Sunteți pe pagina 1din 39

REACTOR MODELING AND SIMULATIONS IN

SYNTHESIS GAS PRODUCTION

Ann M. De Groote and Gilbert F. Froment

Laboratorium voor Petrochemische Techniek


Universiteit Gent, Krijgslaan 281
B9000 Ghent, Belgium

ABSTRACT

Several processes for synthesis gas production are discussed: steam- and
CC^-reforming, partial oxidation and autothermal reforming.
Detailed kinetics are required for the successful simulation of these
processes. Kinetic equations
ff
for steam- and CCVreforming were developed
based upon elementary steps between adsorbed species. Since the reactions
are very fest, severe diffusion limitations occur. These were experimentally
determined. The thresholds for carbon formation according to methane
cracking and the Boudouard reaction were determined in an electrobalance
reactor. Tubular reactors for steam- and CC^-reforming were modeled based
upon a one-dimensional heterogeneous model accounting for internal
diffusion limitations. In these simulations the heat fluxes along the reactor
were calculated from a detailed temperature distribution in the furnace by
means of the program FURNACE.
Tubular reactors for steam- and CCVrefonning were also simulated
based upon a two-dimensional model accounting for both radial and axial
concentration gradients, leading to more accurate bounds for carbon
deposition.
145
Brought to you by | provisional account
Unauthenticated
Download Date | 5/1/18 11:12 PM
Vol. 11, No. 2,1995 Reactor Modeling and Simulations
in Synthesis Gas Production
Furthermore, the catalytic partial oxidation of CH^Oj-mixtures in an
adiabatic fixed bed reactor was simulated based upon the kinetics for total
combustion, steam reforming and the water gas shift reaction.
Finally the catalytic partial oxidation of methane to synthesis gas in an
adiabatic reversed flow reactor was modeled. This is a transient operation
and the traveling waves of temperature and concentrations are simulated for
typical operating conditions.

INTRODUCTION

The conversion of natural gas by steam, oxygen/air or by a mixture of


both, leads to a gas containing H2, CO, COa, H2O, N? when air is used and
eventually unreacted Cut. The components of interest in this gas are CO
and H2, and eventually N2. The synthesis gas can then be used for the
production of ammonia, methanol, phosgene, polycarbonates, formic acid,
acetic acid, oxo-alcohols and intermediates for polyurethane foam plastics.
Eighty per cent of the production cost of these products is determined by the
manufacturing cost of the synthesis gas (Michel, 1989). For the production
of ammonia a synthesis gas with molar ratio of H2 to N2 of 3 is needed. If
the syngas is to be used for methanol production or Fischer-Tropsch
synthesis, a H2/CO-ratio of 2 is required. For oxo-synthesis the H2/CO-ratio
should be equal to 1.
The main commercial process for the production of synthesis gas is the
catalytic steam reforming of natural gas. The catalyst is Ni on AljGj so that
the natural gas has to be hydrodesulfurised. It is fed, together with steam
and eventually recycled carbon dioxide, at typical inlet temperatures between
450 and 650°C to a multi-tubular fixed bed reactor, inserted in a gas fired
furnace. The exit temperature of the process gas ranges from 700 to 950°C,
the pressure between IS and 30 bar (Rastrup-Nielsen, 1993). These
conditions are limited by the tube metallurgy. The reactor tubes have a
length of 10-12 m and an internal diameter of the order of 10 cm. A furnace
contains between 40 and 400 tubes in parallel rows. The tubes are heated by

146
Brought to you by | provisional account
Unauthenticated
Download Date | 5/1/18 11:12 PM
A.M. De Groote and G.F. Froment Reviews in Chemical Engineering

either radiant wall burners, located in the side walls, or long flame burners,
generally located in the ceiling or the bottom. In some reformers long flame
burners are placed in terraces (Foster-Wheeler type) (Froment and Bischoff,
1990).
In steam reforming the following reversible reactions take place: the
endothermic steam reforming reactions

Cff4 + H20 ·* CO + 3 H2 (1)


Cff4 ι- 2 ff2O - CO2 + 4, K2 (2)

and the exothermic water gas shift (WGS)

CO + H20 - CO2 + H2 (3).

Care has to be taken to avoid carbon formation due to methane cracking

cat ~ c * 2 a. (4)

Boudouard reaction
2 CO ~ C + C02 (5)

and CO-reduction

CO + H2 - C+ tf20 (6)

Another process for syngas production is the partial oxidation of natural gas
with oxygen or air. This process combines the exothermic combustion of a
fraction of the natural gas feed

Cff4 + 2 02 »» CO2 + 2 1^0 (7)

with the endothermic steam reforming reactions (1) and (2), the
endothermic CCVrefonning reaction

147
Brought to you by | provisional account
Unauthenticated
Download Date | 5/1/18 11:12 PM
Vol. 11, No. 2, 1995 Reactor Modeling and Simulations
in Synthesis Gas Production
Cff4 + CO2 -> 2 CO + 2 HZ (8)

and the accompanying water gas shift (3) to achieve directly a syngas with
the desired composition. Part of the carbon formed by reactions (4)-(6) can
be burned off by oxygen according to the reaction:

C + O, - CO, (9).

In this process the reactor is refractory lined, so that adiabatic operation


at high pressures and at much higher temperatures than in the steam
reforming reactor is possible. Feeding only 02 and CHLt leads to a syngas
with a low Hj/CO-product ratio, but soot formation is difficult to avoid. Air
can be used instead of oxygen, but this leads to a synthesis gas containing
NZ, which is not easily separated from CO and H2. Air is required when the
synthesis gas is intended for ammonia synthesis. In catalytic partial
oxidation processes the gas is fed at a much lower temperature. With O2
only a CO- and CO2-rich gas for the production of methanol can be
obtained.
When steam is added to the natural gas, together with oxygen or air, the
process is called autothermal (or oxygen) reforming. The total combustion is
either purely homogeneous or catalytic. In case of homogeneous combustion,
a carefully designed burner, achieving efficient mixing of the feed gases, is
required to avoid soot formation. The combustion chamber is followed by a
refractory layer on top of the packed bed of steam reforming catalyst The
combustion gases enter this bed at temperatures around 1200 to 1300 °C.
The exit composition corresponds closely to equilibrium for the steam
reforming and water gas shift reactions.
In the present paper the following processes are modeled:

1. Steam reforming
2. Steam/CO2-reforming
3. Catalytic partial oxidation

148
Brought to you by | provisional account
Unauthenticated
Download Date | 5/1/18 11:12 PM
AM. De Groote and G.F. Frament Reviews in Chemical Engineering

1. Steam reforming

a. Kinetics

Xu and Froment (1989a) studied the steam reforming of methane on a


Topsoe catalyst, containing 15.2% Ni supported on magnesium spinel, in a
tubular reactor. The process can be described by the following triangular
reaction scheme:

(III)

CO:
+H20 -H2
(ID
Reactions (I) and (III) are the endothermic steam reforming reactions,
while reaction (II) is the exothermic water gas shift. By means of model
discrimination they derived a detailed reaction mechanism shown in Figure
1 and consisting of the following reaction steps:
H2O + 1 *- 0-1 + H,
Cfft * 1 - CHt-l
CHt-l «· I - CH^-1 ->· K-l
CH3-1 f I - CH2-1 + H-l
CH2-1 ·fr O-2 -H. CH20-1 + 1
CH2O-1 ·^ 1 ~ CHO-1 + H-I
CHO-1 ·t· 1 ~ CO-1 + H-I SDS (I)
CO-1 πκ 0-2 »· CO2-1 + 1 RDS (II)
CHO-1 Hκ Ο-2 »» C02-l + H-l RDS (III)
CO-1 -* CO + 1
C02-l C02 + 1
-
««
2 H-l H2-l + 1
H2-l τ*
«2 + 1

149
Brought to you by | provisional account
Unauthenticated
Download Date | 5/1/18 11:12 PM
Vol. 11. No. 2.1995 Reactor Modeling and Simulations
in Synthesis Gas Production

HJO}+ I = 0-1 + H.

H2- I -*· 2 H - 1
CH4)+ I
T
CH 4 -I
I

CH 3 -I

+ 0-1 l _ ± O l L _ C H 2 - l + - H 2

CH 2 O-I + I
I
CHO-I + H - I +1

+ H-I CO-I + - H 2 CO-I + H - l


o-i
+ 0-1 T
C02-1 —Ϊ— CO + I CO^-I 4- I

COj| + I
Fig. 1: Detailed reaction mechanism of steam reforming according to Xu
and Froment (1989a).

The corresponding rate equations are:

For CO-production by steam reforming [reaction (I)]:

r, -
(DEN)2

150
Brought to you by | provisional account
Unauthenticated
Download Date | 5/1/18 11:12 PM
AM. De Groote and G.F. Froment Reviews in Chemical Engineering

For the water gas shift reaction (Π):

and for (XVproduction by steam reforming [reaction (ΠΓ)]:

with

JDEW - 1 + J P + JC + P +
(. ιιβ5ο, 13-076)
- 10 r
(J^a-1.784)
Γ
- ίο

The term pn2x (x=2.5, 1, 3.5) in the denominator of the above equations
results from the equilibrium between hydrogen in the gas phase and
diatomically adsorbed hydrogen. Because of this the rate equations are only
valid when there is some H2 in the feed.
The temperature dependence of the rate and adsorption parameters is
given in Table 1. The adsorption constants satisfy the thermodynamic
constraints proposed by Boudart (1967):

• AS°jca < 0 or exp(AS0x/R) = Ax < 1


< S°x.e for non-dissociative adsorption
.0014 AHXa -12.2
or In Αχ = -AS^/R £ (12.2 - 0.0014 AHxJ/R
• usually |Δ8χ8| > 42 J/mol for non-dissociative adsorption.

Carbon formation through reactions (4), (5) and (6) can be a problem in

151
Brought to you by | provisional account
Unauthenticated
Download Date | 5/1/18 11:12 PM
Vol. 11. No. 2.1995 Reactor Modeling and Simulations
in Synthesis Gas Production
Table 1
Temperature dependence of the rate and adsorption constants for the steam
reforming reactions and the water gas shift on a Ni-cataIyst(Xu and
Froment, 1989a)

Rate constants, kj (kmol/kg cat/h)


kj = Aj exp(-E aj /RT) (j=l,2,3)

AJ (dim k.j) E aj (kJ/mol)

*i 4.255 1015 240.1

k2 1.955 106 67.13

*3 1.020 1015 243.9

Adsorption constants, Kx
KX=AX exp(-AHx/RT) (X=CO,H 2 ,CH 4 ,H 2 O)

Ax (dim K x ) ΔΗΧ (kJ/mol)

KCO 8.23 10~5 -70.65

*H2 6.12 10"9 -82.90

K
CH4 6.65 10~4 -38.28

K
H20 1.77 10s 88.68

steam reforming. Methane cracking (4) is endothermic and leads to an


increase in the total number of moles. Therefore, it is favoured by high
temperatures and low pressures. Methane cracking will most likely occur
near the reformer inlet, where the concentration of methane is high.
Reactions (5) and (6) are exothermic and lead to a decrease in the total
number of moles in the gas phase, and are thus favoured by low
temperatures and high pressures. It is most unlikely for these reactions to
occur, unless the temperature is sufficiently low and a significant amount of
carbon monoxide is produced near the reactor inlet.

152
Brought to you by | provisional account
Unauthenticated
Download Date | 5/1/18 11:12 PM
AM. De Groote and G.F. Froment Reviews in Chemical Engineering

Wagner and Froment (1992) studied coke formation and gasification


under steam reforming conditions in an electrobalance reactor. From their
experiments they derived threshold constants for methane cracking

for the Boudouard reaction

^-y^-e'^- 1 '· 081 [*--!

and for CO-reduction

These threshold constants led to the introduction of the following


definition of the coking limit for the methane cracking, the Boudouard
reaction and the CO-reduction:

The definition of Vj is such that coking will occur when its value is lower
than 1. This is an experimental limit, not a thermodynamic one. The
problem arising in the determination of a thermodynamic limit for carbon
deposition is the nature of the carbon. The deposited carbon is neither
graphitic nor carbidic. The ultimate solution is a set of kinetic equations for
the coking reactions. These were also determined by Wagner and Froment
(1992) and by Snoeck and Froment (1994).

b. Diffusion limitations

The intrinsic rates of the steam reforming reactions are very fast, so that
strong diffusion limitations occur inside the catalyst pellet under industrial

153
Brought to you by | provisional account
Unauthenticated
Download Date | 5/1/18 11:12 PM
Vol. 11. No. 2.1995 Reactor Modeling and Simulations
in Synthesis Gas Production

operating conditions. The effective diffiisivity for a component A in reaction


i is related to the molecular diffiisivity, the Knudsen diffusivity and the
tortuosity factor τ by the following equation:

Xu and Froment (1989b) determined the value of the tortuosity factor


under reaction conditions for the above mentioned Ni/MgOA^Os-Topsee
catalyst. They found a value of 3.54. The diffusivities can be calculated from
reliable theoretical expressions.

c. Simulation of an industrial steam reformer


c.l. Reactor model: The one-dimensional heterogeneous reactor model,
accounting for intraparticle diffusion limitations, is used (Froment and
Bischoff, 1990). The model equations are as follows:
continuity equations for methane and carbon dioxide in the gas phase:

energy equation:

dz

pressure drop equation:

dz gdp

with boundary conditions at z=0:

154
Brought to you by | provisional account
Unauthenticated
Download Date | 5/1/18 11:12 PM
AM. De Groote and G.F. Frommt Reviews in Chemical Engineering

The set of differential equations is integrated by means of a 4th order


Runge-Kutta routine.
The continuity equations for the components Cftf and CO2 inside the
catalyst pellets are:

with the boundary conditions

Ρ
'ΐ°^ ^3t - 0 at ξ - 0
ας ας

and
P,.art ~ PCH; at ξ - ι
Ρ,.οα, - Pco, at ξ - l

A planar geometry is used for the catalyst particles, since partial pressure
gradients only occur in a very thin layer near the pellet surface. The
continuity equations for the components inside the pellet are integrated by
means of orthogonal spline collocation. The effectiveness factors for
reactions (1), (2) and(3) can be calculated from:

J>P.)dv
Mi

c.2. Firebox heat transfer: Since the steam reformer tubes are placed in
a gas fired furnace (firebox), the reactor simulation has to be coupled with
the simulation of the firebox in order to account for the temperature
distribution in the furnace.
Rao et al. (1988) developed a zone model, based on the zone method of
Hottel and Sarofim (1967), for the simulation of the firebox. The firebox
heat transfer model is described in detail elsewhere (Rao et al, 1988;
Plehiers and Froment, 1989) and will only be briefly outlined here.

155
Brought to you by | provisional account
Unauthenticated
Download Date | 5/1/18 11:12 PM
Vol. 11, No. 2. 1995 Reactor Modeling and Simulations
in Synthesis Gas Production
The firebox is divided in isothermal surface and volume zones with
uniform properties, in which heat transfer has to be simulated. The energy
balance for a zone i is then given by:

Σ zizi "j - ξ (10)

with ZiZj the total exchange area between zone i and zone j and with Wi the
total emissive power of the hemispherical black body of zone i, given by the
Stefan-Boltzman equation:
IkT Μ /τ *7*
L
"i °• i

The term QiA; in the above energy equation represents the heat leaving zone
i by other means than radiation.
The calculation of the total exchange areas ZjZj consists of three steps.
The first step is the calculation of the view factors between two zones in a
transparent medium, F°ij:

Since analytical integration is impossible, a Monte Carlo routine is


applied to generate the view factors. Plehiers and Froment (1989) found that
a sample of 25000 beams per zone led to an accurate simulation of the heat
transfer in a zone, but that a sample of 250000 beams was required to
predict the heat transfer between two volume zones. The view factor is then
given by:

_o _ number of incidences on zone j


ij
total number of beams emitted from zone i

In the presence of flue gas, absorption and emission of radiation have to


be accounted for. The view factors in a real absorbing medium, Fij, are
derived from those in a transparent medium, F°ij, in the second step. Finally,
in the third step, reflection of radiation is introduced.

156
Brought to you by | provisional account
Unauthenticated
Download Date | 5/1/18 11:12 PM
AM. De Groote and G.F. Frament Reviews in Chemical Engineering

The solution of equation (10) leads to the temperatures in the various


surface and volume zones of the furnace from which the heat flux along the
tubes can be calculated.

c.3. Simulation results: Xu and Froment (1989b) performed a coupled


simulation of the firebox and the reactor of an industrial steam reformer
operating under the conditions given in Table 2. It is assumed that Cyig and
Cs-Cj, present in the natural gas, are immediately converted to Cftj. The
temperature gradient over the external film of the catalyst was found to be
negligible, as already shown before by De Deken et al. (1982).
The simulation results are given in Figures 2 to 4. Figure 2 shows the
tube skin temperatures, the process gas temperature, the total pressure, the
conversion of methane and the conversion of methane to CQj as a function
of the axial position in an industrial steam reformer. The external tube skin
temperature gradually increases from 1000 K to about 1 180 K.
The affinities for equilibrium in the steam reforming reactions and the
water gas shift are given in Figure 3. The affinity for equilibrium in a
reaction i is calculated as:

It is clear from Figure 3 that the water gas shift reaction is already very
close to equilibrium after 3 m in the reformer tubes, while the steam
reforming reactions only reach an affinity for equilibrium of about 80% at
the end of the heated part of the reformer tubes. The difference ΔΤ between
the actual process gas temperature and the value corresponding to
equilibrium, often used in characterising the performance of the steam
reformer, is also shown in Figure 3 and decreases fast to reach a value of 10
K at the end of the heated zone.
The evolution of the effectiveness factors for the steam reforming
reactions through the reactor is presented in Figure 4. A discontinuity occurs
in the profile of the effectiveness factor for the water gas shift (reaction II) at
an axial position of 3.4 m, caused by the reversal of the reaction at the

157
Brought to you by | provisional account
Unauthenticated
Download Date | 5/1/18 11:12 PM
Vol. U. No. 2.1995 Reactor Modeling and Simulations
in Synthesis Gas Production
Table 2
Typical operating conditions of an industrial steam reformer

Feed conditions

Inlet temperature T° ("C) 520


Inlet pressure p° (bar) 29
Inlet flow rate of natural 135
gas per tube F°HG (Nm3/hr)
Inlet flow rate of steam 399
per tube F0^ (Nm3/hr)

Katural gas composition

Mol% CH4 81.5

Mol% N2 14.1

Mol% C2H6 2.8

Mol% C3-C5 0.7

Mol% CO2 1.0

Reactor dimensions

Inner tube diameter dx (m) 0.1016

Outer tube diameter d0 (m) 0.1322

Length of tube L (m) 12

Heated length of tube Lh (m) 11.12

catalyst surface, while there is no such tendency yet inside the catalyst pellet
Further downstream in the reactor, at a position of 9 m, the value of the
effectiveness actor becomes positive again, because now the overall effect of
the reaction inside the catalyst pellet has reversed too. The effectiveness
factors for reactions (I) and (IK) decrease through the reactor, although the

158
Brought to you by | provisional account
Unauthenticated
Download Date | 5/1/18 11:12 PM
AM. De Groote and G.F. Frommt Reviews in Chemical Engineering

p (bar) T (K)
30 j 1200

29 1100

28 1000

27 η 900

26 - 800 0.2

25 700
4 β 8 10 12
Axial position z (m)
Fig. 2: Profiles of methane conversion (XOHM), conversion of methane to
CO2 (XCCM), gas temperature (Tg), internal (TO and external (Tw)
tube skin temperatures and total pressure (pO in an industrial
steam reformer. (Xu and Froment (1989b)).

ATflQ
1.2 80

1-
60
ΟΛ

0.6 H -40
1: CH4 + H2O - CO + 3 H2
2: CO + H2O - CO2 + H2
0.4 3: CH4 + 2 H2O - CO2 + 4 H2
20
0.2 -

4 6 8 10 12
Axial position z (m)
Fig. 3: Approach to equilibrium in the steam reforming reactions and the
water gas shift in an industrial steam reformer. (Xu and Froment
(1989b)).

159
Brought to you by | provisional account
Unauthenticated
Download Date | 5/1/18 11:12 PM
Vol. 11. No. 2.1995 Reactor Modeling and Simulations
in Synthesis Gas Production

0.1

2: CO + H2O - CO2 + H2 )
3: CH4 + 2 H2O - C02 -r 4 H2 !

0.05 -

ui
-OJ05 -

•0.1
4 β β 10 12
Axial position In the reformer tube z (m)

Fig. 4: Effectiveness factors of the steam reforming reactions and water


gas shift reaction along the reactor. (Xu and Froment (1989b)).

rate of reaction (ΠΙ) decreases. The values of TU and η3 are close to those
obtained for a different catalyst by De Deken et al. (1982).
The question often arises whether or not radial temperature and
concentration gradients should be included in the reactor model for an
accurate simulation of the reformer tubes. De Deken et al. (1982) used the 2-
dimensional heterogeneous model of Froment and Bischoff (1990). From
their simulations it appears that an accurate calculation of temperature and
conversion profiles is obtained with the 1-dim model already and that a two-
dimensional reactor model is only required for predicting the zone in which
coke deposition is possible.

2. Steam/CO2-reforming

For certain applications it is necessary to adjust the H2/CO-ratio of the


synthesis gas. This can be done by adding CQz to the feed, but this enhances

160
Brought to you by | provisional account
Unauthenticated
Download Date | 5/1/18 11:12 PM
AM. De Groote and G.F. Froment Reviews in Chemical Engineering

the rate of coke formation. This is illustrated by simulations of steam/CO2-


refonning performed by Snoeck and Froment (1994) for feeds with
COi/CtLj-ratio equal to respectively 0 and 1, for a total pressure of 20 bar
and an inlet temperature of 500°C. The partial pressure of methane was kept
constant The coking limits (as defined earlier) for the methane cracking and
the Boudouard reaction through the reactor are given in Figures 5 and 6. It
is seen that when part of the steam is replaced by CO2? carbon is formed
over the whole length of the reactor by methane cracking. The Boudouard
reaction does not form carbon and neither does CO-reduction. This can be
explained by Figures 7 and 8, indicating that when the amount of steam in
the feed is lowered, the methane conversion is decreased (higher partial
pressure of Cft») and so is the partial pressure of hydrogen. This lowers the
coking limit to values below 1, so that methane cracking is most likely to
occur.
Wagner and Froment (1992) determined the minimum steam to methane
ratio for an industrial steam/CO2-reformer with a feed with CCVCHt-ratio
equal to 1.2. With the thermodynamic approach, based upon the threshold
constants, a minimum H2O/C!l4-ratio of 1.6 is required to avoid coke

Coking limit
5

H20 / CH4 = 2
H20 / CH4 = CO, / CH4 - 0
C02 / CH. = 1

4 β
z(m)
Fig. 5: Coking limit through the catalyst bed for the methane cracking.
(pt = 20bar).
161
Brought to you by | provisional account
Unauthenticated
Download Date | 5/1/18 11:12 PM
Vol. 11. No. 2.1995 Reactor Modeling and Simulations
in Synthesis Gas Production

Coking limit

2 4 β 8 10
z(m)
Fig. 6: Coking limit through the catalyst bed for the Boudouard reaction.
(pt = 20bar).

Partial pressure of methane (bar)


β
Pt = 20 bar
7-

β H 2 0/CH 4 -1
C02/CH4 = 1
4

2- C02/CH4

0 2 4 β 8 10
z(m)
Fig. 7: Partial pressure profiles of methane in the reactor, (pt = 20 bar).

162
Brought to you by | provisional account
Unauthenticated
Download Date | 5/1/18 11:12 PM
AM. De Groote and G.F. Froment Reviews in Chemical Engineering

Partial pressure of hydrogen (bar)


12

10-

4 a 10
2(m)
Fig. 8: Partial pressure profiles of hydrogen in the reactor, (pt = 20 bar).

formation. With the kinetic approach, a steam to methane ratio of 1.2 seems
sufficient.

3. Catalytic partial oxidation and autothermal reforming,

a. Kinetics

A combination of kinetic equations for the total combustion to CÜ2


H2O (7), the steam reforming reactions [(1) and (2)], the water gas shift (3)
and the coking and gasification reactions [(4)-(6) and (9)] is used in the
modeling of the process. No complete set of rate equations obtained on one
and the same catalyst has been published so far.
For the complete combustion of methane to CO2 and H2O, the kinetic
equation developed by Trimm and Lam (1980) for a Pt/Al2C>3-catalyst is
applied in the present paper:

t [OfJ [02] [Of4]


[02]

163
Brought to you by | provisional account
Unauthenticated
Download Date | 5/1/18 11:12 PM
Vol. 11, No. 2. 1995 Reactor Modeling and Simulations
in Synthesis Gas Production
The above equation is combined with the kinetics derived by Xu and
Froment (1989a) for steam reforming and with those determined by Wagner
and Froment (Ί992) and by Snoeck and Froment (1994) for carbon
formation and gasification. A rate equation for the CO2-refonning is not
explicitly included in the set of kinetic equations since with the reaction
scheme, given in the section of steam reforming l.a.) kinetics, this reaction
is a linear combination of CQrproduction by steam reforming (II) and of
the WGS-reaction (ΠΓ).
Another problem is the degree of reduction of the Ni-catalyst required for
the steam reforming reactions and the water gas shift. In other words: does
steam reforming start immediately (parallel to total combustion) or is the
activity for steam reforming gradually increased? If steam reforming is
partially or completely consecutive to the total combustion of methane, the
fraction of the catalyst bed which is in the reduced state should be taken into
account in simulations. This is done by means of a reduction factor,
introduced in previous work (De Groote and Froment, 1994).

b. Reactor model
A one-dimensional heterogeneous model, whereby the temperature
gradients over the external film are neglected, is applied. The model
equations are:

Continuity equations

O Q

-ρ-

164
Brought to you by | provisional account
Unauthenticated
Download Date | 5/1/18 11:12 PM
AM. De Groote and G.F. Froment Reviews in Chemical Engineering

Energy equation
cfT _
dz Ί

Pressure drop equation

As discussed elsewhere (De Groote and Froment, 1994), the intraparticle


diffusion limitations are accounted for by means of constant effectiveness
actors to avoid the time robbing integration of the differential equations
inside the catalyst particle at each increment in the reactor. In all
simulations the values of the effectiveness factors are:
ηΐ = 0.07 for CO-production by steam reforming,
η2 = 0.70 for the WGS-reaction,
η3 = 0.06 for CCVproduction by steam reforming,
η4= O.OS for the total combustion of Cft, to CO2 and H2O,
η5= 0.05 for the Boudouard reaction,
τ|6= 0.05 for methane cracking,
η7= 0.05 for C-gasification by steam, and
η8= 0.05 for C-gasification by oxygen.
Two models are applied, depending on the state of reduction of the Ni-
catalyst: the BV-model and the VDR-model. In the BV-model the catalyst is
of the bivalent type, so that total combustion and steam reforming proceed in
parallel. In the VDR-model steam reforming is partially consecutive to total
combustion, since the degree of reduction of the catalyst bed is taken into
account.

c. Simulation of an industrial adiabatic fixed bed reactor

The partial oxidation of methane with oxygen or air was simulated in a


fixed bed reactor for an inlet temperature of 535°C, a total pressure of 25 bar
and an Oa to OHU molar feed ratio of 0.598. The influence of a

165
Brought to you by | provisional account
Unauthenticated
Download Date | 5/1/18 11:12 PM
Vol. 11. No. 2. 1995 Reactor Modeling and Simulations
in Synthesis Gas Production
steam/methane ratio of 1.4 and of 10 mol% carbon dioxide in the feed is
studied. The molar flow rate of methane, F°CH4, is kept constant in all the
simulations so that only the total molar flow rate is adjusted.
The simulations of the partial oxidation of methane with oxygen have
been described in detail elsewhere (De Groote and Froment, 1994) and only
the temperature profiles through the catalyst bed obtained by the BV-model
(Figure 9) and by the VDR-model (Figure 10) are given here. The maximum
bed temperature predicted by the VDR-model is much higher than that
obtained by the BV-model, because steam reforming starts further
downstream in the reactor. For a feed containing methane and oxygen only,
a peak temperature of about 1500°C, which is detrimental for the Ni-
catalyst, is predicted by the VDR-model. For a CH4/O2/H2O-mixture the
agreement was excellent between simulated and industrial values (De Groote
and Froment, 1994).
The results predicted by the VDR-model for the partial oxidation of
methane with air instead of oxygen are given in Table 3 and Figures 11 to

1600

1400-

1200

1000

800
3:CH4,O2,H2O,C02j
«CH4.O2.CO2
600
2 3 4
Axial position z (m)

Fig. 9: Temperature profiles through the catalyst bed for the partial
oxidation of methane with oxygen at 25 bar. BV-model. Influence
of the addition of steam and carbon dioxide to the feed.

166
Brought to you by | provisional account
Unauthenticated
Download Date | 5/1/18 11:12 PM
A .M. De Groote and G.F. Froment Reviews in Chemical Engineering

2000
Feed: l
1: CH4,02, H2O
2CH4.02
3: CH4,02, H2O, CO2
4: CH4,02, CO2

600
1 U 2
Axial position z (m)

Fig. 10: Temperature profiles through the catalyst bed for the partial
oxidation of methane with oxygen at 25 bar. VDR-model.
Influence of the addition of steam and carbon dioxide to the feed.

1400

1300

1200

1100 -

5 1000
Ό

i 900 Fwd:
1:CH4,O2.N2,H2O
BOO 2:CH4,02.N2
3:CH4,02,N2,H20,COi
4:CH4,O2,N2,CO2
700
0 1 2 3 4 β β
Axial position z (m)
Fig. 11: Temperature profiles through the catalyst bed for the partial
oxidation of methane with air at 25 bar. VDR-model. Influence of
the addition of steam and carbon dioxide to the feed.

167
Brought to you by | provisional account
Unauthenticated
Download Date | 5/1/18 11:12 PM
Vol. 11, No. 2. 1995 Reactor Modeling and Simulations
in Synthesis Gas Production
Tables
Simulation results for the catalytic partial oxidation of methane with air for
an inlet temperature of 535°C VDR-model. Influence of the addition of
steam and CO2 to the feed mixture.

Feed 1 Feed 2 Feed 3 Feed 4

P°t (bar) 25.33 25.33 25.33 25.33

*°t 18375 13498 20417 14998


(Nm3/hr)
H20/CH4 1.4 0.0 1.4 0.0

y°<»2 0.0 0.0 0.1 0.1

*CH4 <*) 94.65 98.17 84.36 88.66

*b2 < * > 98.12 100 92.72 100

*CO < * ) 58.51 70.05 65.69 94.78

XC02 < % ) 32.16 7.67 14.66 -12.60


T
m« < K > 1231 1292 1128 1168

axmax (m) 3.37 5.00 4.43 2.97

Tout (K) 1147 1291 1059 1139

Pt.out 24.33 24.55 24.18 24.52


(bar)
H2/CO 3.33 2.32 2.33 1.34

FCO 2038 2440 2287 3301


(Hm3/hr)

*H2 6785 5648 5524 4435


(Hm3/hr)
C-bal 96.02 79.55 97.47 95.47

19. The maximum temperatures in the bed are considerably lower and
within the operating bounds for the Ni-catalyst. The temperature peak is
shifted further downstream in the reactor and is broadened, mainly due to
the dilution effect. The temperature decrease near the reactor inlet in Figure

168
Brought to you by | provisional account
Unauthenticated
Download Date | 5/1/18 11:12 PM
AM. De Groote and G.F. Froment Reviews in Chemical Engineering

100

1:CH4,02,N2.H20
2:CH4,O2,N2
3:CH4.O2,N2.H20.COa
4:CH4.O2,N2.CO2

0 1 2 3 4 5 6
Axial position z (m)

Fig. 12: Oxygen conversion through the catalyst bed for the partial
oxidation of methane with air at 25 bar. VDR-model. Influence of
the addition of steam and carbon dioxide to the feed.

too

eo

s «o

β 40.-

20 1:Cm.OS,«,HJO
fcCH4.Oa.N2
*CH4,OZ,N2.H2O,COZ
*CH4,O3,K2,CO2

2 3 4 5
Axial petition z (m)

Fig. 13: Methane .conversion through the catalyst bed for the partial
oxidation of methane with air at 25 bar. VDR-model. Influence of
the addition of steam and carbon dioxide to the feed.

169
Brought to you by | provisional account
Unauthenticated
Download Date | 5/1/18 11:12 PM
Vol. 11. No. 2. 1995 Reactor Modeling and Simulations
in Synthesis Gas Production

100
FMd:
1: CH4,02, N2, H2O
2:CH4.0a.NJ
80
3: CH4,02, N2, H2O, CO2
«CH4.O2.N2.C02

60

40-

20-

0 1 2 3 4 5 6
AxM portion z (m)

Fig. 14: Conversion of methane to carbon monoxide through the catalyst


bed for the partial oxidation of methane with air at 25 bar. VDR-
model. Influence of the addition of steam and carbon dioxide to
the feed.

100
FMCt
1:CH«,02,N2.H20
2: CM. O2, HZ
3: CH4,02. N2, H2O, COZ
«CM, 02. na.cn

2 1 4
AIM position z (m)

Fig. 15: Conversion of methane to hydrogen through the catalyst bed for
the partial oxidation of methane with air at 25 bar. VDR-model.
Influence of the addition of steam and carbon dioxide to the feed.

170
Brought to you by | provisional account
Unauthenticated
Download Date | 5/1/18 11:12 PM
A M. De Groote and G.F. Frament Reviews in Chemical Engineering

0.6
Fwd:
1: CH4. O2, N2, H2O
0.5 £CH4,O2,N2
f 3: CH4,02, N2, H2O, CO2I
4.-CK4,O2.N2.COZ
l 04
CH4 + 2 02 — > CO2 + 2 H2O
I
S 03 4 1
f
S 02 -
S.
01 -

0 1 2 3 4 5 0
Axial position z (m)

Fig. 16: Total combustion rate of methane to carbon dioxide and steam
through the catalyst bed for the partial oxidation of methane with
air at 25 bar. VDR-model. Influence of the addition of steam and
carbon dioxide to the feed.

1:CH4.O2.N2,H2O
£CH4,O2,N2
3: CH4,02,«, H2O, CO2
«CW,O2,N2,CO2

2 3 4
Axial position z (m)

Fig. 17: Rate of carbon monoxide production by steam reforming through


the catalyst bed for the partial oxidation of methane with air at 25
bar. VDR-model. Influence of the addition of steam and carbon
dioxide to the feed.

171
Brought to you by | provisional account
Unauthenticated
Download Date | 5/1/18 11:12 PM
Vol. 11. No. 2,1995 Reactor Modeling and Simulations
in Synthesis Gas Production

1:CH4,OJ.N2,H20
£CH4,02,N2
*cm,oa.N2,H20,coi
«CH4.02.N2.C02

0 0.5 1 U 2
AxW petition z (m)

Fig. 18: Net coking rate through the first part of the catalyst bed for the
partial oxidation of methane with air at 25 bar. VDR-model.
Influence of the addition of steam and carbon dioxide to the feed.

Axled position ζ (πι)

Fig. 19: Rate of the coking reactions for feed 2 through the first part of the
catalyst bed for the partial oxidation of methane with air at 25
bar. VDR-model.

172
Brought to you by | provisional account
Unauthenticated
Download Date | 5/1/18 11:12 PM
AM. De Groote and G.F. Froment Reviews in Chemical Engineering

11 is caused by methane cracking, as can be seen in Figures 18 and 19. The


total combustion reaction leads to a significant temperature rise. The axial
position of the top of the temperature peak in the reactor corresponds to the
reactor distance where the heat effects of the steam reforming reactions
become dominant. The temperature increase towards the reactor outlet can
be explained by Figures 12, 16 and 17. Behind the temperature peak total
combustion still occurs until all oxygen is consumed, albeit at a slower rate.
Since the steam reforming reactions have also slowed down because of the
lower temperature, the heat effect of the combustion becomes predominant
again.
For the same inlet temperature as in the corresponding cases with oxygen
only, the conversions of CHj and Qj and the conversion to CO are
considerably lower. This results from the lower reaction rates of total
combustion and steam reforming, caused in turn by the lower temperatures.
For higher inlet temperatures the conversions are increased, but methane
cracking becomes more important, thus leading to serious catalyst deacti-
vation. The simulation results for the partial oxidation of methane with air
with a higher inlet temperature of 635°C, but the same total pressure of 25
bar and initial molar flow rate of methane are presented in Table 4.
It appears from Tables 3 and 4 that the addition of CQ2 and/or H2O leads
to a significant change in the H2/CO-product ratio of the synthesis gas. The
addition of COj increases the C/O- and CYH-ratio of the feed, thus leading to
a H2-poor syngas. In the presence of steam the C/H-feed ratio is decreased
and a synthesis gas with higher Hz/CO-ratio is processed. Changes in the
feed composition are reflected in the carbon content as well. It is clear from
Tables 3 and 4 that coke deposition is inhibited by steam and CO2.
It follows from previous work (De Groote and Froment, 1994) that the
addition of Ü2 to the feed only prevents the carbon deposition due to
methane cracking near the reactor inlet, but increases the rate of methane
cracking in the vicinity of the temperature peak, which is higher.

173
Brought to you by | provisional account
Unauthenticated
Download Date | 5/1/18 11:12 PM
Vol. 11. No. 2.1995 Reactor Modeling and Simulations
in Synthesis Gas Production
Table 4
Simulation results for the catalytic partial oxidation of methane with air for
an inlet temperature of 635°C VDR-model. Influence of the addition of
steam and CQz to the feed mixture.

Feed 1 Feed 2 Feed 3 Feed 4

p°t (bar) 25.33 25.33 25.33 25.33

F°t 18375 13498 20417 14998


(Mm 3 /hr)
H20/CH4 1.4 0.0 1.4 0.0

y°co2 0.0 0.0 0.1 0.1

*CH4 < * ) 98.53 99.59 94.85 93.79

*02 < * > 98.27 99.77 98.14 100.00

*co (*) 61.40 68.03 78.05 99.57

*C02 < * > 28.05 6.68 7.99 -15.82

T
max < K > 1300 1380 1192 1236

axmax (m) 1.60 6 2.02 1.30

Tout W 1215 1380 1122 1178

Pt,out 24.10 24.41 23.88 24.35


(bar)

H2/CO 3.21 2.37 2.13 1.37

FCO 2138 2369 2718 3468


3
(Nm /hr)

*Η2 6864 5617 5795 4735


3
(Nm /hr)
C-bal 90.91 75.11 94.44 92.98

d. Simulation of an adiabatic reversed flow reactor

Overall the catalytic partial oxidation of methane to synthesis gas is an


exothermic process and thus suited for reversed flow operation of a fixed bed

174
Brought to you by | provisional account
Unauthenticated
Download Date | 5/1/18 11:12 PM
A.M. De Groote and G.F. Froment Reviews in Chemical Engineering

reactor. This alternative mode of reactor operation may lead to lower syngas
manufacturing costs than the conventional unidirectional fixed bed reactor,
since external heat exchangers are no longer required.
Partial oxidation reactors operated with reversed flow of the feed have
been investigated by Blanks et al. (1990). In a reactor of the "reversed flow"
type, the catalyst bed is preheated to a uniform temperature and the reactant
mixture is fed at considerably lower temperature, so that transients develop
in the reactor. The cold feed is progressively heated by the catalyst bed. At a
certain depth the exothermic combustion reaction takes off, leading to a
temperature rise in the bed. Consequently, temperature and concentration
waves are generated which move at a constant velocity through the catalyst
bed in the flow direction of the gases. Before the reaction front reaches the
end of the catalyst bed, the flow direction of the gases is reversed, so that the
heat front starts moving in the opposite direction. In this mode of reactor
operation the catalyst bed is used simultaneously as chemical reactor and
heat exchanger.
For the modeling of the reversed flow partial oxidation process a one
dimensional heterogeneous model, accounting for intraparticle and
interfacial gradients, is developed. The model consists of differential
equations for both gas phase and solid phase. Transient terms are
incorporated in the differential equations to enable the modeling of the
traveling waves. A simplified model is used to reduce the tremendous
computer times needed for these simulations. The model equations assuming
a steady state bulk gas phase, a steady state gas phase in the pellet and
constant effectiveness factors are given below. Because of the introduction of
constant effectiveness actors, derived from the steady state simulations, the
integration of the continuity equations over the catalyst particle is no longer
required.
The model equations for the gas phase then become:

• Continuity equation for component j :

175
Brought to you by | provisional account
Unauthenticated
Download Date | 5/1/18 11:12 PM
Vol. 11. No. 2, 1995 Reactor Modeling and Simulations
in Synthesis Gas Production
• Change in the total molar flow rate:

Energy equation:

• Pressure drop equation:

dz pe ε' -"-* dp
The energy equation for the solid phase is:

dTa θ2!*, J^
M
*" dt dz2 v
" i-i

The above set of equations has to be integrated with the following set of
initial and boundary conditions:

0) - yjo
T(z,0) - T0(z)
Γ,(ζ,Ο) - Tf0(z)

y,(Q,t) - yf
p e (0,t) -p t °
Ft(Q,t) - Fc°
T(0,t) - T°
r s (Q,t) - rs°
•(Z, t) - 0 . z-Q.Z

The effectiveness factors are:


ηι = 0.04 for total combustion,
η2 = 0.03 for CO-production by steam reforming,
η3 = 0.06 for the WGS-reaction, and
η4= 0.03 for COj-production by steam reforming.

176
Brought to you by | provisional account
Unauthenticated
Download Date | 5/1/18 11:12 PM
AM. De Groote and G.F. Froment Reviews in Chemical Engineering

An example of the traveling temperature waves during the reversed flow


operation of a partial oxidation reactor is shown in Figure 20 for a semi-
cycle in the permanent regime. The reactor was operated at atmospheric
pressure. The catalyst bed was preheated to a uniform temperature of 800°C.
The feed gas, with a CHi/Oj-ratio of 2 and containing 10% H^, was
introduced to the catalyst bed at a temperature of 300°C. The flow was
reversed every 350 s.
The conversions to hydrogen and carbon monoxide at the reactor exit are
given in Figure 21. The dip following the flow reversal can be explained by
the shape of the temperature profile during the first seconds of a semi-cycle.
The total combustion of methane and the steam reforming reactions are
more or less consecutive, so that the flow has to be reversed before the
temperature peak is at the reactor outlet (Figure 20, curve 8). Because of this
the temperature peak is not found near the reactor inlet just after flow
reversal (Figure 20, curve 1). The total combustion reaction takes off in the
zone of the reactor in front of the temperature peak, thus leading to a

Solid temperature Ts (K) Semi-cycle 6


1400
1:t- O2 s
2:1= 4&2s
3: t »100.2 s
4:t=*14&2*
1200

1000

800

600

400
Ο 0.2 0.4 0.6 0.8 1
Dlmenslonless axial position z/Z
Fig. 20: Solid temperature through the catalyst bed for various moments in
a semi-cycle of the permanent regime.

177
Brought to you by | provisional account
Unauthenticated
Download Date | 5/1/18 11:12 PM
Vol. 11, No. 2.1995 Reactor Modeling and Simulations
in Synthesis Gas Production

100

90 FH FR FR

80

CO
70
Ο
υ
60

50
1300 1500 1700 1900 2100
Clock time t (s)
Fig. 21: Conversion to hydrogen and carbon monoxide at the exit of a
reversed flow reactor for catalytic partial oxidation as a function
of clock time. (FR = flow reversal)

temperature profile with two peaks. As the semi-cycle proceeds, the first
peak develops further by the exothermic total combustion reaction, while the
second peak gradually disappears because of the endothermic reforming
reactions. This leads to a smaller contribution of the reforming reactions to
the overall partial oxidation process. Once the reaction front is completely
developed, it wanders through the catalyst bed at a constant velocity (Figure
20, curves 2 to 8), leading to a conversion to hydrogen of about 75% and to
carbon monoxide of about 74%.
The maximum temperature in the catalyst bed amounts to about 1030°C,
which is not detrimental for the Ni-catalyst. Selectivities, averaged over the
semi-cycle, to H? and CO of respectively 93.5% and 91.5% are obtained,
comparable to those in a classical reactor.

4. CONCLUSIONS

The present review of the simulation of various processes for synthesis


gas production leads to considerable insight into these complex processes

178
Brought to you by | provisional account
Unauthenticated
Download Date | 5/1/18 11:12 PM
A.M. De Groote and G.F. Frament Reviews in Chemical Engineering

and reveals operation and zones where excessive temperatures are obtained.
By adjusting H2O/CH4-, CC^/Cftj- and Oa/CHHfeed ratios, a synthesis gas
with the desired composition can be produced. The presence of steam and
CÜ2 in the feed mixture reduces the amount of carbon deposited on the
catalyst surface during the catalytic partial oxidation. Although CC>2 inhibits
the carbon formation in catalytic partial oxidation, it is not efficient in
reducing coke in steam reforming, where CQz favours methane cracking
through the whole catalyst bed. Recycling H2 to slow down methane
cracking is too expensive and leads to a syngas with high H2/CO ratio,
undesirable for methanol synthesis and Fischer-Tropsch synthesis.
Reversed flow operation of a catalytic partial oxidation reactor has been
shown to be feasible. The conversions and selectivities obtained with this
alternative way of reactor operation are good.

5. REFERENCES

Blanks, R.F., Wittrig, T.S. and Peterson, D.A., Chem. Eng. Sei., 45(8),
2407-2413 (1990).
Boudart, M, Mears, D.E. and Vannice, M.A., Congres International Chimie
Industrielle, Industrie Chimique Beige, 32(1), 281 (1967).
De Deken, J.C., Devos, E.F. and Froment, G.F., "Chemical Reacion
Engineering", inA.C.S. Symp. Series, 196.
De Groote, A.M. and Froment, G.F., Hydrocarbon Processing (submitted
for publication, 1994).
Froment, G.F. and Bischoff, K.B., Chemical Reactor Analysis and Design,
J. Wiley and Sons, New York, 1990.
Hottel, H.C. and Sarofim, A.F., Radiative Transfer, McGraw-Hill, New
York, 1967.
Michel, S., Hydocarbon Processing, April, 37-39 (1989).
Plehiers, P.M. and Froment, G.F., Chem. Eng. Technol., 12,20-26 (1989).
Rao Ramana, M.V., Plehiers, P.M. and Froment, G.F., Chem. Eng. Sei.,
43(6), 1223-1229 (1988).

179
Brought to you by | provisional account
Unauthenticated
Download Date | 5/1/18 11:12 PM
Vol. 11, No. 2,1995 Reactor Modeling and Simulations
in Synthesis Gas Production
R0strup-Nielsen, J.R., Catal. Today, 18,305-324 (1993).
Snoeck, J.W. and Froment, G.F. (unpublished results).
Trimm, D.L. and Lam, C. W., Chem. Eng. Sei.. 35,1405-1413 (1980).
Wagner, E.S. and Froment, G.F., Hydrocarbon Processing, 7,69-78 (1992).
Xu, J. and Froment, G.F., AIChE Journal, 35(1), 88-96 (1989a).
Xu, J. andFroment, G.F.,AIChEJournal, 35(1), 97-102 (1989b).

SYMBOLS

[A] mole percentage of component A %


ΑΪ surface area of zone i m2
Aj pre-exponential factor for rate constants
av external particle surface per unit reaactor volume mp2/mr3
Αχ pre-exponential factor for adsorption constants
axmax axial position where the maximum temperature occurs m
Cp specific heat of fluid kJ/kgK
Cps specific heat of solid kJ/kg K
DeA effective diffusivity mfVm s
di,dti inner tube diameter m
Knudsen diffusivity for component A mfVm s
effective molecular diffusivity of A in a multi-
component mixture mfVm s
do outer tube diameter m
dp particle diameter nip
Eaj activation energy kJ/kmol
f friction factor in Fanning equation
F°, F°t total molar feed rate kmol/hr
F°A molar feed rate of reactant A kmol/hr
F°jj view factor between zones i and j in a transparent
medium
FA flow rate of component A Nm3/hr
Fij view factor between zones i and j in a real, absorbing
medium
180
Brought to you by | provisional account
Unauthenticated
Download Date | 5/1/18 11:12 PM
AM. De Groote and G.F. Frament Reviews in Chemical Engineering

Ft total molar flow rate kmol/mr2s


g acceleration of gravity m/s2
hf gas-solid heat transfer coefficient kJ/mp2 s K
Ts solid temperature K
H& heat of formation of species i kJ/kmol
Hi molar enthalpy of species i kJ/kmol
Ηχ» adsorption enthalpy of species X kJ/kmol
KA adsorption constant of species A
KB* threshold for Boudouard coking bar1
KB° equilibrium constant for the Boudouard reaction
where coking is just detected bar1
KBB equilibrium constant for the Boudouard reaction
where gasification is just detected bar1
Keq.i equilibrium constant of reaction i
kgj,kgj gas/solid mass transfer coefficient mf3/mp2 s
Ki adsorption and equilibrium constants
ki reaction rate constant
KM* threshold constant for methane cracking bar
KM° equilibrium constant for methane cracking where
coking is just detected bar
6
KM equilibrium constant for methane cracking where
gasification is just detected bar
KR* threshold constant for CO-reduction bar1
KRC equilibrium constant for CO-reduction where coking
is just detected bar1
equilibrium constant for CO-reduction where
gasification is just detected bar1
length of reactor m
heated tube length m
mean molecular weight kg/kmol
PA partial pressure of component A bar
PP partial pressure of a product bar
pr partial pressure of a reactant bar

181
Brought to you by | provisional account
Unauthenticated
Download Date | 5/1/18 11:12 PM
Vol. U. No. 2. 1995 Reactor Modeling and Simulations
in Synthesis Gas Production

P, vector of partial pressures


partial pressure of component A inside the pellet bar
total pressure at the reactor outlet bar
Pbp°t total pressure bar
R gas constant kJ/kmolK
RDS rate determining step
η reaction rate kmol/m3 s
Rp equivalent radius of the catalyst particle m
S beam length m
Sx° standard entropy of species X kJ/kmolK
Sx.»° entropy of species X in the adsorbed state kJ/kmolK
Sx,g° standard entropy of gas component X kJ/kmolK
T temperature K
t clock time s
T° initial temperature K
To temperature at the inlet K
Tj internal tube skin temperature K
Tmax maximum temperature K
Tout outlet temperature K
TsO initial solid temperature K
Ts° solid temperature at the inlet K
Tw external tube skin temperature K
Us superficial velocity m3o/m2r s
Veq(i) approach to equilibrium in reaction i
coking limit of reaction i
black body emissive power kW/m2
conversion of methane %
conversion of methane to CO %
conversion of methane to CCh %
conversion of oxygen
y°A mole fraction of species A in the feed
yA mole fraction of species A
Vj0 initial mole fraction of species j

182
Brought to you by | provisional account
Unauthenticated
Download Date | 5/1/18 11:12 PM
AM. De Groote and G.F. Froment Reviews in Chemical Engineering

yiss,yiss m
ole fraction at the solid surface
z axial position ην
^Zj total exchange area between zones i and j m2

Greek symbols

OP stoichiometric coefficient of product


ctf stoichiometric coefficient of reactant
i, j inclination to normal to the surface rad
ε void fraction of packing
sg porosity
ηί,η,ηχ effectiveness factor
Xea axial effective thermal conductivity in a packed bed kJ/nvsK
μ dynamic viscosity kg/m s
v; stoichiometric coefficient of component i
ξ radial position in the pellet m
PB catalyst bulk density kgcat/nv3
pg gas density kg/m3r
PS density of solid kg/rap3
σ Stefan-Boltzmann constant W/m2 K4
τ tortuosity factor
Ω cross section of reactor m2

183
Brought to you by | provisional account
Unauthenticated
Download Date | 5/1/18 11:12 PM

S-ar putea să vă placă și