem. Res. 51, 11955–11964.
based modeling. Transp. Porous Med. 95, 563–580. Sarup, B., Wojciechowski, B.W., 1989. Studies of the Fischer–Tropsch synthesis on a
Delgado, J.M.P.Q., 2006. A critical review of dispersion in packed beds. Heat Mass cobalt catalyst II. Kinetics of carbon monoxide conversion to methane and to
Transfer 42, 279–310. higher hydrocarbons. Can. J. Chem. Eng. 67, 62–74.
Deshmukh, S.R., Tonkovich, A.L.Y., Jarosch, K.T., Schrader, L., Fitzgerald, S.P., Springer (Ed.), 2010. VDI Heat Atlas. VDI-Gesellschaft Verfahrenstechnik und
Kilanowski, D.R., Lerou, J.J., Mazanec, T.J., 2010. Scale-up of microchannel Chemieingenieurwesen, second edition. VDI Gesellschaft.
reactors for Fischer–Tropsch synthesis. Ind. Eng. Chem. Res. 49, 10883–10888. Wang, Y.N., Xu, Y.Y., Li, Y.W., Zhao, Y.L., Zhang, B.J., 2003. Heterogeneous modeling
Dry, M.E., 2002. The Fischer–Tropsch process: 1950–2000. Catal. Today 71, for fixed-bed Fischer–Tropsch synthesis: reactor model and its applications.
227–241. Chem. Eng. Sci. 58, 867–875.
Fuller, E.N., Ensley, K., Giddings, J.C., 1969. J. Phys. Chem. 73. Wu, J., Zhang, H., Ying, W., Fang, D., 2010. Simulation and analysis of a tubular fixed-
Fuller, E.N., Giddings, J.C., 1965. J. Gas Chromatogr. 3. bed Fischer–Tropsch synthesis reactor with co-based catalyst. Chem. Eng.
Fuller, E.N., Schettler, P.D., Giddings, J.C., 1966. Ind. Eng. Chem. 58. Technol. 33, 1083–1092.
Guillou, L., 2005. Synthèse de Fischer–Tropsch en réacteurs structurés àcatalyse Yates, I.C., Satterfield, C.N., 1991. Intrinsic kinetics of the Fischer–Tropsch synthesis
supportée en paroi (Ph.D. thesis), Ecole Centrale Lille. on a cobalt catalyst. Energy Fuels 5, 168–173.
Huff, G.A., Satterfield, C.N., 1984. Intrinsic kinetics of the Fischer–Tropsch synthesis Yaws, C., 1999. Chemical Properties Handbook: Physical, Thermodynamics, Envir-
on a reduced fused-magnetite catalyst. Ind. Eng. Chem. Process Des. Dev. 23, onmental Transport, Safety & Health Related Properties for Organic and
696–705. Inorganic Chemicals. McGraw-Hill Education, New York.
Jess, A., Kern, C., 2009. Modeling of multi-tubular reactors for Fischer–Tropsch Zimmerman, W.H., Bukur, D.B., 1990. Reaction kinetics over iron catalysts used for
synthesis. Chem. Eng. Technol. 32, 1164–1175. the Fischer–Tropsch synthesis. Can. J. Chem. Eng. 68, 292–301.

Potrebbero piacerti anche