You are on page 1of 9

Catalysis Today 118 (2006) 228–236

www.elsevier.com/locate/cattod

Theoretical analysis of conversion enhancement in isothermal


polymeric catalytic membrane reactors
Ju-Meng Zheng, José M. Sousa, Diogo Mendes,
Luis M. Madeira, Adélio Mendes *
LEPAE, Chemical Engineering Department, Faculty of Engineering, University of Porto,
Rua Dr. Roberto Frias, 4200-465 Porto, Portugal

Available online 12 July 2006

Abstract
A theoretical study of polymeric catalytic membrane reactors (PCMR) is performed when a reversible A $ B reaction is taking place inside the
catalytic membrane. The PCMR is assumed to be isothermal, and with a perfectly mixed flow pattern for both retentate and permeate chambers. An
analytical expression is derived to calculate the reactor conversion and the reactor conversion enhancement for the case in which the membrane
permeability is the same for the reactant and product. When the membrane has different permeabilities for the reactant and product, a semi-
analytical solution is presented. The results show that, at least theoretically, conversion above thermodynamic equilibrium can be achieved with the
PCMR when the membrane has higher permeability for the product than for the reactant. It is also pointed out that, in the PCMR, the environment
of the catalyst particle should be quite different from that existing in conventional reactors. This should change the properties of the catalyst particle
and it must be carefully considered.
# 2006 Elsevier B.V. All rights reserved.

Keywords: Polymeric catalytic membrane reactors; Conversion enhancement; Gas phase reversible reaction; Permselectivity; Equilibrium constant

1. Introduction In the development of catalytic membrane reactors (CMR),


some efforts have been reported concerning simulation work.
Membrane reactors (MR), combining the use of membranes Most of these studies were carried out for packed bed
together with catalysts, provide an opportunity to achieve membrane reactors (PBMR) [5,11–20], where the membranes
significant enhancement over the thermodynamic equilibrium are permselective but not catalytically active. For these
conversion. Most of the recent research has been done on systems, it is well recognized that conversion above the
reactors incorporating inorganic membranes, due to the usually thermodynamic equilibrium value can be obtained by
high temperatures involved and the sometimes aggressive selectively removing the product from the reaction zone. In
chemical environments. However, there has been a recent other words, as in all membrane applications, the membrane
interest in polymeric membranes operating at low temperatures with high permeability and permselectivity towards the product
for conducting gas phase [1–5] or liquid phase reactions [6–9]. is favored.
In such conditions, inorganic membranes may be replaced by There are also some studies considering polymeric
less expensive, more versatile polymeric membranes. Further- catalytic membrane reactors (PCMR) where the catalyst is
more, a dense polymer membrane can actively take part in the incorporated inside the membrane, either for liquid phase
reaction by influencing sorption and diffusion behavior of the [21,22] or gas phase reaction [23–29]. When the PCMR was
reactants and products [10]. applied to a gas phase reaction, and whatever the reaction
stoichiometry considered [23–29], it was concluded that a
conversion above the thermodynamic equilibrium value could
be achieved if the reactant diffusion coefficient was lower, or
* Corresponding author. Tel.: +351 22 508 1695; fax: +351 22 508 1449. the sorption coefficient was higher, than that of the product.
E-mail address: mendes@fe.up.pt (A. Mendes). However, the relative conversion variable used in these papers
0920-5861/$ – see front matter # 2006 Elsevier B.V. All rights reserved.
doi:10.1016/j.cattod.2005.12.009
J.-M. Zheng et al. / Catalysis Today 118 (2006) 228–236 229

considered. Conclusions are obtained for the model react-


Nomenclature ion A $ B and extended for other equilibrium limited
reactions.
Di diffusion coefficient of component i (m2/s) In the present study, a PCMR is considered with perfectly
Hi Henry’s sorption constant of component i (mol/ mixed flow pattern in both the retentate and permeate sides
(m3 Pa)) and isothermal operation. The conversion reached in the
kdm direct reaction rate constant in membrane phase PCMR is compared with the thermodynamic equilibrium
(s1)) one, which is the maximum possible conversion attained in a
Keg gas phase reaction equilibrium constant conventional reactor. Moreover, the influence of the
Kem membrane phase reaction equilibrium constant membrane permselectivity, reaction equilibrium constant,
p partial pressure in equilibrium with the concen- Thiele modulus, dimensionless contact time and total
tration inside the membrane (Pa) pressure difference on the performance of the PCMR is
Q volumetric flow rate (m3/s) briefly discussed.
rm apparent reaction rate in the membrane phase
(mol/m3/s)
2. Model development
R universal gas constant (Pa m3/mol/K)
S membrane surface area (m2)
The PCMR considered in the present work is sketched in
T absolute temperature (K)
Fig. 1, and it has been reported elsewhere [23–25]. A
x dimensionless membrane spatial coordinate
hypothetical reversible reaction of the type A $ B is studied.
XPCMR conversion of reactant A in the PCMR
The model development is based on the following main
Xeg thermodynamic equilibrium conversion
assumptions [23–25]:
z membrane spatial coordinate (m)

Greek symbols (1) Steady-state operation.


G dimensionless contact time (2) Isothermal operation.
aB/A membrane permselectivity (3) Perfectly mixed flow pattern on the retentate chambers.
x Relative PCMR conversion (ratio between Assuming a perfectly mixing flow pattern in the retentate
XPCMR and Xeg ) chamber, the permeate chamber should have also a
d thickness of the membrane perfectly mixing flow pattern, once the emerging gas at the
f Thiele modulus for catalyst particle permeate side should have the same composition
f0 Modified Thiele modulus used in present work irrespective of the position in the membrane surface if
g total pressure difference the membrane is homogeneous.
li dimensionless permeability of component i (4) Negligible total pressure drop in the retentate and
ni stoichiometric coefficient of component i permeate chambers.
c dimensionless pressure (5) Negligible external transport limitations in the gas-
z dimensionless volumetric flow rate membrane interface.
(6) Fickian transport across the membrane.
Superscript (7) Henry’s law describing the sorption/desorption equili-
F feed side brium between the bulk gas phase and the membrane
P permeate chamber phase.
R retentate chamber (8) Unitary activity coefficients.
(9) Constant diffusion and sorption coefficients.
Subscript (10) Homogeneous catalyst distribution throughout the mem-
i ith component brane.
ref reference component (reactant A in this paper)

to evaluate the performance of a PCMR considers that the


reaction equilibrium constant in the membrane phase is
numerically equal to the one in the gas phase. This
assumption is very limitative. The present paper uses a
simplified relationship between the reaction equilibrium
constant for the gas phase and for the membrane phase, which
can be obtained from the thermodynamics [30,31]. Inserting
the relation for the equilibrium constants in the previous
model [23–29], it is expected that different results are
obtained. This is precisely the aim of this work where a Fig. 1. Schematic diagram of the polymeric catalytic membrane reactor
hypothetical reversible reaction of the type A $ B is (PCMR).
230 J.-M. Zheng et al. / Catalysis Today 118 (2006) 228–236

(11) Reaction occurs only on the surface of catalyst particles. Partial and total mass balances in the retentate chamber
(12) Equal concentration of reaction species in the catalyst are:
surface and polymer (any relationship could be considered 
QF pFi QR pRi dp 
in principle, but this one simplifies the original problem).  þSDi H i i  ¼ 0
(13) Elementary reaction rate law. RT RT dz z¼0 
(6)
F F
Q p R R
Q p X d pi 
 þS Di H i  ¼0
With these assumptions, the mass balance in the membrane RT RT dz 
i z¼0
is [23–25]:
where is the partial pressure of species i in the feed; pF and
pFi
d2 ðH i pi Þ P
p are the total pressures of the feed and effluent retentate,
Di þ ni r m ¼ 0 i ¼ A; B (1)
dz2 respectively; QF and QR are the total volumetric flow-rates of
where Di is the diffusion coefficient of component i, Hi is the the feed and effluent retentate, respectively; R is the universal
corresponding Henry’s sorption coefficient, pi the gas phase gas constant, T is the absolute temperature and S is the
partial pressure in equilibrium with the concentration of com- membrane surface area.
ponent i inside the membrane (Ci = Hi pi), z the spatial coor- Partial and total mass balances in the permeate chamber are:
dinate perpendicular to the membrane surface, ni the P p

Q pi dp 
stoichiometric coefficient of component i (nA = 1, nB = 1) þSDi H i i  ¼ 0
RT dz z¼d
and rm the membrane phase apparent reaction rate:  (7)
  P P
Q p X d pi 
H B pB þS Di H i  ¼0
r m ¼ kdm H A pA  (2) RT dz 
Kem i z¼d
P
where kdm is the forward reaction rate constant and Kem is the where Q is the total volumetric flow-rate of the effluent
reaction equilibrium constant when the reaction is taking place permeate.
inside the membrane phase. Since we will compare the con- Eqs. (1) and (4)–(7) in dimensionless form are as follows:
version reached in the PCMR with that obtained in a conven- 2  
d2 c i f0 aB=A Keg cB
tional reactor, it is more convenient to use the gas phase þni c  ¼0
reaction equilibrium constant (Keg ). The ratio of the reaction dx2 li 1 þ aB=A Keg A Keg
; (8)
equilibrium constant in the gas phase and in the membrane x ¼ 0; ci ¼ cRi
phase is given by [31]: x ¼ 1; ci ¼ cPi
HB g 
Kem ¼ K (3) dc 
HA e cFi  zR cRi þ G li i  ¼0 (9)
dx x¼0
thus, Eq. (2) may be written as: 
  X dc 
p 1  zR cR þ G li i  ¼0 (10)
r m ¼ kdm H A pA  Bg (4) dx x¼0
Ke

the boundary conditions for Eq. (1) are: P P dci 
z ci þ G li ¼0 (11)
dx x¼1
z ¼ 0; pi ¼ pRi
(5) 
z ¼ d; pi ¼ pPi X dci 
zP cP þ G li ¼0 (12)
where pRi and pPi are the partial pressures of species i in the dx x¼1
retentate and permeate chambers, respectively, and d is the
membrane thickness. where:

PFA R PR PP PR PP PP R PR
cFA ¼ ; cA ¼ A ; cPA ¼ A ; cRB ¼ B ; cPB ¼ B ; cP ¼ ;c ¼ ¼1
Pref Pref Pref Pref Pref Pref Pref
QF R QR P QP
zF ¼ ;z ¼ ;z ¼
Qref Q Qref
 ref 
Di H i DB H B z ; (13)
li ¼ aB=A ¼ ;x ¼
Dref H ref DA H A d
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi vffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
!ffi
  u
k m
D =D u k m
1 SDref H ref RT cR
f0 ¼ d d

A B
¼ dt d 1þ ; G ¼ ; g ¼
g
DA Kem Dref aB=A Ke dQref cP
J.-M. Zheng et al. / Catalysis Today 118 (2006) 228–236 231

the subscripts i and ref refer to the ith and reference compo- when Eqs. (16) and (17) are substituted into the retentate
nents, respectively, and the superscripts F, R, and P refer to the and permeate sides mass balance equations, Eqs. (9)–(12), they
feed, retentate and permeate conditions, respectively. c desig- can be simplified and lead to the following six algebraic
nates the dimensionless pressure, and z the dimensionless equations:
volumetric flow rate. li is the dimensionless membrane perme-
ability. For convenience, we also define aB/A as the membrane zP
1  zR  ¼0 (18)
permselectivity (ratio between product (B) and reactant (A) g
permeability). G is the dimensionless contact time (ratio   
zP 1  aB=A R 1
between the characteristic feed flow time and the characteristic ¼ G aB=A ðcA  cPA Þ þ 1  (19)
permeance time of the reference component) and g is the total g aB=A g
pressure difference (ratio between retentate and permeate side
total pressure). Feed conditions are taken as the reference for cFA  zR cRA  zP cPA
Pref and Qref. Component A is taken as the reference for Dref and aB=A f½coshðf0 Þ  1
¼G
Href. Variable f0 in Eq. (13) is similar to the Thiele modulus 1 þ aB=A Ke g
sinhðf0 Þ
expression for a first-order reversible reaction and slab geo-   
g R P 1
metry of catalyst used by Froment and Bischoff [32]:  ð1 þ Ke ÞðcA þ cA Þ  1 þ (20)
g
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
 
kd 1 
f¼d 1þ (14) p P aB=A f0
Dref Ke z cA ¼ G g ½ðK g þ 1ÞcRA  1
1 þ aB=A Ke sinhðf0 Þ e
 
however, for catalytic membrane systems there is not just one f0 coshðf0 Þ g P 1
 ðKe þ 1ÞcA 
but two surface concentrations (retentate and permeate). It is sinhðf0 Þ g
then not possible to use the generalized Thiele modulus defini-  
1  aB=A R P 1
tion [32]: þ ðcA  cA Þ þ 1  (21)
aB=A g
Z Cs 1=2
r A ðC s Þ
f ¼ d pffiffiffi DA ðCÞr A ðCÞdC (15) cRA þ cRB ¼ 1 (22)
2 C eq

1
for obtaining a consistent Thiele modulus definition for the cPA þ cPB ¼ (23)
present system (where Cs is the concentration on the surface of g
the catalyst particle). In this set of equations, the input parameters are
In this way, variable f0 should be viewed as a modified cFA ; g; G ; f; Keg and aB=A , and the output parameters are
Thiele modulus. A similar expression was used by Winkelman cRA ; cPA ; cRB ; cPB ; zR and zP . Then, the PCMR conversion,
and Beenackers [33], when modeling a gas–liquid reactive XPCMR, defined as the conversion of the reactant species
absorption process, where reactant and product had different (component A) that can be obtained by the PCMR, is calculated
diffusion coefficients in liquid phase. in terms of feed and output gas phase compositions and flow
In the literature, the analytical solutions for Eq. (8) (two rates:
components) are expressed in two different formulations
[25,33]. The solution presented by Winkelman and Beenackers cFA  zR cRA  zP cPA
X PCMR ¼ (24)
[33] is adopted here because it provides an analytical solution cFA
for the set of dimensionless mass balance equations (Eqs. (9)–
The relative PCMR conversion (xPCMR), defined as the ratio
(12)). Winkelman and Beenackers’s solution was originally
between the reactor conversion (XPCMR) and the thermody-
used to describe simultaneous absorption/desorption with
namic gas phase equilibrium conversion (Xge ), is:
reversible first-order chemical reaction in gas–liquid absorp-
tion. It can be adopted and re-written in the dimensionless form X PCMR
xPCMR ¼ (25)
as: Xge
sinh½f0 ð1  xÞ aB=A Keg cRA  aB=A cRB The thermodynamic equilibrium conversion depends on
cA ¼
sinhðf0 Þ 1 þ aB=A Keg the gas phase reaction equilibrium constant and feed
composition:
sinhðf0 xÞ aB=A Keg cPA  aB=A cPB
þ
sinhðf0 Þ 1 þ aB=A Keg
ðKeg þ 1ÞcFA  1
cR þ aB=A cRB cPA þ aB=A cPB Xge ¼ (26)
þ ð1  xÞ A þ x (16) ðKeg þ 1ÞcFA
1 þ aB=A Keg 1 þ aB=A Keg
 R  Basically, Eqs. (18)–(23) can be solved by numerical
cRA  cA ðcA  cPA Þ
cB ¼ cRB þ  R P
þ ðcB  cB Þ x (17) methods [24,25]. However, for a specific case, i.e. when the
aB=A aB=A membrane has the same permeability for the reactant and
232 J.-M. Zheng et al. / Catalysis Today 118 (2006) 228–236

product (aB/A = 1), an analytic solution can be derived, as 3. Results and discussion
follows:
3.1. The membrane permeability is the same for the
R g1 reactant and product
z ¼1G (27)
g
In this condition, the PCMR conversion (XPCMR) and relative
zP ¼ G ðg  1Þ (28) PCMR conversion (xPCMR) can be computed by Eqs. (33) and
(34). An analysis of Eq. (34) shows that the relative PCMR
1 conversion approaches its maximum value of unit when the
cRA ¼ Thiele modulus tends to infinite (f0 ! 1, 9 ! 1, ! 0,
1 þ Keg
þg1
h xPCMR ! 1). This means that a conversion enhancement over
þ 2 2Þ the thermodynamic equilibrium will never occur when the
G ð
h  þ G ½g þ ð1=gÞ þ ð1=G Þ  2ð
h  1Þ þ g
membrane permeability is the same for the reactant and
ð1 þ Keg ÞcFA  1 product. Since the reaction product has the same transport
 (29)
1 þ Keg properties as the reactant in the membrane, there is no
 separation effect and it is impossible to obtain a conversion
1 1 enhancement relative to the thermodynamic equilibrium value.
cPA ¼
g 1 þ Keg When the membrane has the same permeability for the
þ reactant and product, the relative PCMR conversion depends
h2  2 Þ þ G ½g þ ð1=gÞ þ ð1=G Þ  2ð
G ð h  1Þ þ g only on the Thiele modulus, dimensionless contact time and

g F
ð1 þ Ke ÞcA  1 total pressure difference. This is illustrated in Fig. 2. The
 (30) relevant variables used in the calculation are: Keg ¼ 1 and
1 þ Keg
cFA ¼ 1(cFB ¼ 1  cFA ¼ 0). It can be seen that the relative
reactor conversion always increases with the Thiele modulus.
cRB ¼ 1  cRA (31)
In general, the Thiele modulus reflects the relative rate of the
reaction processes compared to the diffusion across the
1 membrane. A low Thiele modulus value indicates that species
cPB ¼  cPA (32)
g move through the membrane with small probability of reaction.
A high Thiele modulus, on the other hand, means that species
where
are rapidly consumed at the catalyst surface, and thus just a
f0 cosh ðf0 Þ f0 small fraction permeates through the membrane, leading to a
h ¼ ; ¼ high PCMR conversion. We also can find that the relative
sinh ðf0 Þ sinh ðf0 Þ
reactor conversion is greatly influenced by the dimensionless
Substituting Eqs. (27)–(32) into Eqs. (24)–(25), both the PCMR contact time. The relative reactor conversion increases with the
conversion and the relative PCMR conversion, for the case in dimensionless contact time mainly because the loss of reactants
which the membrane has the same permeability for the reactant decreases as the retentate stream flowrate decreases and/or
and the product (aB/A = 1), can be calculated with the following membrane permeability increases. The other conclusion that
expressions: can be made from Fig. 2 is that the total pressure difference can

 
½1  G þ ðG =gÞ
h þ G ðg  1Þ þ ðg  1Þ½1  G þ ðG =gÞ ð1 þ Keg ÞcFA  1
X PCMR ¼ 1 (33)
G ð 2 2
h  Þ þ G ½g þ ð1=gÞ þ ð1=G Þ  2ðh  1Þ þ g ð1 þ Keg ÞcFA

xPCMR enhance or penalize the conversion, depending on the Thiele


½1  G þ ðG =gÞ
h þ G ðg  1Þ modulus values. In the low Thiele modulus regime, where
þ ðg  1Þ½1  G þ ðG =gÞ reaction controls, a low total pressure difference will decrease
¼1 the reactant loss and thus favor the reactor conversion. On the
h2  2 Þ þ G ½g þ ð1=gÞ þ ð1=G Þ  2ð
G ð h  1Þ þ g
other hand, at high Thiele modulus values (diffusion controlled
(34) regime), a large total pressure difference will increase the
When the reactant and product have different permeabilities driving force of the diffusion process. However, the influence of
in the membrane (aB/A 6¼ 1), the set of equations (Eqs. (18)– the total pressure difference is inconspicuous in most
(23)) are solved by a Matlab program. A fast and accurate conditions. A notable influence of the total pressure difference
simulation can be obtained for such a simple algebraic on the reactor conversion enhancement can only be observed
equations system. when the dimensionless contact time is high.
J.-M. Zheng et al. / Catalysis Today 118 (2006) 228–236 233

(curves 1–3). On the other hand, when the reactant permeates


through the membrane faster than the product (aB/A < 1), the
reaction equilibrium shift is hindered by the ‘negative’
membrane separation effect: the higher permeability of the
reactant through the membrane penalizes the conversion. In such
a condition, the higher the membrane permselectivity for the
reactant (aB/A  1), the more seriously it penalizes the reactor
conversion (curves 3–5).
When the product permeates through the membrane faster
than the reactant (aB/A > 1), it can be seen that the relative
conversion reaches a peak well above the thermodynamic value
in a medium Thiele modulus region and then the relative
conversion gradually decreases to the unit when the Thiele
modulus further increases. This happens because in the medium
Thiele modulus range, in which the PCMR operates in the
intermediate regime, when the product is selectively removed
Fig. 2. Influence of the Thiele modulus, dimensionless contact time and total from the reaction zone the reaction equilibrium is shifted
pressure gradient on the relative PCMR conversion (aB/A = 1). forward and thus a reactor conversion enhancement is attained.
As the Thiele modulus further increases, the reaction becomes
so fast that the membrane separation effect is getting more and
3.2. The permeabilities are different for the reactant and more offset. Finally, when the Thiele modulus approaches
product infinity, the reaction reaches equilibrium instantaneously
throughout the membrane and the reactor conversion
3.2.1. Effect of the permselectivity approaches the thermodynamic equilibrium value.
Fig. 3 shows the relative reactor conversion enhancement As discussed so far, it can be concluded that a PCMR
obtained in the PCMR when the membrane has different conversion exceeding the thermodynamic equilibrium can be
permselectivities for the reactant and product. The other achieved by selecting a membrane with a higher permeability
parameters used in the simulations are: g = 100, ’ = 0.15, Keg ¼ for the product than for the reactant. In other words, the PCMR
1 and cFA ¼ 1. conversion enhancement does not depend on the diffusion or
As can be seen from this figure, a PCMR conversion well sorption coefficients separately but depends on the membrane
above the thermodynamic equilibrium value can be achieved permselectivity, which is the product of both parameters.
when the membrane has a higher permeability for the product
than for the reactant. This is due to the membrane positive 3.2.2. Effect of the gas phase reaction equilibrium constant
separation effect. When the product is selectively removed from Fig. 4 shows the influence of the gas phase reaction
the reaction zone (aB/A > 1), the reaction equilibrium is shifted equilibrium constant on the relative reactor conversion when
forward and thus a reactor conversion enhancement is obtained. g = 100, ’ = 0.15, cFA ¼ 1 and aB/A = 5. For a wide range of
The higher is the membrane permselectivity (aB/A  1), the Thiele modulus, we can see that a higher reactor conversion
more significant is the reactor conversion enhancement enhancement can be obtained for the reaction with lower

Fig. 3. Influence of the membrane permselectivity and Thiele modulus on the Fig. 4. Influence of the reaction equilibrium constant and Thiele modulus on
relative PCMR conversion. the relative PCMR conversion (aB/A > 1).
234 J.-M. Zheng et al. / Catalysis Today 118 (2006) 228–236

Table 1
Maximum PCMR conversion and conversion enhancement (cFA ¼ 1)
System Xeg max
XPCMR xmax
PCMR Conditions for xmax
PCMR

aB/A = 10, Keg ¼ 1 0.5 0.9091 1.8182 (1) g ! 1; (2) total permeation condition
(TPC); (3) medium Thiele modulus
aB/A = 5, Keg ¼ 1 0.5 0.8333 1.6667
aB/A = 5, Keg ¼ 5 0.8333 0.9615 1.1539
aB/A = 5, Keg ¼ 0:2 0.1667 0.5 3
aB/A  1 1 f0 ! 1

equilibrium constant. A high equilibrium constant means the Fig. 5 clearly shows that, even though the reaction product
equilibrium conversion is high itself. So, it is difficult to obtain permeates through the membrane faster than the reactant
a high conversion enhancement. Considering the extreme case (aB/A = 5), when the dimensionless contact time is small
of Keg ! 1 (irreversible reaction), we can never obtain a (G = 0.02 in this case), a reactor conversion over the
conversion enhancement. On the other hand, it was found that, thermodynamic equilibrium value can only be achieved when
in some conditions, the reaction with a high equilibrium the total pressure difference is large enough (g = 100 and
constant leads to higher reactor conversion enhancement in 1000 in this case). As compared with Fig. 2, we can see
high Thiele modulus region. from Fig. 5 that for medium to high Thiele modulus values
Assuming a feed condition of cFA ¼ 1, we found that the the total pressure difference plays a more important role
maximum relative conversion depends on the membrane when the membrane has different permeabilities for the
permselectivity and gas phase reaction equilibrium constant reactant and product. This is a direct consequence of the
(Table 1). When the permeability of the product is smaller than membrane separation effect since the transport of reaction
or equal to that of the reactant (aB/A  1), the maximum species through the membrane is related to their concentration
relative reactor conversion is the unit, which can be approached gradient.
when the Thiele modulus tends to infinite (f0 ! 1, Fig. 6 illustrates the influence of the dimensionless contact
xPCMR ! 1). For the case in which the product permeates time on the relative conversion. For the case in which the
through the membrane faster than the reactant (aB/A > 1), the product permeates through the membrane faster than the
maximum conversion above the thermodynamic equilibrium reactant, the relative conversion increases significantly with
can be obtained for: (1) a very large total pressure difference the dimensionless contact time. This is expected, because a
(g ! 1); (2) the total permeation condition (TPC) [24,25] and high dimensionless contact time means that the fraction of
(3) medium Thiele modulus range. As an example, Table 1 species that enter into the membrane is higher, which will
gives the maximum PCMR conversion and maximum relative reduce the loss of the reactant. From Fig. 6, it can also be seen
PCMR conversion in various combinations of membrane that the maximum relative conversion can be approached at
permselectivity and gas phase reaction equilibrium constant. the total permeation condition when a large total pressure
Basing on simulated results (not all of these results are gradient is applied (g = 100 in this case). The value of the
shown in Table 1), a simple correlation was found which relates maximum relative conversion is consistent with the predic-
the maximum relative conversion with the membrane perms- tion of Eq. (35).
electivity and gas phase reaction equilibrium constant:

aB=A þ aB=A Keg


xmax
PCMR ¼ ðaB=A  1; cFA ¼ 1Þ (35)
1 þ aB=A Keg

3.2.3. Effect of the total pressure difference and


dimensionless contact time
As pointed in previous sections, it is possible to obtain a
PCMR conversion exceeding the thermodynamic equilibrium
value by selecting a membrane with a higher permeability for
the product than that for the reactant. However, to what extend
the reactor conversion will be enhanced is tightly related to
other conditions, such as the Thiele modulus, the total pressure
difference and the dimensionless contact time. As the influence
of the Thiele modulus on the reactor conversion has been
extensively discussed, in this section it is shown how the total
pressure gradient and dimensionless contact time affect the Fig. 5. Influence of the total pressure difference and Thiele modulus on the
PCMR performance. relative PCMR conversion (aB/A > 1).
J.-M. Zheng et al. / Catalysis Today 118 (2006) 228–236 235

pattern, was studied for a reversible A $ B gas phase reaction.


The PCMR conversion was calculated both analytically, for the
case in which the membrane has the same permeability for the
reactant and product, and numerically, otherwise. It was found
that all the factors which may affect the performance of the
PCMR can be combined into six dimensionless groups: Thiele
modulus, dimensionless contact time, total pressure difference,
feed composition, gas phase reaction equilibrium constant and
membrane permselectivity. The simulation results show that, at
least theoretically, the PCMR can overcome the thermody-
namic equilibrium limitations when the reaction product
permeates preferentially through the membrane. This conclu-
sion should hold for any equilibrium limited reaction.
The performance of catalyst particles is affected by the
membrane properties in the PCMR. It is clear that a change of
the ratio between the reactant and product sorption coefficients
Fig. 6. Influence of the dimensionless contact time and Thiele modulus on the should change the ratio between the forward and reverse
relative PCMR conversion (aB/A > 1).
membrane phase reaction rate constants. Accordingly, the
3.3. Effect of the membrane properties on the catalyst Thiele modulus value could be changed. At this point, it is
performance difficult to make a precise prediction of the PCMR
performance.
Previous sections show that, theoretically, a conversion
enhancement over the thermodynamic equilibrium one can be Acknowledgements
obtained by selecting a membrane with higher permeability for
the reaction product than that of the reactant. However, in this Ju-Meng Zheng is grateful to the Portuguese Foundation for
analysis, the catalyst is considered as a ‘black box’. How the Science and Technology (FCT) for the postdoctoral grant
specific conditions, especially the membrane properties affect (reference: SFRH/BPD/14489/2003). Financial support by
the catalyst state and properties is not taken into account. As FCT through the project POCTI/EQU/44994/2002 is also
pointed by Miachon and Dalmon [34], the environment of the acknowledged.
catalyst in the polymeric catalyst membrane reactor may be
quite different from that existing in conventional reactors. This References
originates changes of the catalyst properties.
[1] C. Liu, Y. Xu, S. Liao, D. Yu, Y. Zhao, Y. Fan, J. Membr. Sci. 137 (1997)
The analysis above has shown that the PCMR conversion 139.
greatly depends on the Thiele modulus, which is defined as [2] D. Fritsch, K.-V. Peinemann, J. Membr. Sci. 99 (1995) 29.
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
[3] H. Gao, Y. Xu, S. Liao, R. Liu, J. Liu, D. Li, D. Yu, Y. Zhao, Y. Fan, J.
f0 ¼ d kdm =DA ð1 þ 1=aB=A Keg Þ (for the A $ B reaction
Membr. Sci. 106 (1995) 213.
studied in the present work). In this definition, for a specific [4] S. Ziegler, J. Theis, D. Fritsch, J. Membr. Sci. 187 (2001) 71.
gas phase reaction and a selected membrane, the exact values of [5] J. Huang, L. El-Azzami, W.S.W. Ho, J. Membr. Sci. 261 (2005) 61.
the membrane thickness (d), diffusivity (DA) and permselec- [6] S. Wu, J.-E. Gallot, M. Bousmina, C. Bouchard, S. Kaliaguine, Catal.
Today 56 (2000) 113.
tivity (aA/B) and reaction equilibrium constant (Keg ) can be well [7] H.L. Frisch, S. Maaref, H. Deng-Nemer, J. Membr. Sci. 154 (1999)
estimated. However, the membrane phase direct reaction rate 33.
constant (kdm ) is a factor which is affected by the membrane [8] S. Kaliaguine, C. Bouchard, S.Q. Wu, J. Shu, Canadian Patent 2,206,626
properties and it should be different in comparison to (1998).
conventional systems, e.g. CSTR or packed bed reactor. As [9] G. Langhendries, G.V. Baron, I.F.J. Vankelecom, R.F. Parton, P.A. Jacobs,
Catal. Today 56 (2000) 131.
shown above, Kem ¼ kdm =kim ¼ H B =H A Keg . So, a change of the [10] I.F.J. Vankelecom, P.A. Jacobs, Catal. Today 56 (2000) 147.
sorption coefficients will change the ratio kdm =kim (possibly both [11] K. Mohan, R. Govind, AIChE J. 34 (1988) 1493.
kdm and kim change), as Keg depends only on the temperature. In [12] G. Saracco, J.W. Veldsink, G.F. Versteeg, W.P.M. van Swaaij, Chem. Eng.
this regard, it is impossible to anticipate how the membrane Sci. 50 (1995) 2833.
properties will alter the catalyst performance (kdm ) and [13] G. Saracco, V. Specchia, Chem. Eng. Sci. 55 (2000) 3979.
[14] H.W.J.P. Neomagus, G. Saracco, H.F.W. Wessel, F. Versteeg, Chem. Eng.
consequentially the Thiele modulus and PCMR conversion, J. 77 (2000) 165.
even qualitatively. How the polymeric membrane properties [15] M.P. Harold, C. Lee, Chem. Eng. Sci. 52 (1997) 1923.
affect the catalyst performance will be the aim of a future work. [16] A.G. Dixon, Catal. Today 67 (2001) 189.
[17] M.K. Koukou, N. Papayannakos, N.C. Markatos, AIChE J. 42 (1996)
4. Conclusions 2607.
[18] M.K. Koukou, G. Chaloulou, N. Papayannakos, N.C. Markatos, Int. J.
Heat Mass Transfer 40 (1997) 2407.
The performance achieved by an isothermal polymeric [19] T.M. Moustafa, S.S.E.H. Elnashaie, J. Membr. Sci. 178 (2000) 171.
catalytic membrane reactor, with a perfectly mixed flow [20] C. Hermann, P. Quicker, R. Dittmeyer, J. Membr. Sci. 136 (1997) 161.
236 J.-M. Zheng et al. / Catalysis Today 118 (2006) 228–236

[21] J. Vital, A.M. Ramos, I.F. Silva, H. Valente, J.E. Castanheiro, Catal. Today [29] J.M. Sousa, A. Mendes, Catal. Today 104 (2005) 336.
67 (2001) 217. [30] J.M. Smith, H.C. Van Ness, M.M. Abbott, Introduction to Chemical
[22] A.A. Yawalkar, V.G. Pangarkar, G.V. Baron, J. Membr. Sci. 182 (2001) Engineering Thermodynamics, six ed., McGraw-Hill, 2001 (chapter
129. 13, p. 494-495).
[23] J.M. Sousa, P. Cruz, A. Mendes, J. Membr. Sci. 181 (2001) 241. [31] F. Denbigh, The Principles of Chemical Equilibrium, fourth ed., Cam-
[24] J.M. Sousa, P. Cruz, A. Mendes, Catal. Today 67 (2001) 281. bridge University Press, 1981 (chapter 10, p. 301).
[25] J.M. Sousa, P. Cruz, F.D. Magalhães, A. Mendes, J. Membr. Sci. 208 [32] G.F. Froment, K.B. Bischoff, Chemical Reactor Analysis and Design,
(2002) 57. second ed., John Wiley & Sons, 1990 (chapter 3, pp. 162–163).
[26] J.M. Sousa, A. Mendes, Catal. Today 82 (2003) 241. [33] J.G.M. Winkelman, A.A.C.M. Beenackers, Chem. Eng. Sci. 48 (1993)
[27] J.M. Sousa, A. Mendes, Chem. Eng. J. 95 (2003) 67. 2951.
[28] J.M. Sousa, A. Mendes, J. Membr. Sci. 243 (2004) 283. [34] S. Miachon, J.-A. Dalmon, Top. Catal. 29 (2004) 59.

You might also like