You are on page 1of 14

Pergamon

Chemical Engineering Science, Vol. 52, No. 10, pp. 1609-1622, 1997
1997 Elsevier Science Ltd. All rights reserved

PII: $1111O9-25119(91~)11@511-8

Printed in Great Britain

0009-2509/97 $17.00 + 0.00

The fluidized-bed membrane reactor for


steam methane reforming: model
verification and parametric study
A. M. Adris,* C. J. Lim and J. R. GracC
Department of Chemical Engineering, University of British Columbia, Vancouver, BC,
Canada V6T 1Z4
(Received 19 April 1996; in revised form 25 November 1996; accepted 27 November 1996)
Abstraet--A new approach is presented for the modelling of a fluidized-bed membrane reactor
(FBMR). The model considers the two-phase nature of the fluidized-bed reactor system and the
parallel reactions taking place in stream methane reforming, as well as selective permeation
through the walls of membrane tubes immersed in the bed. The model is based on the
two-phase bubbling bed model with allowance for some gas flow in the dense phase. Plug flow
is assumed for the combined sweep gas and permeating hydrogen flowing through the membrane tubes. Freeboard non-isothermal effects and reactions are also taken into account. The
coupled differential equations for the fluidized bed and membrane tubes are solved numerically.
The model is in very good agreement with experimental data, both with and without permeation, obtained in a pilot-scale reactor system. Parametric investigations demonstrate the
effect of key operating variables and design parameters over a wide range, The model is also
tested for its sensitivity to changes in hydrodynamic parameters. Increasing the permeation of
hydrogen through the membrane tubes is of key importance in achieving high methane
conversions and in minimizing adverse reactions in the freeboard region. Hydrodynamic and
kinetic properties have limited influence for the conditions studied. 1997 Elsevier Science
Ltd. All rights reserved
Keywords: Fluidization; membranes; permeation; reforming; hydrogen; reactor modelling.

1. INTRODUCTION
Steam reforming of light hydrocarbons, especially
natural gas, is an industrially important chemical reaction and a key step for producing hydrogen and
syngas for ammonia and methanol production, hydrocracking and hydrotreating, oxo-alcohol and
Fischer-Tropsch synthesis and other important processes in the petroleum and petrochemical industries.
However, industrial fixed-bed steam reformers suffer
from several problems which seriously affect their
operation and performance. These include low catalyst effectiveness due to internal mass transfer resistance in the large catalyst particles, low heat transfer
rates, large temperature gradients and thermodynamic equilibrium constraints. These problems with
conventional reformers can be alleviated by using
a fluidized-bed membrane reactor (FBMR) system for
steam methane reforming (SMR) (Adris et al., 1994).
This new reactor system combines several advantages

* Present address: Department of Research and Technology Support, SABIC R&D, P.O. Box 42503, Riyadh 11551
Saudi Arabia.
, Corresponding author.

of fluidized beds as chemical reactors, in particular


catalyst bed temperature uniformity, improved
heat transfer and virtual elimination of intracatalyst
diffusional limitations, with advantages offered by
permselective membrane technology, in particular
shifting the conventional thermodynamic equilibrium
and in situ separation and removal of a desirable
reaction product in a very pure state.
In a related study (Adris, 1994), a pilot-scale reforming plant of internal diameter 102 mm with a hydrogen capacity of 6 m3[STP]/h, able to operate at temperatures up to 750C and pressures up to 1.5 MPa,
was commissioned to examine the concept of the new
fluidized-bed membrane reactor system for steam reforming of natural gas and to study its properties. The
present paper considers the modelling of the FBMRSMR reactor.
The first attempt to model the FBMR system was
conducted by Adris et al. (1991), who simulated an
industrial-scale fluidized-bed membrane reformer in
order to explore the potential of this new reactor
system and compare it with conventional reforming
units. This earlier model was based on the Orcutt
model with perfect mixing of the dense phase gas
(Orcutt et al., 1962) and with a number of simplifying

1609

A.M. Adris et al.

1610

assumptions which made the system of equations similar to those for a CSTR. Those simplifications were
adequate for 'gross' prediction of the overall reactor
performance.
In order to represent the hydrodynamics better, the
simulation work in the present paper is based on the
two-phase bubbling bed reactor model (Grace, 1984)
with allowance for some gas flow (assumed to be in
plug flow) in the dense phase. This assumption was
also used in a second version of the Orcutt model
(Orcutt et al., 1962). Both versions of the model consider selective removal of hydrogen by permeation
through membrane tubes immersed in both the dense
bed and the dilute phase. Reaction taking place over
entrained catalyst particles in the freeboard zone is
also taken into account in the present model.

with both the bubble phase and the dense


phase. A 'sweep gas' such as steam is used
inside the membrane tubes to maintain a low
partial pressure of hydrogen there, helping to
promote permeation of hydrogen through the
walls of the tubes.
(7) Sweep and permeating gases are in plug flow
inside the membrane tubes.
(8) Gases leaving the bed surface travel in plug
flow and undergo chemical reaction in the
freeboard region over the surface of entrained
catalyst particles.
(9) The concentration of solids in the dilute freeboard phase decays exponentially with height.
(10) Hydrogen is the only species which permeates
through the membrane tube walls.
(11) Ideal gas behaviour is assumed.
(12) The temperature is assumed to be constant in
the dense bed region, but it may vary in the
freeboard region.

2. MODELDEVELOPMENT
2.1. Model assumptions
The phases considered in the present model are
illustrated schematically in Fig. 1. The model is based
on the following assumptions:
(1) Steady-state conditions are assumed.
(2) The dense catalyst bed is considered to be
composed of two phases, a bubble phase and
a dense phase.
(3) Axial dispersion of gas is ignored in both
phases.
(4) Reaction occurs mostly in the dense phase.
However, the bubbles contain some solids
which contribute to the overall reaction.
(5) In view of their small size, the diffusional resistance inside the catalyst particles is neglected.
(6) The interior of the membrane tube is taken as
a separate region which exchanges hydrogen

2.2. Model formulation


2.2.1. Dense catalyst bed equations. A mole balance
on component i (where i equals 1 for methane, 2 for
steam, 3 for carbon monoxide and 4 for carbon dioxide) in the bubble phase gives

dnib
dh = kiqabebA(Cid

-- Cib ) +

~bPsARib.

For i = 5 (hydrogen), the component mass balance in


the bubble phase is given by

dnib
dh = klqabg'bA(Cia -- Cib)
Pn ) Cevlg,b(Prl~
o s _ po~s)
+ (~bPsARib -- ~e(--~H

(2)
other
~L_ product

__[_hydrogen and

.~ stream

Permbate
~sweepgas
=

Lr----J

Freeboard
zone

oo

ph~sel D

Permeate
.~.F.~._~~.~
i
i gas
Membrane
tubes

(1)

z=o

~ h=H

~reactants

I~llSe

bed
h=O

zone

Fig. 1. Schematic showing phases considered in the fluidized-bed membrane reactor model.

1611

Modelling of a fluidized-bed membrane reactor


where C~pz is the equivalent permeation capacity per
unit length of the membrane tube, defined as the
external surface area of the membrane tubes divided
by the tube wall thickness per unit length.
Similarly, a mole balance on component i (i = 1-4)
in the dense phase gives
dnia
dh = kiqab~'bA(Cib

--

Cid) "q- Oap, ARia.

(3)

-- Cid)

+ ~tpsARid -

{pH'~
e \MH/]

x C,t,,(l --

s --

eb)(Pa

pO;S).

(4)

Finally, a mole balance on hydrogen (component


i = 5) in the separation phase (i.e. inside the membrane tubes) gives
dh

MHH {(1

-- eb)(Pd 5 --

+ eb(pO.b5 _ pO;5)}.

dni, = qJ~ARiz.
dz

(6)

For i = 5 (hydrogen) the dilute-phase equation is

dn,~dz- VzAR,z - *e ~

Cep,(P.z -- e~; 5) (7)

where z is the distance above the bed surface. The


boundary condition at z = 0 is n~ = n~d + n~b, evaluated at h =/-/. u?z is the solids mass concentration in
the freeboard zone given as: u?z = Ez/Uo, while E~ is
the entrainment flux of solids at a distance z above the
bed surface, given by Wen and Chen (1982) as

For hydrogen, the corresponding equation is


dnid
dh - kiqabebA(Cib

freeboard region gives

pO~5)
(5)

Boundary conditions at h = 0
are n~b =n~,
(Uo - Umi)/Uo, nld = niiU~f/Uo and ni~ = 0, where it
is assumed that the gas flow through the dense phase
is that needed for minimum fluidization.
It is common practice in modelling steam reforming
reactors to consider the conversion of two key components and then to obtain the concentrations of
the other components by applying stoichiometric
relations (Xu and Froment, 1989; Adris et al., 1991;
Soliman et al., 1992). The present model, however,
considers the change in the molar flow of each component separately so that differences in diffusivities can
be accounted for rigorously.
The reaction rate terms, Rib and Rid in eqs (1)-(4),
represent the rates of formation of component i in the
bubble or dense phase, essentially equal to the intrinsic rates since the external mass transfer resistance is
negligible for the size of catalyst particles considered.
Each R~ is the combination of the rates of formation
and disappearance of this component through the
three principal parallel reactions taking place in steam
reforming given in Appendix A. Reaction rate expressions and constants, equilibrium constants and adsorption constants are also given in Appendix A. The
Xu and Froment (1989) rate expression has been utilized due to its general nature and its ability to describe the kinetics of the reforming reaction over
a wide range of operating conditions (Elnashaie et al.,
1990).
The hydrodynamic parameters required to solve
eqs (1)-(5) are obtained using standard relations from
the literature as summarized in Appendix B.
2.2.2. Dilute phase (i.e. freeboard region) equations. A mole balance on component i (i = 1-4) in the

Ez = E~ + (Eo -- Eo~)e -"cz

(8)

where Eo is the entrainment flux of solids at the bed


surface, Eo~ is the entrainment flux of solids above the
transport disengagement height (TDH) and a~ is an
overall decay constant, taken here as 4 m - 1 as suggested by Geldart (1986). E0 and E~ were also calculated according to the procedure given by Geldart
(1986).
The selective separation of hydrogen by the membrane tubes in the freeboard zone is described by
a modified form of eq. (5):
dn,~ _ , ( p , "~
dzz
e \ M H ] Cept(P~ -- pO~5).

(9)

2.3. Solution algorithm


The solution algorithm is composed of three modules:
(1) The first module solves for reaction and permeation in the catalyst bed zone.
(2) The second module solves for reaction and permeation in the part of the freeboard zone which
includes membrane tubes.
(3) The third module solves for the reaction in any
part of the freeboard zone that does not include
permeable membrane tubes.
Each module first calculates the parameters needed
by the differential equations describing the phenomena taking place in that part of the reactor and
then solves the model differential equations by calling
a numerical routine N L D E Q D which uses the
Runge-Kutta method, with variable step size to ensure solution accuracy. All variables are defined as
double precision variables to improve the accuracy
when dependent variables are varying steeply. Due to
a delicate balance between the three terms involved in
the differential equations describing the change in
molar rates in the bubble and dense phases, integration had to be started with a very small step size,
typically 10 -1 m. Temperature profiles were based
on experimental measurements in the dense bed and
in the freeboard (Adris, 1994), with the temperature

1612

A. M. Adris et al.

assumed to be uniform below the bed surface. An


experimental value of the expanded bed height was
also used, while the permeation capacity was based on
measurements carried out on palladium tubes in
a separate permeation rig (Adris, 1994). The limiting
case of a fluidized-bed reactor without membrane
separation can be solved by setting Cept = O.

3. MODELVALIDATION
3.1. Model predictions vs experimental data
Experimental results for the pilot-scale FBMRSMR investigation have been reported elsewhere
(Adris, 1994; Adris et al., 1994b) for steam-to-carbon
ratios from 2.3 to 4.2, reaction temperatures from 720
to 930 K and pressures from 0.69 to 0.98 MPa. The
model developed above can be tested using these
experimental data both with and without selective
permeation. The volume fraction of the bed occupied
by solids dispersed in the bubble phase is taken as
0.5% of the volume fraction of the bed occupied by
bubbles, i.e. ~b = 0.005 eb (see Grace, 1986; Kunii and
Levenspiel, 1991). A very good match between experimental results and model predictions was obtained
as shown by Fig. 2. Deviations between the model
predictions and the experimental data are somewhat
larger for the reaction runs without permeation,
where predictions are more sensitive to any error in
temperature, because temperature is the major factor
determining conversion when there is no selective
separation.
For reaction-permeation runs, any error in the
temperature reading affects the predictions to a lesser
extent because the reacting mixture is shifted from
equilibrium due to the irreversible in situ hydrogen
separation. The contribution of permeation to the
overall conversion, together with the use of effective
permeability constants measured and fitted in the
experimental study (Adris et al., 1994b), explain the
better match between model predictions and experimental data for the reaction-permeation runs. The

0.8
0.7-

No permeation

>~ 0.60

0.5-

mean absolute value of the deviation of the predicted


methane conversion from the experimental conversion was 2.4% for reaction-permeation experiments
and 4.2% for runs without permeation.
Typical comparisons between simulation predictions and experimental data are given in Table 1 for
a run without permeation and for a reaction-permeation experiment. It is clear from this table that the
model provides good predictions for the reaction conversions, the total hydrogen yield and the permeating
hydrogen flow. There are, however, differences in the
outlet gas composition, especially in the ratio between
carbon oxides produced by the reaction. This may be
due to the difference between the catalyst used here
and that used by Xu and Froment (1989) to develop
the kinetic rate expressions employed in the simulation. Different reforming catalysts have different abilities to catalyse the water-gas shift reaction (RostrupNielsen, 1984). Therefore, the fraction of the carbon
monoxide converted to carbon dioxide by this reaction may vary depending on the catalyst. However,
this variation should not have a major influence on
the predicted overall methane conversion, whether to
carbon monoxide or carbon dioxide.
The results presented in Table I also indicate (by
difference) the contribution made by the entrained
catalyst particles in the splash zone (i.e. region immediately above the dense bed upper surface) to the
reaction conversion. Because of the heater design in
our experimental reactor, the splash zone temperature
is higher than the bed temperature, and therefore the
catalyst particles in the freeboard contributed positively to the reaction by about 7% of the overall
conversion. In an industrial reformer, however, the
effect of the entrained catalyst particles is likely to be
negative due to lower temperatures in the freeboard
zone, unless an equilibrium shift is affected by means
of permselective membranes as discussed below or
unless the freeboard region is specially heated.
The model developed and validated above includes
standard hydrodynamic equations and correlations
for bubbling fluidized beds and utilizes kinetic rate
expressions from the literature and experimentally
measured temperatures, bed heights and Cept values.
Otherwise there are no fitted parameters. Note that
with the abundance of catalyst in this reactor system,
the reaction conversion is highly influenced by the
thermodynamic equilibria, as well as by hydrogen
permeation rates.

~1~ 0.4-

~ 0.2~ 0.1
13. 0.0

0.0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

Experimental methane conversion


Fig. 2. Experimental vs predicted methane conversion. For
experimental details, see Adris (1994) and Adris et al. (1994b).

3.2. Sensitivity analysis


The model is examined here for its sensitivity to
four estimated parameters: solids concentration in the
bubble phase, distribution of feed gas between the
bubble and dense phases, catalyst activity and bubble
size. Permeation was omitted at this stage to avoid its
contribution which may obscure the effect of these
parameters. Sensitivity tests were restricted to
the catalyst bed conversion, excluding the freeboard
effects which vary from one reactor to another
depending on the reactor configuration and catalyst

Modelling of a fluidized-bed membrane reactor

1613

Table 1. Comparison of model predictions with experimental data for two typical runs. Operating
conditions: T = 541C, P = 0.64 MPa, @b = 0.005eb, Ua = Umf
Experimental
data
(w/o penn.)

Model
predictions
(w/o perm.)

0.400
0.182
NA
1.421
0

0.391
0.177
0.363
1.471
0

Exit methane conversion


Exit steam conversion
Methane conversion at bed surface
Total hydrogen yield
Permeate hydrogen flow, mol/h
Outlet gas composition,
volume % (dry basis):
CH4
CO
CO2
He

24.0
0.8
16.5
58.7

Experimental
Model
data
predictions
(with perm.) (with perm.)
0.419
0.200
NA
1.515
2.31

24.8
1.0
14.9
59.3

0.421
0.183
0.382
1.550
2.39

22.4
0.8
16.8
60.0

22.7
1.2
15.3
60.8

NA: not available.

Table 2. Effects of solids concentration in the bubble phase on the methane conversion:
T = 700C, P = 1.5 MPa, Fc = 60 mol/h, S/C = 3.5
Volume fraction of bed
occupied by bubble phase
solids, qbb

0.0

0.0001

0.0005

0.001

0.005

Methane conversion

0.569

0.570

0.573

0.575

0.579

Table 3. Model sensitivity to gas flow through the dense phase: T = 700C,
P = 1.5 MPa, Fc = 60 mol/h, S/C = 3.5
Fraction of gas flowing
through dense phase,
Ga/A U,,,s

0.01

O.1

0.5

1.0

5.0

Methane conversion

0.554

0.555

0.562

0.568

0.579

particle-size distribution. The bed temperature was


assumed to be uniform in these tests, and all variables,
except the one under investigation, were maintained
constant in each case.
3.2.1. Model sensitivity to solids concentration in the
bubble phase. The volume fraction of the bed occupied by bubble phase solids, ~b, was varied from 0 to
0.005. Methane conversion was predicted for five different cases as presented in Table 2. The increase in
reaction conversion with increasing solids concentration in the bubble phase may be attributed to a reduction in bubble by-passing, i.e. to conversion in the
bubble phase over the surface of the dispersed catalyst
particles. The insensitivity of the results to ~b
arises because the conversion rapidly approaches
an equilibrium value, with limited influence of hydrodynamics.
3.2.2. Model sensitivity to gas distribution between
bubble and dense phases. Simulation results presented

above considered the gas flow through the dense


phase to be that needed for minimum fluidization,
which is the reference case considered in this study.
The sensitivity of the model to this assumption was
examined by varying the gas flow though the dense
phase at h = 0 from five times to 1% of the flow
needed for minimum fluidization. The simulation resuits summarized in Table 3 indicate that the model is
not very sensitive to this parameter. When the flow
through the dense phase is taken as 1% of that needed
for minimum fluidization, approximating the stagnant bed assumption adopted by Kunii and Levenspiel (1969) and Grace (1984), the methane conversion
is predicted to be reduced, but by very little.
Again, bubble by-passing accounts for this change
in methane conversion upon changing the gas distribution between the bubble and dense phases. The
simulated runs in this study are characterized by low
superficial gas velocities, only about six times the
minimum fluidization velocity. It is anticipated that
the sensitivity of the model to this parameter would be

1614

A. M. Adris et al.

even less at higher superficial gas velocities more


typical of industrial fluidized-bed reactors.
3.2.3. Model sensitivity to catalyst activity. The
kinetic rate expressions used to estimate the rates of
reaction were developed (Xu and Froment, 1989) for
a different catalyst with a different nickel content. The
catalyst used in our work might have a different
activity. While the difference in catalyst activity between two commercial catalysts is unlikely to affect
the reaction rate estimation by more than 50% (E1nashaie et al., 1990), the activity was changed here
over a wide range to examine its effect on the model
predictions. Table 4 compares five cases where the
reaction rate coefficients, kl to k3, were varied from
0.025 to 5.0 times the values predicted by the equations given in Appendix A.
The predicted methane conversion for these five
cases indicates that the catalyst activity has a limited
effect on the model predictions, within the range of
activity of commercial catalysts for the catalyst inventory investigated. When the catalyst activity was reduced by 50%, the conversion decreased only from
0.808 to 0.804. Larger changes were predicted when
the activity was varied over a wider range. The
changes in methane conversion shown in Table 4 are
due to the interaction between the chemical reaction
and the mass exchange between the two phases (Adris,
1994), i.e. both thermodynamics and interphase mass
transfer play important roles.
3.2.4. Model sensitivity to bubble size estimation. The Mori and Wen (1975) equation (see Appendix B) for estimating the bubble diameter, like other
widely used relations for bubble diameter estimation,
was developed for beds without internals. In the experimental reactor system (Adris et al., 1994b), however, membrane tubes are placed vertically in the
reactor and the bubble diameters could well differ
from the estimated ones. In this section the bubble
diameter predicted by the Mori and Wen equation
was multiplied by factors of 0.4, 0.6, 0.8, 1.0 and 1.2 to
study the impact of bubble size on the conversion

predictions. Table 5 gives the predicted values for


methane conversion in each case. The bubble size is
seen to have little influence on the reaction conversion
for the conditions investigated.
3.3. Gas distribution between phases
Accounting for a change in the number of moles of
gas due to chemical reaction complicates fluidizedbed reactor modeling (Irani et al., 1980; Kai and
Furusaki, 1987). This is an important consideration
for steam methane reforming where, as seen from eqs
(A1) and (A3) in Appendix A, the reactions cause
a significant increase in molar gas flow. A hydrodynamic investigation of the distribution of additional gas flow resulting from the increase in the total
number of moles due to reaction (Adris et al., 1993)
showed that at least some of the additional flow ends
up in the bubble phase. It is not clear, however,
whether the additional moles generated stay in the
dense phase for some distance or transfer quickly to
the bubble phase.
The model outlined in the present study assumes
that the gas flows through both the bubble phase and
the dense phase change with height due to the increase
in the total molar flow rate caused by the reforming
reactions. Since the reactions occur mostly in the
dense phase, most of the additional gas volume is
generated there. The extra moles are distributed between the two phases by interphase mass transfer and,
possibly also, by bulk transfer of excess moles generated to the bubble phase, with some of the extra moles
also removed due to permeation.
The change in the gas flow through the bubble and
dense phases along the bed height is plotted in Fig. 3
for three different situations, calculated based on the
model presented in this work. The plot shows that the
gas flow through both phases increases with height
when there is no permeation, with the dense phase
share of the excess gas being about 6 5 0 . An increase
followed by a slight decrease is exhibited for reaction-permeation with a limited permeation capacity
(Cep = 1.0 km); in this case the dense phase accommodates about 5 0 0 of the excess gas, the bubble

Table 4. Model sensitivity to catalyst activity level: T = 800C, P = 1.5MPa,


Fc = 60 mol/h, S/C = 3.5
Rate coefficients,kt to k3,
multiplied by

0.025

0.1

0.5

1.0

5.0

Methane conversion

0.731

0.786

0.804

0.808

0.817

Table 5. Model sensitivity to estimated bubble size: T = 700C, P = 1.5 MPa,


Fc = 60 mol/h, S/C = 3.5
Bubble diameter factor

0.4

0.6

0.8

1.0

1.2

Methane conversion

0.579

0.579

0.577

0.570

0.559

1615

Modelling of a fluidized-bed membrane reactor


1.0

0.0

0.1

0.2

0.3

zx

0.4

a4

0.9-

2.O E

~
e"
tJ~

0.8 ~

R
ia.

" 0.7Q.

1.8 "~

(/)

J~
--I

"1o

0.6-

r'-

'

0.4-

J=
'-I

0.5-

0.3-

1.6

No permeation

ii=

---W-- Cep=1.0 km

1.4

Cep=l 0 km
0.2
0.0

0.'1

.o

0.'2
0.'3
Height coordinate, m

0.4

Fig. 3. Change of gas flow through bubble and dense phases along the reactor for no permeation and for
two different permeation capacities: T = 800C, P = 1.5 MPa, Fc = 80 mol/h, S/C = 3.5, Ps = 0.4 MPa,
Fs = 80 mol/h.

Table 6. Model predictions for reforming reaction with and without permeation considering variable gas
flow in both phases and constant dense-phase gas flow assumptions: Fc = 80 mol/h, S/C = 3.5, T = 800C,
P = 1.5 MPa, Ps = 0.4 MPa, Fs = 80 mol/h (freeboard reaction ignored)
Permeation capacity, km

Model assumption
Methane conversion
Steam conversion

10

1.0

Variable gas
flow in both
phases

Constant
dense phase
gas flow

Variable gas
flow in both
phases

Constant
densephase
gas flow

0.774
0.323

0.752
0.314

0.803
0.337

0.781
0.328

phase about 35%, with the balance being removed as


hydrogen by the membrane tubes. When the permeation capacity is increased to 10 km, the gas flow
steeply decreases after an initial increase, with most of
the excess gas removed by hydrogen permeation
through the membrane tubes.
In a variant of the model presented above, the gas
flow through the dense phase was maintained constant, i.e. Ud was maintained constant, so that all
the excess gas generated in the dense phase and
not removed by the permeable membrane tubes was
accommodated by the bubble phase. An adjustment
was carried out between each step in the integration.
Table 6 gives the model predictions for two cases
considering this 'constant dense phase gas flow'
assumption compared with the predictions of the
original model.
It is clear from Table 6 that the division of excess
gas distribution between the bubble and dense phases
can affect the model predictions to a small but appreciable extent. Given the available experimental

results which are dominated by reaction equilibrium


and permeability, it would be difficult, if not impossible, to discriminate between the two models based on
reaction experiments. Separate experiments designed
to determine the gas flow in the separate phases are
required to discriminate between these two rival assumptions, i.e. to determine how extra moles of gas
generated by the reactions are distributed between the
two phases.
4. P A R A M E T R I C

INVESTIGATION

A parametric investigation was conducted to obtain better insights into the effect of major operating
variables and design parameters influencing the performance of the reactor. In addition, the performance
of the FBMR system was explored beyond the range
of parameters which could be studied experimentally
due to limitations imposed by economic and safety
considerations.
The reaction-permeation option has been used for
the predictions in this section, with the contribution of

1616

A.M. Adris et al.

the freeboard ignored and the temperature assumed


to be uniform, except where explicitly specified otherwise. The model with variable gas flow in both phases
is used. Variables and parameters studied and their
ranges are as follows: temperature: 400-800C; pressure: 0.3-2.7 MPa; steam-to-carbon molar feed ratio:
1.5-5.5; methane molar feed rate: 20-100 mol/h; permeation capacity, Cep: 0.4-7.0 km; sweep gas flow
rate: 40-120mol/h; separation side pressure:
0.1-0.9 MPa.
The investigation focused on the effect of changing
these parameters and variables on the conversion of
methane. Another parameter, the fraction of hydrogen separated from the reaction domain through the
membrane tubes, is also useful here in indicating the
extent of the equilibrium shift. This parameter is responsible for altering the composition of the reacting
mixture by changing its hydrogen content. The extent
of this change depends on the portion of the hydrogen
produced which is removed from the reaction system
by permeation. It is referred to here as the 'hydrogen
fraction separated'.

1.00 -

e.o,^
~
~>
oo
t t~
t-.,
~

0.90"
0.a50.a0.

~7

Q
0
~7

0.75.
0.70-

Equilibrium
Cep=0.4 km

Cep=3.0 km

~7

Cep=7.0km

o'.5
1'.o
1'.5
21o
2'.5
Reactor pressure, MPa

0.0

3.0

Fig. 4. Effect of reactor pressure on the methane conversion


for different permeation capacities: T = 800C, Fc =
60 mol/h, S/C = 3.5, Ps = 0.4 MPa, Fs = 80 mol/h.
1.0

0.50 .~

O
C 0.8-

._o

4.1. Effect of operating variables


4.1.1. Operating pressure. In conventional reformers, higher reforming reaction conversions are
favoured by lower reactor pressures, an obvious consequence of LeChatelier's principle. In an FBMR
system, the SMR reaction is accompanied by a separation process which is favoured by higher pressures.
Therefore, the net effect of increasing the reactor pressure depends on the balance between the reaction and
permeation processes. Factors like the permeation
capacity, temperature and steam-to-carbon ratio all
have significant impacts on this balance; the net effect
could be a decrease, no change or an increase in
methane conversion.
Fifteen cases were simulated here to predict the
effect of pressure over the range 0.3 to 2.7 MPa at
three different membrane equivalent permeation capacities, 0.4, 3.0 and 7.0 km. Simulation results together with the corresponding equilibrium conversions at the reactor-bed conditions for each operating
pressure are plotted in Fig. 4. The conversions in Fig.
4 indicate the two opposing effects which the reactor
pressure has on the overall conversion. At low pressure, the methane conversion is lower than its equilibrium value due to the insignificant extent of hydrogen
separation, as well as due to strong bubble by-passing
caused by the high superficial gas velocity. As the
pressure increases, the impact of the permeation capacity becomes more important and the extent of the
equilibrium shift is determined by the membrane capacity. The cases shown in Fig. 4 with high permeation capacity (Cep = 7.0 km) exhibit the strong
potential of the F B M R for SMR reactions. The 0.96
methane conversion obtained here at 800C and
2.7 M P a exceeds the equilibrium conversion by about
25%. Such a high conversion could only be obtained
at a much higher operating temperature, about 130C
higher, in a conventional reforming reactor.

0.95-

Equilibrium conversion

FBMR conversion

Hydrogen fraction separated

>~ 0.5C

~ 0.4.

.2
o.,~
e.-

g
~ 0.2.
0.0

0.45 ~ "

0.35

"1-

4~o

5~o
0~o
r~0
Reactor temperature, C

800

0.30

Fig. 5. Effect of reactor temperature on the methane conversion and separation by permeation: P = 1.5 MPa, Fc =
60 mol/h S/C=3.5, Ps = 0.3 MPa, Fs = 80 mol/h, Cep= 7.0 km.
4.1.2. Operating temperature. Increasing the operating temperature has a positive effect on both the
reaction and the permeation processes, since both
depend on temperature in an Arrhenius fashion. Both
the methane conversion and the hydrogen fraction
separated by permeation always increase with temperature as shown in Fig. 5. The increase in membrane permeability with increasing temperature does
not necessarily match the corresponding increase in
reaction rate and equilibrium constant. Therefore, the
extent of the equilibrium shift varies with temperature
according to the temperature dependence of the permeability rate constant. The membrane capacity
again contributes to the extent to which the reaction is
shifted from the thermodynamic equilibrium. The reactor operating temperature is limited in practice by
the membrane tube lifetime and depends on the exit
gas composition requirements for downstream processing.
4.1.3. Steam-to-carbon molar feed ratio (S/C). The
steam-to-carbon molar feed ratio is another variable
with two opposing effects on the FBMR. Increasing

1617

Modelling of a fluidized-bed membrane reactor


2

3
i

0.8-

0.5-

5
t

0.95 -

t .oo

0.6

Equilibrium
With permeation

No permeation

" ~ 0.90O~
0.4-

0.4

0.85-

o
0.3-

0.3

4'o

6'0

8'0

1~o

0.80t-

0
o.7s0.2

0.2

Steam-to-carbon m o l a r feed ratio

Fig. 6. Effect of steam-to-carbon molar feed ratio at constant methane flow on the reaction conversion and hydrogen
separation: T = 600C, P = 1.5 MPa, Fc = 60 mol/h, Ps =
0.3 MPa, Fs = 80 mol/h, Cep = 3.0 km.

the S/C ratio increases the reaction conversion thermodynamically. However, a higher steam concentration in the reacting mixture reduces the hydrogen
concentration, thereby diminishing the driving force
for permeation. The choice of S/C is constrained at its
lower and upper limits by carbon formation and catalyst re-oxidation, respectively. In addition, palladium
has a tendency to form oxides in an oxidizing atmosphere (Tsotsis et al., 1993), and high steam-to-carbon
ratios may affect the tube life.
In the simulation, the steam-to-carbon molar feed
ratio was changed from 1.5 to 5.5 by increasing the
flow of steam at a constant methane flow. The resulting methane conversion and hydrogen fraction separated for these two cases are presented in Fig. 6. The
net effect is an overall increase in methane conversion.
However, a strong negative effect on the selective
separation process is exhibited in Fig. 6 by the steep
decrease in the hydrogen fraction separated, from 0.52
at S/C = 1.5 to 0.27 at S/C = 5.5, indicating a reduction in the extent of the equilibrium shift.
4.1.4. Reactor throughput. Increasing the reactor
throughput (i.e. overall gas flow rate) affects the performance of the FBMR negatively: (a) by increasing
bubble by-passing and (b) by reducing the magnitude
of the equilibrium shift due to a reduction in the
fraction of hydrogen separated by permeation at the
same membrane capacity. The effect of this variable
was examined by performing five simulations where
the methane feed rate was increased from 20 to
100 mol/h at a constant steam-to-carbon ratio. Results are shown in Fig. 7 confirming the anticipated
negative effect. The predicted methane conversion
without membrane separation is plotted on the same
graph for comparison. The gap between the conversion with and without permeation is diminished as the
methane flow rate increases, due to the decline in the
role played by membrane separation in shifting the
equilibrium. At the highest methane flow the FBMR
gives a lower conversion than the equilibrium value,

0.70

2'0

M e t h a n e flow, mol/h

Fig. 7. Effect of reactor throughput on methane conversion:


T = 800C, P = 1.5 MPa, Ps = 0.4 MPa, Fs = 80 mol/h,
S/C = 3.5, Cep = 3.0 km.

04s

40

60

80

.~ 0.44-

100

120
i

0.38

0.36

a.

e-

0.32 .o_
~
m

0.30 4::
r-

0.40

0.34

043.

..
~ 0.42,~
0.41-

"0

4'0

;o

8'0

Methane conversion
Hydrogen separation

1~o

1~o

0.28 L~
"10
0.26 '3"
0.24

Sweep gas flow rate, mol/h

Fig. 8. Effect of sweep gas flow on the reaction conversion


and hydrogen separation: T = 600C, P = 1.5 MPa, Fc =
60 mol/h, S/C = 3.5, Ps = 0.3 MPa, Cep = 3.0 km.

indicating that the membrane separation is insufficient to compensate for bubble by-passing.
4.2. Effect of membrane-side design parameters
The effect of the sweep gas flow rate was studied by
changing its value from 40 to 120 mol/h. The predicted methane conversions and hydrogen fractions
separated are plotted against sweep gas flow in Fig. 8.
The increase in sweep gas flow reduces the hydrogen
partial pressure on the separation side, thereby increasing the permeation driving force and leading to
higher rates of hydrogen removal from the reaction
domain. The extent of the increase in conversion is,
however, relatively small, with a tripling of the sweep
gas flow only increasing the conversion from about
0.415 to 0.434.
The separation side pressure was varied from 0.1 to
0.9 MPa to explore its influence. The predictions are
plotted in Fig. 9 showing that an increase in the
separation side pressure causes a decrease in conversion. This is because an increase in the hydrogen
partial pressure on the separation side reduces the

1618

A. M .

0.91

0.8

0.9

was also examined. As expected, the overall conversion increases significantly as the membrane capacity
increases. The effect is not limited by other factors in
the model. Results are plotted in Fig. 10. However, in
practice tubes must be separated by a distance of at
least 20 dp to 30 dp to maintain good fluidization
(Grace, 1982), and this puts an upper limit on the
number of tubes. Also the wall thickness must be
sufficient to allow the tubes to withstand mechanical
forces and erosion in the bed. Like the other two
separation side parameters discussed above, Ce~ must
be optimized on an economic basis, considering the
cost of the membrane material together with the cost
of generating and recovering the sweep gas, most
likely steam.

0.5~

") 0.89

Adris et al.

8 o.~

0.4

[]

I0.87

r-I~

0.3 }

0.86

o12

0.85

o14

o18

Separation side pressure,

0.2

oi.
MPa

Fig. 9. Effect of separation side pressure on the reaction


conversion and hydrogen separation: T = 800C, P =
1.5 MPa, Fc = 60 mol/h, SIC = 3.5, Fs = 80 mol/h, Cev =
3.0 km.

2
i

1.1111

4
i

6
i

1.0

0.95.
tO
'~

0.80.

0.85t-

4.3. Effect of membrane separation on freeboard reactions


One of the important advantages of the FBMR
system for reversible reactions is its ability to suppress
undesirable reactions in the freeboard region by shifting the thermodynamic equilibrium by product removal mainly in the dense bed. This property of the
FBMR is demonstrated here by four simulation runs
where, for illustration purposes, the freeboard was
assumed to be cooler than the average bed temperature by 75C, while the mass of entrained catalyst
was assumed to be 4 times that estimated for our
reaction-permeation experiments. The first run was
performed without hydrogen separation; in the second and third, hydrogen was selectively removed using two different membrane capacities, while the
fourth had membrane tubes extending into the freeboard region. Table 7 gives the conditions and predicted results.
Table 7 shows that freeboard reactions reduce the
methane conversion in the first case by about 4%,
with the resulting exit conversion falling between the
equilibrium conversion at bed conditions and that
corresponding to the freeboard conditions. In the
second case, methane conversion at the bed surface
approaches equilibrium by virtue of membrane separation, with freeboard reactions affecting the conversion in a negative way, but to a lesser extent than for
the first case. When the membrane capacity is increased in the third column, the reaction conversion
exceeds the equilibrium limits in the bed and very

O.S ~
0.6

e-.

0.4 ~

O.SO.

o.rs-

0
070

Methane conversion
Hydrogen sepamlJon

0.2 ~
"r
oo

Equivalent permeation capacity, Cop, km


Fig. 10. Effectofmembranecapacity on the reaction conversion and hydrogen separation: T = 800C, P = 1.5 MPa,
Fc = 60 mol/h, SIC = 3.5, Ps = 0.3 MPa, Fs = 80 mol/h.

permeation driving force. The extent of this negative


effect is again limited, with the ninefold increase in the
separation side pressure only reducing the methane
conversion by about 4%. It should be remembered,
however, that higher membrane side pressures enable
the use of thinner membrane tube walls, thereby enhancing permeation.
The effect of the capacity of the membrane tubes,
expressed as the equivalent permeation capacity, Cep,

Table 7. Predicted effect of membrane separation on limiting reaction reversal in the freeboard:
P = 1.5 MPa, TR = 800C, T1b = 725C, Fc = 80 mol/h, S/C = 3.5, Fs = 80 mol/h, Ps = 0.4 MPa,
mass of catalyst entrained in the freeboard = 0.03 kg
Permeation capacity in dense bed, km
Permeation capacity in dilute phase, km
Methane conversion at bed surface
Equilibrium conversion at bed conditions
Methane conversion at reactor exit
Equilibrium conversion at freeboard
conditions

2.80

6.73

2.80

0
0.774
0.831
0.735
0.647

0
0.822
0.831
0.798
0.647

0
0.857
0.831
0.848
0.647

1.2
0.822
0.831
0.819
0.647

Modelling of a fluidized-bed membrane reactor


little reverse reaction is predicted in the freeboard
region. Indeed, the overall conversion remained
higher than the equilibrium conversion for the bed
conditions and much higher than the equilibrium
conversion at the freeboard conditions.
It is clear that removal of product hydrogen from
the dense bed by means of permeable membranes can
significantly reduce adverse freeboard effects. Further
membrane separation in the freeboard region could
also be helpful. In the final case, the membrane capacity in the dense catalyst bed is the same as in the
second case, but Cep is augmented by a further 1.2 km
of membrane in the freeboard. This addition leads to
almost no reduction in methane conversion in the
freeboard, despite the reduced temperature there. For
the conditions explored here, membrane surfaces in
the freeboard zone do not contribute to the net thermodynamic equilibrium shift and therefore cannot
lead to methane conversions higher than at the bed
surface. They do, however, cause a reduction in the
reverse reaction. The distribution of membrane capacity between the bed and the freeboard is important in
designing a fluidized-bed membrane reactor, with the
optimum distribution depending on operating conditions and catalyst properties, as well as on what is
downstream of the reforming process.
5. CONCLUSIONS
A comprehensive mathematical model has been
developed for the fluidized-bed membrane reactor
(FBMR) system and successfully validated against
pilot-plant data for steam methane reforming (SMR).
The model is utilized to examine the effect of operating variables beyond those which could be realized
experimentally. The predictions demonstrate the role
played by membrane separation in exceeding equilibrium conversions and in suppressing freeboard reactions. The membrane capacity and their distribution
between the dense catalyst bed and the freeboard
region are important design parameters in FBMR
systems.
Hydrodynamic parameters have only a limited effect on the model predictions for the conditions explored because of the dominance of reaction thermodynamics and selective permeation as the two major
phenomena influencing conversion within the FBMR
system. More experimental measurements are required to discriminate between two rival assumptions
explored in this work regarding the distribution of gas
flow generated by the increase in the number of moles
due to reaction,

ab
ac

NOTATION
specific surface area of gas bubbles,
m2/m 3
overall entrainment decay constant,

m- 1
A
Ci

reactor cross-sectional area, m 2


concentration of component i, mol/
m3

Cep

1619

equivalent permeation capacity: membrane surface/wall thickness, m2/m


equivalent permeation capacity per
Cepl
unit length, m2/m 2
bubble size, a function of height dedb
fined by eq. (B1), m
maximum bubble diameter, m
db,,,
initial bubble size produced at the disdbo
tributor level, m
mean particle diameter, m
dp
reactor diameter, m
D
denominator of the kinetic rate exDeN
pression defined by eq. (A8)
effective molecular diffusivity of comDie
ponent i, m2/s
entrainment flux of solids at the bed
Eo
surface, kg/m 2 s
entrainment flux of solids at distance
Ez
z above the bed surface, kg/m 2 s
fluidized-bed membrane reactor
FBMR
molar feed rate of methane equivaFc
lent, mol/s
molar flow rate of sweep gas, mol/s
Fs
#
acceleration of gravity, m/s 2
volumetric flow, m3/s
G
vertical coordinate measured from
h
distributor plate, m
expanded bed height, m
H
static bed height, m
Ho
heat of reaction at 298 K, J/mol
AH298
rate coefficients of reactions (A1) and
kl, ka
(A3), respectively, mol MPaS/kgc,t s
rate coefficient of reaction (A2),
k2
mol/kgc,t s MPa
equilibrium constant for reactions
K1, K3
(A1) and (A3), MPa 2
equilibrium constant for reaction (A2)
K2
dimensionless
kcH,, kco, kH2 adsorption constants for CH4, CO
and H2, respectively, MPa-1
dissociative adsorption constant for
kmo
H20, dimensionless
interphase mass exchange coefficient
klq
for component i, m/s
molecular weight of hydrogen, kg/mol
Mn
molar flow rate of component i, mol/s
ni
number of orifices in the grid
Nor
reactor total pressure, MPa
P
Pl-lb,Pna, Pm partial pressure of hydrogen in the
bubble, dense and separation phase,
respectively, MPa
p~
total pressure on the separation side,
MPa
Pi
partial pressure of component i, MPa
ideal gas constant, 8.314 J/mol K
R
RR1, RR2, RR3 rates of reactions (A1), (A2) and (A3)
respectively, mol/kgcat s
Ri
rate of formation of component i,
mol/kg~,t s
steam-to-carbon molar feed ratio, diS/C
mensionless

A. M. Adris et al.

1620
SMR
T
TDH

Ub
Ud

uo
U~s

Greek letters
~mf
~b
~)b, ~)d

(I) e

Pn
Ps

Subscripts
b
d
f
fb
i
m
mf
o
R
s
z

steam methane reforming


temperature, C
transport disengaging height, m
bubble rising velocity, m/s
superficial gas velocity through the
dense phase, m/s
superficial gas velocity, m/s
superficial gas velocity at minimum
fluidization, m/s
freeboard zone vertical coordinate, m

bed voidage at minimum fluidization


volume fraction of bed occupied by
bubbles
volume fraction of bed occupied by
solids in bubble and dense phases, respectively
effective permeation rate constant,
m2/s MPa -5
hydrogen density, kg/m 3
particle density, kg/m 3
solids concentration in the freeboard
zone, kg/m 3

bubble phase
dense phase
feed
freeboard
integer (1-5) denoting gaseous components
maximum
minimum fluidization
at distributor plate
reaction side
separation side
at height z in freeboard zone
REFERENCES

Adris, A. M. (1994) A fluidized bed membrane reactor


for steam methane reforming: experimental verification and model validation. Ph.D. dissertation,
Univ. of British Columbia, Vancouver, Canada.
Adris, A. M., Elnashaie, S. S. E. H. and Hughes, R.
(1991) A fluidized bed membrane reactor for the
steam reforming of methane. Canad. J. Chem.
Engng 69, 1061-1070.
Adris, A. M., Lim, C. J. and Grace, J. R. (1993). The
effect and implications of gas volume increase
due to reaction on bed expansion, bubbling and
overall conversion in a fluidized bed reactor.
A.I.Ch.E. Annual Meeting, St. Louis, November 7-12, 1993.
Adris, A. M., Grace, J. R., Lim, C. J. and Elnashaie,
S. S. E. H. (1994a) Fluidized bed reaction system for
steam/hydrocarbon gas reforming to produce hydrogen. U.S. Patent No. 5,326,550.
Adris, A. M., Lim, C. J. and Grace, J. R. (1994b) The
fluidized bed membrane reactor (FBMR) system:
a pilot scale experimental study. Chem. Engng Sci.
49, 5833-5843.

Davidson, J. F. and Harrison, D. (1963) Fluidised


Particles. Cambridge University Press, Cambridge,
U.K.
Elnashaie, S. S. E. H., Adris, A. M., A1-Ubaid, A. S.
and Soliman, M. A. (1990) On the non-monotonic
behaviour of methane steam reforming kinetics.
Chem. Engng Sci. 45, 491-501.
Geldart, D. (1986) Particle entrainment and carryover. In Gas Fluidization Technology, ed. D. Geldart, Chap. 6, pp. 123-153. Wiley, Chichester, U.K.
Grace, J. R. (1982) Fluidized bed hydrodynamics. Section 8.1 in Handbook of Multiphase Systems, ed. G.
Hetsroni, pp. 8-5 to 8-64. Hemisphere, Washington.
Grace, J. R. (1984) Generalized models for isothermal
fluidized bed reactors. In Recent Advances in the
Engineering Analysis of Chemically Reacting System, ed. L. K . Doraiswamy, pp. 237-255. Wiley
Eastern, New Delhi, India.
Grace, J. R. (1986) Fluid beds as chemical reactors. In
Gas Fluidization Technology, ed. D. Geldart, pp.
285-339. Wiley, Chichester, U.K.
Irani, R. K., Kulkarni, B. D. and Doraiswamy, L. K.
(1980) Analysis of fluid bed reactors for reaction
involving change in volume. Ind. Engng Chem. Fundam. 19, 424-428.
Kai, T. and Furusaki, S. (1987) Methanation of carbon dioxide and fiuidization quality in a fluid bed
reactor - - the influence of decrease in gas volume.
Chem. Engng Sci. 42, 335-339.
Katsuta, H., Farraro, R. J. and McLeUan, R. B. (1979)
The diffusivity of hydrogen in palladium. Acta Metallurgica 27, 1111-1114.
Kunii, D. and Levenspiel, O. (1969) Fluidization Engineering, Wiley, New York, U.S.A.
Kunii, D. and Levenspiel, O. (1991) Fluidization Engineering, 2nd Ed. Butterworth-Heinmann, MA,
U.S.A.
Miwa, K., Mori, S., Kato, T. and Muchi, I. (1972)
Behaviour of bubbles in gaseous fluidized beds. Int.
Chem. Engng 12, 187-194.
Mori, S. and Wen, C. Y. (1975) Estimation of bubble
diameter in gaseous fluidized beds. A.1.Ch.E.J. 21,
109-115.
Orcutt, J. C., Davidson, J. F. and Pigford, R. L. (1962)
Reaction time distributions in fluidized catalytic
reactors. Chem. Engng Prog. Syrup. Ser. 85, 1-15.
Rostrup-Nielsen, J. R. (1984) Catalytic steam reforming. In Catalysis Science and Technology, eds. J. R.
Anderson and M. Boudart, Vol. 4. Springer, Berlin,
Germany.
Sit, S. P. and Grace, J. R. (1981) Effect of bubble
interaction on interphase mass transfer in gasfluidized beds. Chem. Engng Sci. 36, 327-335.
Soliman, M. A., Adris, A. M., Elnashie, S. S. E. H and
A1-Ubaid, A. S. (1992) Intrinsic kinetics of
nickel/calcium aluminate catalyst for methane
steam reforming. J. Chem. Technol. Biotechnol. 55,
131-138.
Tsotsis, T. T., Champagniem, A. M., Minet, R. G. and
Liu, P. K. T. (1993) Catalytic membrane reactors.
In Computer Aided Design of Catalysts, Chemical
Industries, Vol. 51, eds. E. Rober Becker and
Caruso, J. Perira, pp. 471-551. Marcel Dekker,
New York, U.S.A.
Wen, C. Y. and Chen, L. H. (1982) Fluidized bed
freeboard phenomena--Entrainment and elutriation. A.I.Ch.E.J. 28, 117-128.

1621

Modelling of a fluidized-bed membrane reactor


Wilke, C. R. a n d Lee, C. Y. (1995) Estimation of
diffusion coefficients for gases and vapors. Ind.
Engn# Chem. 47, 1253-1257.
Xu, J. and F r o m e n t , G. F. (1989) M e t h a n e steam
reforming, m e t h a n a t i o n and w a t e r - g a s shift: I. Intrinsic kinetics. A . I . C h . E . J . 35, 88-96.

Values of the various constants are given in Table A1.


Equilibrium constants:
K1 = exp (-26,830/T + 30.114)
K2 = exp (4,400/T -- 4.036)

K3 = K1K2

A P P E N D I X A: R E A C T I O N R A T E E Q U A T I O N S A N D
PARAMETERS

The three main reactions taking place in steam reforming


are as follows:
CH4 + H20 = CO + 3H2;
-AH298 = - 206.0 kJ/mol CH4

A P P E N D I X B: H Y D R O D Y N A M I C

PARAMETERS

The model uses the Mori and Wen (1975) correlation


which accounts for bubble growth due to coalescence to
estimate bubble size as a function of height, i.e.

(A1)

db= dbm -- (db,, -- dbo)e -'3h/

(B1)

CO + H 2 0 = C O 2 + H2;
-

AH298 = 41.0 kJ/mol CO

(A2)

where db,, is the maximum bubble diameter at the given gas


flow rate given by

CH4 + 2H20 = CO2 + 4H2;

db,,= 1.64[A(Uo -- U,,I)] '4


--AH298 = -- 164.9 kJ/mol CH 4.

(B2)

(A3)

The net rate of formation of the various components are

and dbo is the initial bubble size produced at the distributor


level which can be estimated (Miwa et al., 1972) as

RcH, = - (RR1 + Ra3)


1.38 FA(Uo -- U,s)l'4
dbo=~L
~
j
.

Rn2o = -- (RR1 + RR2 + 2RR3)

(B3)

Rco = RR1 -- RR2


Rco2 = RR2 + RR3
Rn: = 3RR1 + RR2 + 4RR3

(A4)

where RR1,RR2 and RR3 are the intrinsic rates for reactions
(AI), (A2) and (A3), respectively. The reaction rate expressions developed by Xu and Froment (1989) are used to
estimate the individual reaction rates. The functional forms
of these expressions are
{~ocH.PHzo pO.Sp

RRI = k l l
\

\ /
H-~lCO)/D2N

p2.5

(PcoPH2o
RR2 = k2 \ Pn2
I
RR3:k3~.-~2

(AS)

H2

Pco2\ / 2

~')/DEN

(A6)

"[Pcu'P22 pO.Sp
H2 CO2\ / 2
K1K------2-)/DEN

(i7)

where

[Pn~o\

DEu = 1 + kcoPco + kn2Pu2 + ken, Pert, + kn2o - z - - - | .


/

[rH~

Here Nor is the number of orifices in the grid, with No~ = 116
in the pilot-scale reactor used in the present study. The above
relations for estimating the bubble size were developed for
freely bubbling beds, i.e. without internals. The presence of
vertical internals in the bed may significantly affect the
bubble size. The sensitivity of the model predictions to
bubble size estimations is explored in this paper.
Because of the net increase in the total number of moles
due to reaction, the superficial gas velocity, Uo, changes with
height. Equations (B1) and (B2) are therefore re-used at each
step in height to recalculate db. The model uses experimentally measured values for the expanded bed height, H, as
discussed in an earlier paper (Adris et al., 1993). Because the
bubble diameter increases with height, the ratio, ab, of
bubble surface area per volume, given by the relation,
ab = 6/db, decreases with height.
A mean value is used for the volume fraction of bed
occupied by solids dispersed in the bubble phase, ~b. Kunii
and Levenspiel (1969) found that 0.001eb < ~b <0.01eb,
where eb is the volume fraction of the bed occupied by
bubbles, which can be estimated by

(A8)

eb = (Uo - U,.f)/Ub.

Table A1. Rate expression parameters and equilibrium constants due


to Xu and Froment (1989) used in this work

Constants

Pre-exponential factors

kl
k2
k3
kco
kcm
kn2o
ka2

9.49 1015
4.39 106
2.29 x 1015
8.23 x 10- 5
6.65 x 10- s
1.77 x 105
6.12 10 -9

* Units are given in the Notation.

Activation energies and


heats of absorption*
240.1
67.13
243.9
70.65
- 38.28
88.68
-- 82.9

(B4)

A. M. Adris et al.

1622

Here the bubble rising velocity is calculated by the commonly used relation (Davidson and Harrison, 1963):
Ub = Uo -- Umf + 0.711(g.db) 1/2.

(B5)

The volume fraction occupied by the dense phase solids, Od,


is approximated by including all of the solids in the clouds,
wakes and emulsion in the dense phase and assuming the
dense-phase voidage to be constant and equal to emI" Thus,
O~ can be estimated (Grace, 1986) as
Od = (1 - eb)(1 -- t~I)'

(B6)

The interphase mass exchange coefficient for component i,


ki~, is calculated using the semi-empirical equation of Sit and
Grace (1981):
Umy

k,q = ~

['4Diee,mIVb~ 1/2

+ ~ - - )

(B7)

where D~e, the effective molecular diffusivity of component


i in the gas mixture, is estimated based on the average
composition of the bubble phase and dense phase [i.e.
(Cib + Cid)/2] using a relation given by Wilke and Lee (1995).

You might also like