Sie sind auf Seite 1von 9

international journal of hydrogen energy 35 (2010) 71427150

Available at www.sciencedirect.com

journal homepage: www.elsevier.com/locate/he

Theoretical comparison of packed bed and fluidized bed


membrane reactors for methane reforming
Fausto Gallucci*, Martin Van Sintannaland, J.A.M. Kuipers
Fundamentals of Chemical Reaction Engineering Group, Faculty of Science and Technology, (IMPACT) University of Twente,
Enschede, The Netherlands

article info

abstract

Article history:

In this theoretical work the performance of different membrane reactor concepts, both

Received 16 November 2009

fluidized bed and packed bed membrane reactors, has been compared for ultra-pure

Received in revised form

hydrogen production via methane reforming. Using detailed theoretical models, the

8 February 2010

required membrane area to reach a given conversion and the prevailing temperature

Accepted 11 February 2010

profiles have been compared. The extent of mass and heat transfer limitations in the

Available online 15 March 2010

different reactors has been evaluated, and strategies to decrease (or avoid) these limitations have been proposed for the fluidized bed membrane reactor concept. The results

Keywords:

show that the packed bed membrane reactor requires in some conditions double

Hydrogen production

membrane area with respect to the fluidized bed membrane reactor.

Membrane reactors

2010 Professor T. Nejat Veziroglu. Published by Elsevier Ltd. All rights reserved.

Fluidized bed
Heat and mass transfer limitations

1.

Introduction

The recent advances in Polymer Electrolyte Membrane Fuel


Cells (PEMFC) for small or medium scale applications make
the production of ultra-pure hydrogen a challenging topic in
energy conversion. On an industrial scale, most of the
hydrogen is currently produced via steam reforming of
methane (SRM). The traditional SRM process consists of
different process steps such as feed gas preheating and pretreatment (for example hydrodesulphurisation), primary and
secondary reformers (often multi-tubular fixed-bed reactors)
and high and low temperature shift converters, CO2 removal
and methanation units. Often a PSA (Pressure Swing Adsorption) unit is used to achieve the desired hydrogen purity. In
view of thermodynamic limitations and the high endothermicity of steam reforming, heat transfer at high temperatures (850950  C) is required, where excess of steam is used
to avoid carbon deposition (typical feed H2O/CH4 molar ratios

25) [1]. For the production of ultra-pure hydrogen for small


scale application, the large number of process units with
complex heat integration and the associated uneconomical
downscaling make this route inefficient. A high degree of
process integration and process intensification can be
accomplished by integrating hydrogen perm-selective
membranes in the steam reformer [2,3]. Via the integration
of hydrogen perm-selective membranes, the number of
process units can be strongly decreased and the total required
reactor volume can be largely reduced, while higher methane
conversions and hydrogen yields beyond thermodynamic
equilibrium limitations can be achieved, at lower temperatures and with higher overall energy efficiencies [47]. Pdbased membrane reactors for pure hydrogen production
have been proposed in literature for different reaction
systems such as methanol reforming [8,9], ethanol reforming
[10], and autothermal reforming [11]. In particular, Gallucci
and Basile [9] have demonstrated the feasibility of packed bed

* Corresponding author. Tel.: 31 53 489 2370; fax: 31 53 489 2882.


E-mail address: F.Gallucci@tnw.utwente.nl (F. Gallucci).
0360-3199/$ see front matter 2010 Professor T. Nejat Veziroglu. Published by Elsevier Ltd. All rights reserved.
doi:10.1016/j.ijhydene.2010.02.050

international journal of hydrogen energy 35 (2010) 71427150

self supported membrane reactors for different fuels. Simakov


and Sheintuch [11] have developed a small scale autothermal
membrane reformer by coupling an exothermic reaction
(carried out in a separate compartment of the reactor) with the
endothermic methane reforming with hydrogen recovery
through Pd-based membranes.
Both packed bed membrane reactors [12,13] and fluidized
bed membrane reactors [1416] have already been presented
in literature for the reforming of methane and advantages and
disadvantages of both concepts have already been discussed.
However a direct quantitative comparison of the two concepts
at the same conditions is lacking. In this paper a direct
comparison between the two concepts is performed for ultrapure hydrogen production via methane reforming using
detailed theoretical models. The extent of mass and heat
transfer limitations in the different reactors is evaluated, and
strategies to decrease (or avoid) these limitations are
proposed.

2.

Reactor configurations

2.1.

Fluidized bed membrane reactor concept

A schematic representation of the considered fluidized bed


membrane reactor configuration is shown in Fig. 1. The
methane steam reforming takes place in a fluidized bed
operated in the bubbling regime. Pure hydrogen is recovered
via dead-end Pd-based membranes inserted into the fluidized
bed. The hydrogen recovery through the membranes shifts
both the methane reforming and water gas shift reactions
towards the products resulting in higher conversion and
hydrogen yield compared with a conventional reformer.
A pressure difference between the reaction side (fluidized bed)
and the permeation side (membrane lumen) is applied in

7143

order to increase the hydrogen permeation through the


membranes. With the fluidized bed membrane reactor
a virtually isothermal condition can be achieved and bed-tomembrane mass transfer limitations are largely avoided. On
the other hand, bubble-to-emulsion phase mass transfer
limitations and the extent of gas back-mixing could deteriorate its performance. In particular, the use of membranes
inside the reactor could decrease the extent of back-mixing
and can also help in decreasing the bubble diameter,
enhancing the bubble-to-emulsion phase mass transfer. With
the help of a two-phase phenomenological reactor model, the
effect of bubble-to-emulsion phase mass transfer limitations
and gas back-mixing are quantified and a possible strategy to
decrease these limitations is proposed.

2.2.

Packed bed membrane reactor concept

A typical tube-in-tube packed bed membrane reactor configuration is considered (see Fig. 2). The catalyst is packed in the
tube side of the membrane while pure hydrogen is recovered
in the shell side of the reactor. Also in this case, a pressure
difference between the reaction side and the permeation side
is applied.
The reactor has been studied with both a 1D model and
a detailed 2D model in order to identify the extent of wall-to-bed
heat transfer limitations and the bed-to-membrane mass
transfer limitations (concentration polarization) and their effect
on the temperature profiles and reactor performance. The
influence of the reactor and particle dimensions is investigated.

3.

Reactor models

In both reactor concepts the reactions considered are the


following:
CH4 H2 O5CO 3H2

(1)

CO H2 O5CO2 H2

(2)

Where the reaction rate expressions are taken from Numaguchi et al. [17]:

r1

Fig. 1 Scheme of the fluidized bed membrane reactor.



k1 pCH4 pH2 O  p3H2 pCO =Keq;1
p1:596
H2 O

Fig. 2 Scheme of the packed bed membrane reactor.

(3)

7144

r2

international journal of hydrogen energy 35 (2010) 71427150



k2 pCO pH2 O  pH2 pCO2 =Keq;2
pH2 O

Bubble phase component mass balances1


(4)
usb;n1 AT rb;n1  usb;n AT rb;n 

The contribution by homogeneous gas phase reactions can


safely be neglected for this reaction system.
The membranes considered in this study are Pd-based thin
dense layer supported membranes prepared by electroless
plating by ECN (Energy research Center of the Netherlands).
The hydrogen permeation rate through the palladium
membranes follows the Richardson equation:
JH2 P0e $eEa =RT $

0:5 
0:5 i
h
plumen
 pshell
H2
H2

(5)

where the values of the apparent activation energy Ea and preexponential factor Pe0 (in agreement with [18]) are
12,540 J mol1 and 2.21  1003 mol s1 m2 Pa0.5, respectively.

nc
X



Kbe;i;n Vb;n rb;n wb;i;n  we;i;n

i1

nc
X
f00membrane
Mw;i Amembrane eb;n
i;mol

 0

w
SFQ
 wb;i;n SFQ
e;i;n
i1

Emulsion phase component mass balances1


nc
X


use;n1 AT re;n1  use;n AT re;n 
Kbe;i;n Vb;n rb;n wb;i;n  we;i;n
i1
nc
X

}membrane
fi;mol
Mw;i Amembrane 1


 eb;n
0i1
1
nrxn
X


@
nj;i rj AVe;n rp;n 1  ee  we;i;n SFQ  wb;i;n SFQ 0
j1

(8)
Transfer term

3.1.

Fluidized bed membrane reactor model

Q use;n1 AT re;n1  use;n AT re;n 

A typical 1D two-phase model for a membrane assisted


fluidized bed reactor has been used for the simulation of the
fluidized bed membrane reactor based on the following main
assumptions:
 Dead-end hydrogen perm-selective membranes are integrated in the reactor.
 The reactor consists of two-phases, viz. the bubble phase
and the emulsion phase.
 The gas flowing through the emulsion phase is considered
to be completely mixed in each section and at incipient
fluidization conditions.
 The bubble phase gas is assumed to be in plug flow (i.e., large
number of CSTRs), where the bubble size and the bubble rise
velocity change for each section.
 The heterogeneous reactions (methane steam reforming
and water gas shift reactions) take place only in the emulsion phase, assuming that the bubble phase is free of catalyst particles.
 Gas removed from the fluidized bed via membranes is
assumed to be extracted from both the emulsion phase and
the bubble phase, distributed according to the local bubble
fraction. The gas extracted from the emulsion phase is
subsequently instantaneously replenished via exchange
from the bubble phase (to maintain the emulsion phase at
minimum fluidization conditions) (following Deshmukh
et al. [19,20]).
 A uniform temperature is assumed throughout an entire
section of the fluidized bed, assuming no heat losses to the
surroundings (adiabatic conditions) and no heat transfer
limitations between the bubble phase and the emulsion
phase [21,22].
The mass and heat balance equations are as follows:
Total mass balance

i1

f00membrane
Mw;i Amembrane 1
i;mol

o
 eb;n 0

Kbe;i;n Vb;n rb;n wb;i;n  we;i;n

i1

where
use;n AT ue;n AT 1  eb;n
usb;0 AT utot AT eb;0
use;0 AT utot AT 1  eb;0
Energy balance (in case of energy supply inside the reactor)
nc


X
T
Hi feed usb;n0 AT rb;i;n0 use;n0 AT re;i;n0
i1

nc
X



HTi out usb;nN AT rb;i;nN use;nN AT re;i;nN

i1

(


nc
X


HTi out f00membrane
Mw;i AT eb;n
i;mol

i1

f00membrane
Mw;i AT 1
i;mol

)

 eb;n
E0

10

where E depends on the kind of energy supply used (see e.g.


Gallucci et al. [15]). All the parameters used are described
elsewhere [15].

3.2.

Packed bed membrane reactor 1D and 2D models

The axial temperature and concentration profiles in both


reaction and permeation compartments were modeled with
a 1D reactor model. The mass and energy conservation
equations read:
Mass conservation equations:


v rsg usg

CMS D

2pri
J
s2cell  pr20

Note that


6

f00membrane
Amembrane 1  eb;n
i;mol

i1
nc
X

vz

usb;n1 AT rb;n1  usb;n AT rb;n use;n1 AT re;n1  use;n AT re;n


nc n
X
f00membrane

Mw;i Amembrane eb;n


i;mol

nc
X

SFx

x if x > 0
:
0 if x  0

(11)

international journal of hydrogen energy 35 (2010) 71427150



v rtg utg

2
CMS D J
vz
ri

s

vwsj;g
v s s vwj;g

rsj;g
rg Dz
rsg usg
vz
vz
vz


!
t
t
6 1  3tg
vw
vw
v
j;g
j;g

jtj
rtg utg
rt Dt
vz
dtp
vz g z vz


2
1  wsj;g CMs D J  jtj rtj;s 0
ri

(12)
(13)

rg


h
i
2
sg  lg  mg V$uI  mg Vu VuT
3


r0  r
3r 30 1  30 exp  6
dp

(15)

(17)

The model equations in 2D axisymmetrical cylindrical coordinates are:


Continuity equation
(18)

Total momentum balance equation

vt

(19)

Friction coefficient
1  32 mg
1  3 3juj
b 150
1:75 3
33
rg d2p
3
dp

(20)

where
juj

q
u2r  u2z

Sr;i 1  3rS Mi

nr
X

ni;j rj

for i 1; 2; .; nc

(26)

j1

Energy balance

 The gas bulk can be described as an ideal Newtonian fluid.


 The porosity profiles have been accounted in the model.




 
3rg u V$ 3rg uu 3Vp  b3rg u  V$ 3sg 3rg g

(25)

where source terms equals:

The remaining constitutive correlations needed to close the


model are summarised elsewhere (see Smit et al. [23])
The 2D model consists of a pseudo-homogeneous, 2D
reactor model consisting of the total gas phase continuity and
NavierStokes equations augmented with gas phase component mass balances and the overall energy balance (see e.g.
Tiemersma et al. [12]). The model is based on a standard
dispersion model which describes the gas phase mass and
energy transport as convective flow with superimposed radial
and axial dispersion.
The following assumptions have been made in this model:

v






v
3rg wi V$ 3rg uwi V$ rg Di $Vwi Sr;i
vt


0
Dr;i
with Di
0 Dz;i

(16)


tw


v tw vTtw
2r0
tw vT

2
astw Ts  Ttw
leff
rtw
g Cp;g
vt
vz
vz
r0  r2i



v3rg
V$ 3rg u 0
vt

(24)

Component mass balance

X

vTt

vTt
v t vTt
3tg rtg Ctp;g rtbulk Ctp;g
 rtg utg Ctp;g


rtj;g Htj;g
leff
vt
vz vz
vz
j



2ri
attw Tt  Ttw
2
r0  r2i

(23)

Porosity profile (Hunt and Tien [24])


(14)

X

vTs

vTs v s vTs
3sg rsg Csp;g rsbulk Csp;g
 rsg usg Csp;g


rsj;g Hsj;g
leff
vt
vz vz
vz
j



2
attw Ttw  Tt
ri

(22)

Newtonian fluid

Energy conservation equations:



2pr0
2
astw Ttw  Ts
scell  pr20

Mg p
ideal gas
RTg

7145

(21)

vT


Cp;g V$ 3rg uT
vt


0
l
with li r
0 lz

3rg Cp;g 1  3rS Cp;S


V$li $VT Sh

(27)

where source terms equals:


Sr;i 1  3rS

nr
X

rj DHj

for j 1; 2; .; nr

(28)

j1

4.

Results and discussion

All the simulations for both the FBMR and PBMR have been
performed without sweep gas and considering vacuum in the
permeation side. A comparison between the two reactor
concepts is carried out based on the membrane area required
for a target conversion, because it is anticipated that
membrane costs is the most important parameter in determining the reactor investment costs. A first comparison has
been made between the fluidized bed membrane reactor
model and the 1D packed bed reactor model at ideal conditions: isothermal conditions and absence of mass and heat
transfer limitations, i.e., the number of grid cells of the 1D
model is set equal to number of CSTRs in the MAFB model. The
results show that as expected in these conditions the two
reactors give identical performance in terms of membrane
area required for a given conversion. In this way it has been
verified that the two models are working properly. The
following simulations have been performed with a heat flux
through the reactor walls. The main difference between the
fluidized bed and the packed bed membrane reactors is
related to the heat management. In fact, for the fluidized bed
membrane reactor it is well known that a virtually isothermal
condition can be achieved while for the packed bed
membrane reactor a temperature drop in the first part of the
reactor is always observed irrespective of the profile of
temperature at the reactor wall.

international journal of hydrogen energy 35 (2010) 71427150

980

980

960

960

Temperature, K

Temperature, K

7146

940
dp = 0.0005 m
dp = 0.0015 m
dp = 0.0025 m

920

900

880

860
0.0

Wall temperature
Reaction temperature

940

920

900
Pre-reforming zone

880

0.2

0.4

0.6

0.8

1.0

1.2

1.4

1.6

1.8

860
0.0

2.0

0.2

0.4

0.6

0.8

1.0

1.2

1.4

1.6

1.8

2.0

z, m

z, m

Fig. 5 Axial temperature profile in a packed bed


membrane reactor with pre-reforming zone.

A typical result for the axial temperature profile in the


packed bed membrane reactor is reported in Fig. 3, indeed
showing a temperature drop of 80100 K in the first part of the
reactor which can give stability and sealing problems for the
membrane. In fact, the membrane material should withstand
a large axial temperature gradient, which might cause the
detachment of the Pd-based layer from the support with
consequent loss in perm-selectivity. Moreover, the first part of
the membrane is not effectively used since it is working at
a relatively low temperature which, following Richardsons
equation, results in a lower hydrogen permeation flux. The
decrease in temperature at the beginning of the reactor also
gives a decrease in the reaction rate. The result is an increase
of the membrane area required for a specified conversion
compared to the case of constant temperature. In particular,
the membrane area required increases by about 21% if
compared with an isothermal operation (which is only
possible in a fluidized bed membrane reactor). The effect of
the particle size on the temperature profile is negligible as also

indicated in the same figure. By changing the particle size the


combination of the temperature change and the change in the
effectiveness factor is counterbalanced so that the final
conversion is practically the same for different particle
diameters (see Fig. 4).
A way to overcome the problem of the temperature drop is
the use of a pre-reforming zone (in our case about 2025 cm)
where no membrane is applied. In this case (pre-reforming
section 25 cm) the membrane is used at an almost constant
temperature (maximum temperature difference 28 K) so that
the stability problems are prevented and the membrane is
effectively used resulting in a lower membrane area needed
for a given conversion (i.e., slightly longer packed bed, but
smaller membrane area). The results are shown in Fig. 5. In
these conditions the increase of membrane area with respect
to an ideal fluidized bed membrane reactor is 13%.
Another difference that can occur between a packed bed
and a fluidized bed is mass transfer limitations between the
bed and the membrane wall which are present in the packed
bed but not in the fluidized bed. To investigate the extent and

1.0

0.06

0.8

0.05

H2 weight fraction, -

CH4 conversion, -

Fig. 3 Axial temperature profile in a packed bed


membrane reactor.

0.6

dp = 0.0005 m
dp = 0.0015 m
dp = 0.0025 m

0.4

0.03

0.02

0.2

0.0
0.0

0.04

0.2

0.4

0.6

0.8

1.0

1.2

1.4

1.6

1.8

z, m

Fig. 4 Methane conversion for different particle size in


a packed bed membrane reactor.

2.0

0.01
0.0

z/L = 0.1
z/L = 0.5
z/L = 0.8
0.2

0.4

0.6

0.8

r/R, -

Fig. 6 Radial profile of the H2 weight fraction for the


isothermal operation mode.

1.0

7147

international journal of hydrogen energy 35 (2010) 71427150

0.060

z/L = 0.1

H2 weight fraction, -

0.055
0.050
0.045
0.040
0.035
0.030

Original H2 permeation rate


2 * Original H2 permeation rate
5 * Original H2 permeation rate

0.025
0.0

0.2

0.4

0.6

0.8

1.0

r/R, -

Fig. 7 Radial profile of the H2 weight fraction for the


isothermal operation mode at different hydrogen
permeabilities.

the influence of the concentration polarization a 2D model


was used.
First the radial H2 concentration profiles have been calculated at different axial positions at isothermal conditions. As
can be seen in Fig. 6, radial concentration profiles are present
but not really pronounced. It can be concluded that for the
present membranes and for small membrane diameters (1 cm
in the simulation shown in the figure), the bed-to-wall mass
transfer limitations have a negligible influence on the
membrane area.
In view of further developments and optimization of Pdbased membranes, higher membrane fluxes will become
possible in the near future. Whether concentration polarization will occur with increased permeability was investigated
numerically. Simulation results where the membrane
permeability was increased with a factor of 2 and 5. As shown
in Figs. 7 and 8, an increase of hydrogen permeability causes
a significant increase of the concentration polarization, even
in membrane tubes with a diameter as small as 1 cm.

Fig. 9 Schematic representation of the membrane reactor


concept with bubble increasing in size.

Fig. 8 reports the relative H2 weight fraction (defined as the


actual H2 weight fraction divided by the H2 weight fraction at
the catalyst center) as a function of the radial position. The
results reported in the Fig. 8 suggest that at higher membrane
permeabilities mass transport limitations to the membrane
wall will negatively affect the reactor performance resulting in
an increased H2 slip through the reactor exhaust. The 2D
model can be further applied to quantify the effects of mass

0.980
0.975

0.8
z/L = 0.1

CH4 conversion, -

Relative H2 weight fraction, -

1.0

0.6

0.4
Original H2 permeation rate
0.2

0.0
0.0

2 * Original H2 permeation rate


5 * Original H2 permeation rate

0.2

0.4

0.6

0.970
No mass transfer limitations
0.965
0.960
0.955

Mass transfer limitations calculated


as Fluidized Bed without internals

0.950

0.8

r/R, Fig. 8 Relative H2 weight fraction for the isothermal


operation mode at different hydrogen permeabilities.

1.0

0.945
1e+0

1e+1

1e+2

1e+3

1e+4

1e+5

Factor multiplying the mass transfer coefficient, -

Fig. 10 Effects of bubble-to-emulsion phase mass transfer


limitations on the conversion (FBMR).

7148

international journal of hydrogen energy 35 (2010) 71427150

CH4 conversion, -

0.980

Table 1 Comparison between staged fluidized bed and


1D packed bed.

0.975

T,  C

0.970

700

0.965

0.960

0.955

0.950

Membrane area, m2

Fig. 11 Membrane area needed for a given conversion in


case of mass transfer limitations (FBMR).

transfer limitations at different membrane diameters, and


different operating conditions.
Concerning the fluidized bed membrane reactor, an
important transfer limitation affecting its performance is the
mass transfer limitation between the bubble phase and the
emulsion phase. In fact, the gas transported inside the bubbles
should be exchanged with the emulsion phase to react. A high
mass transfer limitation (low mass transfer coefficient)
between the bubble phase and the emulsion phase results in
a larger gas slip via the bubble phase and a lower conversion
degree. In our fluidized bed membrane reactor model the
bubble-to-emulsion phases mass transfer coefficient is
calculated with the equations derived for a fluidized bed
without internals. Although the internals (solid membranes)
should enhance the mass transfer characteristics of the bed,
at the moment, reliable equations for bubble-to-emulsion
phases mass transfer coefficient for fluidized bed with
inserts are not available.
For a fair comparison with the packed bed membrane
reactor, the fluidized bed has been simulated with the same
membrane area but also with the same bed length. As
a matter of fact, the bubble-to-emulsion phase mass transfer
limitation increases with increasing bubble diameter, which

Fluidized bed (5 stages)

Packed bed (1D)

20

3.24

3.94

itself increases by increasing the reactor length as schematically indicated in Fig. 9.


As a result of this bubble increase, the methane conversion
decreases as indicated in Fig. 10. The figure shows that the
methane conversion decreases by increasing the mass
transfer limitations. In case of mass transfer limitations
calculated as a fluidized bed reactor without internals (worst
case) the methane conversion decreases tremendously as
compared with the case without mass transfer limitations
(previously indicated as ideal condition for fluidized bed
reactor). The figure also shows that by improving the mass
transfer by a factor of 10 results in a conversion close to the
ideal case without mass transfer limitations.
In order to achieve the same conversion degree of a fluidized bed membrane reactor without mass transfer limitations
the membrane area installed in the reactor needs to be
increased as indicated in the following Fig. 11.
The membrane area required in case of mass transfer
limitations increases 2.4 times with respect to the case
without limitations as reported in the figure.
However, it has to be pointed out that the use of membranes
inside the bed leads to a decrease of the bubble size (both
because of gas extraction through the membranes and because
of break-up of bubbles by the solid membrane tubes) and
a consequent decrease of the mass transfer limitations. Fig. 10
shows that a decrease of 10 times in the mass transfer limitations is enough to reach the limit conversion required. Thus,
a more detailed experimental work on the determination of
the bubble-to-emulsion phase mass transfer coefficient in
a fluidized bed with internals should be carried out.
On the other hand, even considering the worst case
(bubble-to-emulsion phase mass transfer coefficient equal to
a fluidized bed without internals) the mass transfer problem in
the fluidized bed can be easily circumvented. In fact, the mass
transfer resistance is higher when the bubble diameter
1.0

0.980
0.975

0.9

CH4 Conversion, -

CH4 conversion, -

P, bar

0.970
0.965
0.960
0.955

Staged Fluidized bed


2D isothermal model

0.8

0.7

0.6
0.950
0.945

10

Number of stages

Fig. 12 Conversion reached for a given area in case of


mass transfer limitations for different stages (FBMR).

12

0.5
0.05

0.10

0.15

0.20

0.25

z/L, -

Fig. 13 Comparison between a staged fluidized bed and


a packed bed with 2D model. 5 bar reaction pressure.

international journal of hydrogen energy 35 (2010) 71427150

Table 2 Increase of membrane area with respect to the


FBMR due to bed-to-membrane mass transfer limitation.
Conversion
degree

Preaction 1
bar

Preaction 5
bar

Preaction 10
bar

0.975
1

85.4%
98.7%

64.5%
81.3%

60.0%
79.5%

becomes larger, and the bubble diameter increases with


increasing of the bed height, so that we can reduce the bubble
diameter by inserting stagers such as meshing wires at
different reactor heights (i.e., staging the fluidized bed reactor).
In Fig. 12 the conversion in a fluidized bed with mass transfer
limitations is shown for different numbers of stages. For these
simulations the axial position of the stagers has not been optimized. This means that the distance between two stages is
constant for one simulation and it is given by dividing the total
length of the reactor by the number of stages. In a future work an
optimization of the axial position for stagers will be carried out.
The area used in this simulation was kept the same as was
needed in the case of no mass transfer limitations. From the
figure it can be seen that the conversion required can be
achieved already with 34 stages. Thus, dividing the reactor in
different stages completely circumvents the problems of mass
transfer limitation for the fluidized bed membrane reactor.
A direct comparison (in terms of membrane area required)
between the performances of a staged fluidized bed membrane
reactor and a packed bed membrane reactor simulated in
1D (no effects of bed-to-wall mass transfer limitations) is
reported in Table 1 (for a 50 Nm3/h hydrogen production).
The packed bed membrane reactor needs around 22%
larger membrane area if compared with a staged fluidized bed
membrane reactor. Moreover, the membrane in the packed
bed is exposed to an axial temperature variation of about 100 K
in the first 25 cm of reactor length, with possible stability and
sealing.
The fluidized bed becomes even more advantageous
compared to packed bed membrane reactor when the effects of
concentration polarization on the performances of packed bed
are considered. A comparison in terms of methane conversion
between the staged fluidized bed membrane reactor and the
packed bed membrane reactor simulated with the 2D
isothermal model (effect of bed-to-wall mass transfer) is
reported in Fig. 13. The effect of bed-to-wall mass transfer
limitations in the packed bed reactor results in a decrease of
methane conversion compared to the fluidized bed membrane
reactor. An overview of the membrane reactor increase due to
the effects of bed-to-wall mass transfer limitations is reported
in Table 2. In the worst case (complete conversion required and
1 bar reaction pressure) the packed bed membrane reactor
requires almost double the membrane area with respect to the
staged fluidized bed membrane reactor.
Finally, we can state that the more evident advantages of
a fluidized bed reactor with respect to a packed bed membrane
reactor are: constant temperature along the reactor and better
heat integration (see Gallucci et al. [14,15]), no mass transfer
limitation between the fluidized bed and the membrane surface.
Some disadvantages such as the erosion problems and
horizontal membrane sealing should be further studied
experimentally.

5.

7149

Conclusions

In this work, two different membrane reactor concepts for


the H2 production via methane steam reforming have been
compared via detailed models. It has been shown that both
concepts may suffer from mass transfer limitations. For the
fluidized bed membrane reactor the mass transfer limitations
occur between the bubble phase and the emulsion phase. The
effect of these mass transfer limitations on the membrane
area required is quite significant. However, these mass transfer limitations can be easily circumvented by staging the
fluidized bed with consequent break-up of bubbles and
decrease of mass transfer limitations. For the packed bed
membrane reactor, the mass transfer limitations occur
between the catalytic bed and the membrane area (concentration polarization). These mass transfer limitations cannot
be easily avoided (not even with membrane tube diameters as
small as 1 cm), and the packed bed membrane reactor requires
in some cases double the membrane area with respect to the
staged fluidized bed operated at the same conditions. Moreover, a 2025% more membrane area is required by the packed
bed (with respect to the fluidized bed) because of the temperature profile prevailing in the packed bed. With the advance of
the development of H2 perm-selective membranes with everincreasing permeabilities, the advantages of fluidized bed
membrane reactor become more and more pronounced.

Acknowledgment
The authors are grateful to the Dutch Ministry of Economic
affairs for financial support of this work in the EOS program
(Project EOSLT05010).

Table of symbols
AT
Area of bed cross section, m2
Amembrane,n Membrane surface area per cell, n, m2
CSTR
Continuously stirred tank reactor,
Particle diameter, m
dp
Cp
Heat capacity, J/(kg K)
D
Dispersion coefficient, m2/s
Gas diffusivity, m2/s
Dg
Bubble phase fraction,
eb
Emulsion phase fraction,
ee
Activation energy for hydrogen permeation, J/mol
Ea
g
Gravitational acceleration (9.81), m/s2
Hj
Enthalpy of specie j, J/mol
Enthalpy of component i at temperature T at position
HTi;x
x, J/mol
J
Permeation flux through membrane, mol/(m2 s)
Mass flux component j, mol/(m2 s)
jj
Reaction rate constant for ith reaction
ki
Bubble-to-emulsion phase mass transfer coefficient
Kbe,i,n
for component I in cell, n, s1
Keq,i
Equilibrium constant for jth reaction [depending on
the reaction]
Molar mass for component i, kg/mol
Mw[i]
CMD
Average molar mass, kg/mol

7150

pi
P0e
R
rj
ri
r0
Sh
Sr,i
T
u
us
V
vj,i
wz

international journal of hydrogen energy 35 (2010) 71427150

Partial pressure for component i, atm


Pre-exponential factor for permeation of Pd
membrane, mol/(s m2 Pa0.5)
Gas constant (8.3145), J/(mol K)
Reaction rate for jth reaction, mol/(kgcat s)
Inner tube radius, m
Outer tube radius, m
source/sink term for heat balance, J/(m3 s)
source/sink term for mass balance, kg/(m3 s)
Temperature, K
Mixture velocity, m/s
Superficial gas velocity, m/s
Volume, m3
Stoichiometric coefficient for jth reaction and ith
component,
weight fraction,

Greek
a
heat transfer coefficient, J/(m2 K s)
b
friction factor,
r
Density, kg/m3
3
Porosity,
Emulsion phase porosity,
3e
l
Thermal conductivity, J/(m K s)
Effective thermal conductivity, J/(m K s)
leff
Viscosity of gas, Pa s
mg
Stress tensor, kg/(m s)
sg
Molar flux component i through the membrane per
f00membrane
i;mol
cell, mol/(m2 s)
Subscripts
0
Reactor inlet,
b
Bubble phase,
e
Emulsion phase,
g
gas phase,
i
Component i,
j
Number of reaction,
n
Number of CSTRs for emulsion or bubble phase,
r
radial co-ordinate,
s
solid phase,
z
axial co-ordinate,

references

[1] Rostrup-Nielsen JR. Syngas in perspective. Catal Today 2002;


71(34):2437.
[2] Adris AM, Elnashaie SSEH, Hughes R. Fluidized bed
membrane reactor for the steam reforming of methane. Can
J Chem Eng 1991;69(5):106170.
[3] Kikuchi E. Palladium/ceramic membranes for selective
hydrogen permeation and their application to membrane
reactor. Catal Today 1995;25(34):3337.
[4] Gallucci F, Paturzo L, Fama` A, Basile A. Experimental study of
the methane steam reforming reaction in a dense Pd/Ag
membrane reactor. Ind Eng Chem Res 2004;43:92833.
[5] Gallucci F, Paturzo L, Basile A. A simulation study of the
steam reforming of methane in a dense tubular membrane
reactor. Int J Hydrogen Energy 2004;29:6117.
[6] Oklany JS, Hou K, Hughes R. A simulative comparison of
dense and microporous membrane reactors for the steam
reforming of methane. Appl Catal A Gen 1998;170:1322.

[7] Patil CS, van Sint Annaland M, Kuipers JAM. Fluidised bed
membrane reactor for ultrapure hydrogen production via
methane steam reforming: experimental demonstration and
model validation. Chem Eng Sci 2007;62(11):29893007.
[8] Fu C-H, Wu JCS. Mathematical simulation of hydrogen
production via methanol steam reforming using doublejacketed membrane reactor. Int J Hydrogen Energy 2007;
32(18):48309.
[9] Gallucci F, Basile A. Pd-Ag membrane reactor for steam
reforming reactions: a comparison between different fuels.
Int J Hydrogen Energy 2008;33(6):167187.
[10] Gallucci F, Van Sint Annaland M, Kuipers JAM. Pure hydrogen
production via autothermal reforming of ethanol in
a fluidized bed membrane reactor: a simulation study. Int J
Hydrogen Energy 2010;35(4):165968.
[11] Simakov David SA, Sheintuch Moshe. Demonstration of
a scaled-down autothermal membrane methane reformer for
hydrogen generation. Int J Hydrogen Energy 2009;34(21):886676.
[12] Tiemersma TP, Patil CS, van Sint Annaland M, Kuipers JAM.
Modelling of packed bed membrane reactors for autothermal
production of ultrapure hydrogen. Chem Eng Sci 2006;61(5):
160216.
[13] Gallucci F, Comite A, Capannelli G, Basile A. Steam reforming
of methane in a membrane reactor: an industrial case study.
Ind Eng Chem Res 2006;45(9):29943000.
[14] Gallucci F, Van Sint Annaland M, Kuipers JAM. Autothermal
reforming of methane with integrated CO2 capture in a novel
fluidized bed membrane reactor. Part 1: experimental
demonstration. Top Catal 2008;51(14):13345.
[15] Gallucci F, Van Sint Annaland M, Kuipers JAM. Autothermal
reforming of methane with integrated CO2 capture in a novel
fluidized bed membrane reactor. Part 2: comparison of
reactor configurations. Top Catal 2008;51(14):14657.
[16] Mahecha-Botero A, Grace JR, Jim Lim C, Elnashaie SSEH,
Boyd T, Gulamhusein A. Pure hydrogen generation in
a fluidized bed membrane reactor: application of the
generalized comprehensive reactor model. Chem Eng Sci
2009;64(17):382646.
[17] Numaguchi T, Kikuchi K. Intrinsic kinetics and design
simulation in a complex reaction network: steam reforming.
Chem Eng Sci 1988;43:2295301.
[18] Lin YM, Liu SL, Chuang CH, Chu YT. Effect of incipient
removal of hydrogen through palladium membrane on the
conversion of methane steam reforming experimental and
modeling. Catal Today 2003;82:12739.
[19] Deshmukh SARK, Laverman JA, Cents AHG, Van Sint
Annaland M, Kuipers JAM. Development of a membrane assisted
fluidized bed reactor. 1. Gas phase back-mixing and bubble-toemulsion phase mass transfer using tracer injection and
ultrasound experiments. Ind Eng Chem Res 2005;44:595565.
[20] Deshmukh SARK, Laverman JA, Van Sint Annaland M,
Kuipers JAM. Development of a membrane assisted fluidized
bed reactor. 2. Experimental demonstration and modeling
for the partial oxidation of methanol. Ind Eng Chem Res
2005;44:596676.
[21] Deshmukh SARK, Van Sint Annaland M, Kuipers JAM. Heat
transfer in a membrane assisted fluidised bed with
immersed horizontal tubes. Int J Chem React Eng 2005;3:A1.
[22] Patil CS, Van Sint Annaland M, Kuipers JAM. Experimental
study of a membrane assisted fluidized bed reactor for H2
production by steam reforming of CH4. Chem Eng Res Des
2006;84:399404.
[23] Smit J, Van Sint Annaland M, Kuipers JAM. Feasibility study
of a reverse flow catalytic membrane reactor with porous
membranes for the production of syngas. Chem Eng Sci 2005;
60(24):697182.
[24] Hunt ML, Tien CL. Non-darcian flow, heat and mass transfer
in catalytic packed-bed reactors. Chem Eng Sci 1990;45:5563.

Das könnte Ihnen auch gefallen