You are on page 1of 6

Energy & Fuels 2007, 21, 35933598

3593

Sequential Simulation of a Fluidized Bed Membrane Reactor for the


Steam Methane Reforming Using ASPEN PLUS
A. Sarvar-Amini, R. Sotudeh-Gharebagh,*,, H. Bashiri, N. Mostoufi, and A. Haghtalab
Process Design and Simulation Research Centre, School of Chemical Engineering, UniVersity of Tehran,
P.O. Box 11155-4563, Tehran, Iran, and Chemical Engineering Department, Faculty of Engineering,
Tarbiat Modares UniVersity, Tehran, Iran
ReceiVed June 22, 2007. ReVised Manuscript ReceiVed August 17, 2007

A simulation model is developed using ASPEN PLUS to predict the performance of a fluidized bed membrane
reformer. Because there are physical and chemical phenomena interacting in the fluidized bed membrane
reformer, two submodels seem necessary in the model. These submodels are the hydrodynamic and reaction
submodels. The hydrodynamic submodel is based on the dynamic two-phase model, and the reaction submodel
is derived from the literature. The reformer is divided into two regions: a dense bed and freeboard. The dense
bed is divided into several sections. At each section, the flow of the gas is considered as the plug flow through
the membrane and bubble phases and perfectly mixed through the emulsion phase. The sets of the experimental
data were used from the literature to validate the model. Close agreement was observed between the model
predictions and experimental data. This model can be used for the simulation of nonideal fluidized bed membrane
reactors inside the ASPEN PLUS process simulator.

Introduction
Steam reforming of hydrocarbons is the leading industrial
process for producing hydrogen and synthesis gases for ammonia and methanol production, hydrocracking and hydrotreating, FischerTropsch synthesis, and other important processes
in petroleum refining and petrochemical industries.1 Typical
industrial reformers are formed of hundreds of fixed bed tubes
packed with nickel catalyst particles within gas-fired furnaces.2
Industrial fixed-bed steam reformers have several drawbacks,
such as thermodynamic equilibrium limitations, carbon formation on the catalyst, a large temperature gradient, and low heattransfer rates, which seriously affect their operation and
performance. In recent years, new membrane-based fluidized
bed reformers have been under development to overcome these
shortcomings.
Considerable attention has been paid to the fluidized bed
membrane reformers (FBMRs) as multifunctional reformers.3
The main advantages of FBMR are a uniform temperature in
the bed, an increase in the hydrogen production because of
changes in thermodynamic equilibrium conditions, an enhancement of methane conversion, a simultaneous reaction and
separation of hydrogen, an elimination of intercatalyst diffusion
limitations, a good heat-transfer capability, and a more compact
* To whom correspondence should be addressed. Fax: (0098) 21-66461024. E-mail: sotudeh@ut.ac.ir.
University of Tehran.
On sabbatical leave (20062007) at the Chemical Engineering Department, Qatar University, Doha 2713, Qatar.
Tarbiat Modares University.
(1) Adris, A. M.; Lim, C. J.; Grace, J. R. The fluidized-bed membrane
reactor for steam methane reforming: Model verification and parametric
study. Chem. Eng. Sci. 1997, 52, 16091622.
(2) Chen, Z.; Prasad, P.; Yan, Y.; Elnashaie, S. S. E. H. Simulation for
steam reforming of natural gas with oxygen input in a novel membrane
reformer. Fuel Process. Technol. 2003, 83 (13), 235252.
(3) Abashar, M. E. E. Coupling of steam and dry reforming of methane
in catalytic fluidized bed membrane reactors. Int. J. Hydrogen Energy 2004,
29, 799808.

design.3 Despite these advantages, very few attempts have been


reported in the literature on the industrial commercialization of
FBMRs because of technological difficulties, complexity in the
scale up, and economics.4 In addition, there are some difficulties
with experiments to study the effects of operational conditions
on the performance of FBMRs, such as the controllability of
operating variables, proper maintaining of highly expensive
membrane tubes, complex operational concerns, and tedious
experiments involved. Furthermore, the planning and implementation of experiments, the analysis of the full array of
interactions between system components and bed materials,
internal surfaces, and measurement devices, and interpretation
of results to diagnose abnormal operating conditions present
operators with a range of difficult intellectual challenges. When
the reformer is modeled, these effects could be investigated
while saving time without a further need to generate highly
expensive data.
Modeling the FBMR as a nonideal system is useful to
investigate the effect of key operating parameters, optimization,
and scale up of such reformers. Because two types of phenomena, i.e., physical and chemical, coexist in FBMRs, two
submodels, describing these two phenomena, should simultaneously be included in the modeling: the hydrodynamic submodel, which explains the physical phenomena, and the reaction
submodel, which describes the chemical reactions occurring in
the reformer. Several approaches have been introduced in the
literature for modeling FBMRs that can be contributed to high
complexity in the hydrodynamic and nonideality of these
reformers. Adris et al.1 developed a two-phase bubbling bed
model for steam methane reforming with the gas permeation
via the membranes. They developed a model in which a plug
flow reactor (PFR) for both the bubble and dense phases is
(4) Deshmukh, S. A. R. K.; Heinrich, S.; Morl, L.; Vn Sint Annaland,
M.; Kuipers, J. A. N. Membrane assisted fluidized bed reactors: Potentials
and hurdles. Chem. Eng. Sci. 2007, 62, 416436.

10.1021/ef7003514 CCC: $37.00 2007 American Chemical Society


Published on Web 09/28/2007

3594 Energy & Fuels, Vol. 21, No. 6, 2007

SarVar-Amini et al.

Figure 2. Features of a dense bed in the fluidized bed membrane model.

Figure 1. Typical FBMR.

assumed. This model was later extended by Dogan et al.5 and


Rakib et al.6 for the autothermal steam reforming with oxygen
to investigate the effect of oxygen. Mleczko et al.7 developed
a bubble assemblage model based on the work of Kato et al.8
for the FBMR. In their model, emulsion and bubble phases are
considered to be composed of several continuously stirred tank
reactors (CSTRs). Chen et al.9 developed a steady-state onedimensional PFR model for the circulating FBMR.
Various process simulators, such as ASPEN PLUS and
HYSYS, are employed for industrial process simulation. These
process simulators only include standard, ideal reactors, such
as the PFR and CSTR. Despite the vast application of the
commercial simulator in the chemical industries, a fluidized bed
membrane reactor unit does not exist in the simulators. The
ideal reactors representing different hydrodynamic phases are
available in the commercial process simulators; therefore, it
would possible to introduce the nonideal fluidized bed membrane reactors in such simulators by the proper combination of
these ideal reactors with the external or internal subroutine. In
this work, the FBMR is modeled using the sequential modular
approach by dividing it into several sections in series. At each
section, the flow of gas is considered as a plug flow through
the bubbles and membrane and perfectly mixed through the
emulsion phase. The standard modules available in ASPEN
(5) Dogan, M.; Posarac, D.; Grace, J.; Adris, A. M.; Lim, C. J. Modeling
of autothermal steam methane reforming in a fluidized bed membrane
reactor. Int. J. Chem. Reactor Eng. 2003, 1, 112.
(6) Rakib, M. A.; Alhumaizi, K. I. Modeling of a fluidized bed
membrane reactor for the steam reforming of methane: Advantages of
oxygen addition for favourable hydrogen production. Energy Fuels 2005,
19 (5), 21292139.
(7) Mleczko, L.; Ostrowski, T.; Wurzel, T. A fluidized bed membrane
reactor for the catalytic partial oxidation of methane to synthesis gas. Chem.
Eng. Sci. 1996, 51, 3187.
(8) Kato, K.; Wen, C. Bubble assemblage model for fluidized bed
catalytic reactors. Chem. Eng. Sci. 1969, 24, 1351.
(9) Chen, Z.; Elnashaie, S. S. E. H. Steady-state modeling and bifurcation
behavior of circulating fluidized bed membrane reformerregenerator for
the production of hydrogen for fuel cells from heptane. Chem. Eng. Sci.
2004, 59 (18), 39653979.

Figure 3. Schematic simulation diagram of the reformer.

PLUS could be combined together to represent the real behavior


of the fluidized bed membrane reactors. Moreover, several
calculator blocks are provided along with the reactor modules
for the calculation of hydrodynamic parameters of the reformer.
Furthermore, the model has to be easily applicable in the ASPEN
PLUS process simulator.
Modeling
A schematic diagram of a typical FBMR is shown in Figure
1. As shown in this figure, methane is premixed with steam

Sequential Simulation of a FBMR

Energy & Fuels, Vol. 21, No. 6, 2007 3595


Table 1. Correlations and Equations Used in the Modeling

parameter
1

formula

Archimedes number
Ar )

minimum fluidization velocity

bubble velocity

bubble rise velocity

emulsion velocity

Fgdv3(Fp - Fg)g
2

Wen and Yu17

Rem ) [(33.7)2 + .0408Ar]0.5 - 33.7


Ub ) U0 - Ue + ubr
Kunii and Levenspiel
ubr ) 0.711(gdb)

U0 - Ub
1-

bubble mean diameter

0.3H
D

db ) dbm - (dbm - db0)exp -

)
Mori and Wen

bubble maximum diameter

bubble initial diameter

bubble-phase fraction for Geldart


B particles14

10

11

12
13

emulsion-phase porosity for Geldart


B particles14

18

dbm ) 1.64[A(U0 - Umf)]

0.4

1.38 A(U0 - Umf)


db0 ) 0.2
Nor
g

) 0.534 1 - exp -

bubble-phase porosity for Geldart


B particles14

b ) 1 - 0.146exp -

solid entrainment in the freeboard and


expansion zone
interphase mass-transfer coefficient

U0 - Umf
0.413

e ) mf + 0.2 - 0.059exp -

Miwa et al.16

)]

U0 - Umf
0.429

U0 - Umf
4.439

Cui et al.10

)
Kunii and Levenspiel

kiq )

4DiemfUb
umf
+
3
db

mole flow rate of component i


from the emulsion to bubble phase

Ni ) Askiq(Ci,emulsion - Ci,bubble)

15

mole flow rate of component i


from the bubble to emulsion phase

Ni ) Askiq(Ci,bubble - Ci,emulsion)

16

hydrogen permeation rate


QH2 ) Qeff
effective permeability constant of hydrogen

and fed into the bottom of the bed (lower region), where the
reforming reactions take place. Inside the bed, a number of
hydrogen perm-selective membrane tubes are inserted. Among
these membrane tubes, the catalyst is fluidized and the steam
reforming of methane takes place. The produced hydrogen
selectively permeates through hydrogen membranes and is then
carried away by sweep gas, such as steam in the hydrogen
membrane tubes.20 The reforming reactions are then completed
in the upper region (the freeboard and expansion zone) of the
reformer, where the entrainment and carry-over of catalyst
particles may have a considerable effect on reforming. The rate
of reforming in these regions is directly affected by hydrodynamics. Consequently, these two regions should be considered
in the modeling: the dense bed, which is operating under the
bubbling fluidization regime, and a more dilute upper region,

( )( )[
FH
MH

Am

13

0.5

14

17

13

12

Ue )
6

reference

Sit and Grace

15

PHR 2 - PHM2

Katsuta et al.19

Qeff ) pQ0exp(-Ep/RT)

i.e., the freeboard and expansion zone. The following is a


summary of the steps involved in the modeling of these regions:
Lower Region. The lower region of the reformer (dense bed)
consists of three phases: emulsion, bubble, and membrane
phases, as shown in Figure 2. To model the dense bed by
considering the complexity of hydrodynamics, the movement
of gas through the bubble and membrane phases are considered
as the plug flow, while the emulsion phase is assumed as the
completely mixed phase. The assumptions made in developing
the equations of the model for the dense bed are summarized
as follows: (1) The hydrodynamics of both phases are characterized by a dynamic two-phase theory of fluidization.10 (2)
Reactions occur in both bubble and emulsion phases only. (3)
(10) Cui, H. P.; Mostoufi, N.; Chaouki, J. Characterization of dynamic gas
solid distribution in the fluidized beds. Chem. Eng. J. 2000, 79, 135143.

3596 Energy & Fuels, Vol. 21, No. 6, 2007

SarVar-Amini et al.

Table 2. Descriptions of ASPEN PLUS Calculators Used in the Simulation


calculator name

purpose

Calc1

to calculate volumes, the voidage, and the flow rate for ideal reactors
(CSTR and PFR) in the bed
to calculate voidage in the freeboard
to calculate the average flow rate of components separated from the PFR
added to the membrane tube and CSTR
to calculate the average flow rate of components separated from the
CSTR and added to the membrane tube and PFR

Calc2
Calc3
Calc4

equations used
111 (Table 1)
12 (Table 1)
13 and 1517 (Table 1)
13, 14, 16, and 17 (Table 1)

The bubbles reach their equilibrium size quickly above the


distributor. (4) The temperature is assumed to be constant in
the dense bed. (5) Hydrogen is the only species transferred to
the membrane. (6) The plug flow is considered for sweep gas
and hydrogen in the membrane.
To develop a sequential modular model, the dense bed is
axially divided into several sections, where each section consists
of a membrane tube and two ideal reactors: a PFR to represent
the gas flow through the bubbles and a CSTR to represent the
gas flow through the emulsion. Following reactions taking place
in PFR and CSTR:11
CH4 + H2O S CO + 3H2

(1)

CO + H2O S CO2 + H2

(2)

CH4 + 2H2O S CO2 + 4H2O

(3)

As illustrated in Figure 3, the reactions take place at each


section in CSTR and PFR followed by the mass transfer among
the effluents of each section (bubble, emulsion, and the
membrane tube).
The ASPEN PLUS simulator was used as a platform to solve
the reaction and hydrodynamic submodels simultaneously. The
intrinsic capabilities of the simulator were used in developing
the model, i.e., at each section, the PFR and CSTR were used
to model the gas flow behavior through the bubble, membrane
and emulsion phases, respectively. Along with these ideal unit
operation blocks provided by ASPEN PLUS, calculator blocks
have also been introduced for the calculation of the voidage,
volume, and flow rate of gas through the different phases at
each section, as described in Table 2. Upon the determination
of hydrodynamic parameters for individual phases at each
section, ASPEN PLUS solves material and energy balance
equations and predicts physical and thermodynamic properties.
Effluents of the phases at the end of each section (i.e., unreacted
CH4, H2O, and reforming byproducts, such as CO, CO2, and
H2O) are then entered into the hierarchical block called MASS
TRANSFER, where the mass transfer among the phases are
calculated, as illustrated in Figure 4. As the mass-transfer
calculations are completed among the phases, the corresponding
flow distribution is then properly calculated to move to the next
section (i + 1). The calculation is continued upward until the
top of the bed is reached.
Upper Region. The upper region of the reformer consists of
two different zones, including the freeboard and expansion zone.
Entrainment of solid particles in these regions has a considerable
effect on the overall conversion of FBMR. The variation of the
(11) Kaihu, H.; Hughes, R. The kinetics of methanesteam reforming
over Ni/R-Al2O catalyst. Chem. Eng. J. 2001, 82, 311328.
(12) Adris, A. M.; Lim, C. J.; Grace, J. R. The fluidized bed membrane
reactor (FBMR) system: A pilot scale experimental study. Chem. Eng. Sci.
1994, 49, 58335843.
(13) Kunni, D.; Levenspiel, O. Fluidization Engineering, 2nd ed.;
Butterworth-Heineman: Boston, MA, 1991.
(14) Grace, J. R. Fluid beds as chemical reactors. In Gas Fluidization
Technology; Geldart, D., Ed.; Wiley: Chichester, U.K., 1986; pp 285
339.

Figure 4. Schematic diagram of the MASS TRANSFER block.

voidage with the height in the upper region of the reformer is


dependent upon the transport disengagement height (TDH) and
can be calculated according the procedure given by Kunni and
Levenspiel.13 The assumptions made in developing the model
for the upper region are summarized as follows: (1) PFR is used
for the modeling of the freeboard and expansion zone. (2)
Hydrogen is permeated to the membrane tube at the exit of the
reactor. (3) The amount of voidage in the freeboard is calculated
using a calculator block shown in Table 2. The calculation
procedure is the same as the lower region by considering two
phases in the calculations.
Results and Discussion
Results of the simulation model presented in this work are
compared to the experimental data reported by Adris12,21 in terms
of methane conversion, hydrogen yield, and hydrogen permeation through the membrane.
In Figure 5, the model prediction is compared to the
experimental conversion data at different temperatures for 15
sections. As seen in the figure, the methane conversion is
increased with an increase in the reformer temperature because
of the endothermic behavior of the reforming reactions (except
reaction number 2). Moreover, the permeability of hydrogen
(15) Sit, S. P.; Grace, J. R. Effect of bubble interaction on interphase
mass transfer in gas fluidized beds. Chem. Eng. Sci. 1981, 36, 327335.
(16) Miwa, K.; Mori, S.; Kato, T.; Muchi, I. Behaviour of bubbles in
gaseous fluidized beds. Chem. Eng. J. 1972, 12, 187194.
(17) Wen, C. Y.; Yu, Y. H. A generalized method for predicting the
minimum fluidization velocity. AIChE J. 1996, 12, 610612.
(18) Mori, S.; Wen, C. Y. Estimation of bubble diameter in gaseous
fluidized beds. AIChE J. 1975, 21, 109115.
(19) Katsuta, H.; Farraro, R. J.; McLeUan, R. B. The diffusivity of
hydrogen in palladium. Acta Metall. Sin. (Eng. Lett.) 1979, 27, 11111114.
(20) Chen, Z.; Yan, Y.; Elnashaie, S. S. E. H. Catalyst deactivation and
engineering control for steam reforming of higher hydrocarbons in a novel
membrane reformer. Chem. Eng. Sci. 2004, 59, 19651978.
(21) Adris, A. M. A fluidized bed membrane reactor for steam methane
reforming: Experimental verification and model validation. Ph.D. dissertation, University of British Columbia, Vancouver, Canada, 1994.

Sequential Simulation of a FBMR

Figure 5. FBMR conversion as a function of the temperature at different


numbers of sections (steam/carbon ratio ) 2.4; P ) 0.98 MPa; sweep
gas flow rate ) 80 mol/h; sweep gas pressure ) 0.4 MPa; methane
flow rate ) 74.2 mol/h).

through the membrane is increased by an increase in the


temperature, leading to a shift in thermodynamic equilibrium
and, consequently, to an increase in the methane conversion.
This figure also shows that by dividing the reformer into 45
sections, a close agreement between the predicted and experimental conversion can be found. It is worthy to mention that,
by further increasing the number of sections, the rate of the
mass transfer among the emulsion phase, bubble phase, and
membrane becomes close to the actual reformer. However, there
is a limitation in the number of sections, because the increase
in the number of sections alters the hydrodynamics of the emulsion phase from well mixed to plug flow. Under these conditions, the model overpredicts the experimental data as expected.
It is also important to mention that the number of sections is
not the same for other FBMRS and may be adjusted depending
upon the hydrodynamics and the reactions involved for a given
system.
Figure 6 shows a comparison between predicted and experimental hydrogen yields at different temperatures for 15
sections. As shown in this figure, the hydrogen yield is increased
by an increase in the temperature because of an increase in the
methane conversion. This figure also shows that the reformer
with 4 or 5 sections has a good agreement with the experimental
data, as also confirmed in Figure 5.
The predicted methane conversion is compared to the
experimental data at different steam/methane ratios for 15
sections in Figure 7. As shown in this figure, the methane
conversion is increased with an increase in the steam/methane ratio. This is due to the fact that an increase in the steam/
methane ratio causes the reforming reactions to use the excess
steam and could be the reason of an increase in the methane
conversion. However, with an increase in the steam flow rate,
the hydrogen partial pressure in the reformer is reduced, leading
to a decrease in the permeability of hydrogen and in the methane
conversion. Because of the lower membrane capacity used in
the experimental study reported by Adris et al.,12,21 the changes
in the methane conversion as a result of the reduction in the
permeability are not significant. This figure also shows that the
reformer with 4 or 5 divisions has a good agreement with
the experimental data in terms of the steam/methane ratio,
confirming the findings reported in Figures 5 and 6.

Energy & Fuels, Vol. 21, No. 6, 2007 3597

Figure 6. Hydrogen yield as a function of the temperature at different


numbers of sections (steam/carbon ratio ) 2.4; P ) 0.98 MPa; sweep
gas flow rate ) 80 mol/h; sweep gas pressure ) 0.4 MPa; methane
flow rate ) 74.2 mol/h).

Figure 7. Methane conversion as a function of the steam/methane ratio


at different numbers of sections (temperature ) 650 C; P ) 0.69 MPa;
sweep gas flow rate ) 80 mol/h; sweep gas pressure ) 0.4 MPa;
methane flow rate ) 41.2, 53, and 74.2 mol/h).

Figure 8 shows the parity plot among the experimental hydrogen


permeation rate, the model predictions, and the values calculated
by Adris et al.12,21 As shown in this figure, the average hydrogen
permeation rates calculated by the model [on the basis of the mean
value of effectiveness factor ( ) 0.39) reported by Adris et al.12,21]
are in close good agreement with the experimental data and the
value calculated by Adris et al.12,21 The small deviation between
the models developed in this work with the values reported by
Adris et al.12,21 may be contributed to the different hypothesis used
in the modeling between these two approaches. The results of F-test
analysis also confirms that the model is providing close results in
predicting the behavior of the fluidized bed reformer.
Conclusions
A fluidized bed reactor model was developed for predicting
the performance of FBMRs. The model consists of several ideal

3598 Energy & Fuels, Vol. 21, No. 6, 2007

Figure 8. Comparison of experimental data with the models in terms


of the hydrogen permeation rate through the membrane.

reactors, which are combined in an appropriate manner, so that


the flow of the gas is considered as the plug flow through the
bubble and membrane phases and perfectly mixed through the
emulsion phase in each section. The catalytic reaction expressions and DTP parameters were adapted as reaction and
hydrodynamic submodels, respectively. The ASPEN PLUS
process simulator has been used to solve these two submodels
simultaneously. The model prediction was compared to experimental data in terms of the methane conversion, the hydrogen
yield, and the hydrogen permeation through the membrane, and
close agreement was found. Using the approach proposed in
this study, the prediction of the nonideal fluidized bed membrane
reactors behavior at different operating conditions becomes
possible in ASPEN PLUS. The results of this study could be
used to represent the nonideal fluidized bed membrane system
inside the process simulators.
Acknowledgment. Financial support from the College of
Engineering, Qatar University (project number 06019P) is greatly
acknowledged. Dr. S. Al-Asheh, head of the research committee,
and Dr. H. Alfadala, the dean of the college, are highly appreciated
for their help. The authors are also grateful for the critical and
helpful comments of the anonymous reviewers.
Nomenclature

db ) Bubble mean diameter (m)


dbo ) Bubble initial diameter (m)

SarVar-Amini et al.

dbm ) Bubble maximum diameter (m)


D ) Reformer diameter (m)
dF ) Particle diameter (m)
B ) Bubble phase
F ) Mole flow rate (mol/s)
G ) Acceleration of gravity (m/s2)
H ) Bed height (m)
M ) Membrane tube
U0 ) Superficial velocity at the onset of fluidization (m/s)
Ub ) Bubble velocity (m/s)
ubr ) Bubble rise velocity (m/s)
Ue ) Emulsion gas velocity (m/s)
Umf ) Minimum fluidization velocity (m/s)
Vb ) Bubble-phase volume (m3)
Ve ) Emulsion-phase volume (m3)
VPFR ) PFR volume (m3)
Die ) Diffusivity of component i in the gas mixture (m2/s)
A ) Reformer surface area (m2)
As ) Surface area between phases (m2)
Am ) Membrane surface area in each section (m2)
kiq ) Interphase mass-exchange coefficient (m/s)
Ci,dense ) Average concentration of component i in the dense
phase (kmol/m3)
Ci,bubble ) Average concentration of component i in the bubble
phase (kmol/m3)
T ) Temperature (K)
R ) Universal gas constant (kJ mol1 K1)
Ep ) Activation energy for permeation (kJ/mol)
E ) Emulsion phase
) Permeation effectiveness factor
Qeff ) Effective permeability constant of hydrogen (mol m1
s1 Pa0.5)
R
PH2
) Hydrogen average partial pressure in the reactor (Pa)
M
PH2
) Hydrogen average partial pressure in the membrane
(Pa)
Ar ) Archimedes number
Fg ) Gas mixture density (kg/m3)
Fp ) Solid particle density (kg/m3)
) Gas viscosity (kg m1 s1)
dv ) Solid particle diameter (m)
i ) Section number
Greek Symbols
e ) Emulsion-phase voidage
b ) Bubble-phase voidage
) Bubble-phase fraction
EF7003514

You might also like