You are on page 1of 8

International Journal of Hydrogen Energy 26 (2001) 551–558

www.elsevier.com/locate/ijhydene

Water–gas shift reaction over porous catalyst: temperature


and reactant concentration distribution
Menderes Levent ∗
Department of Chemical Engineering, Faculty of Engineering, Ataturk University, 25240 Erzurum, Turkey

Abstract
This work represents the results of the theoretical study of the distribution of concentration and temperature in porous
catalyst, carried out over spherical iron oxide=chromia catalyst pellets for the water–gas shift reaction (WGSR). The kinetic
information has been taken from an earlier study (G.C. Chinchen et al., Appl. Catal., 12 (1984) 97) for the above reaction.
The mass and heat transfer distribution coupled with chemical reaction, outside and within the catalyst particles for water–gas
shift reaction was estimated in this present study. A computer program was written to calculate both concentration pro6le and
temperature pro6le for the water–gas shift reaction. ? 2001 International Association for Hydrogen Energy. Published by
Elsevier Science Ltd. All rights reserved.

Keywords: Water–gas shift reaction; Temperature pro6le; Concentration pro6le; Spherical catalyst

1. Introduction description of the factors that determine the e:ectiveness of


porous catalysts.
In order to understand transport phenomena in porous In 6rst synthetic ammonia plant, carbon monoxide was re-
media, one may study 9uid 9ow with chemical reaction moved by liquefaction and scrubbing with hot caustic-soda
through porous catalysts, particularly for the most important solution. Nitrogen was supplied from a cryogenic air sepa-
chemical reactions, since in a major fraction of these, the ration unit. However, it was soon realized that the carbon
catalyst is a porous heterogeneous substance. monoxide liquefaction process was unsuitable for use on
The phenomena of the catalysed reaction are important such a large scale, and a catalytic process was developed
in predicting reaction rates and the e:ect of the catalyst on which converted carbon monoxide to carbon dioxide(and
the conversion rate of the reactant. Nevertheless, transport additional desired hydrogen) by reaction with steam accord-
phenomena at the catalyst surface and within the catalyst ing to the equation
are still poorly understood. So, reactant concentration and
CO + H2 O  CO2 + H2 (1)
temperature distribution connected with catalyst particles
are calculated for the range of temperatures and pressures This reaction was termed the “water–gas shift reaction”
of industrial application. (WGSR). The reaction was 6rst reported in the literature by
Mass transfer within catalyst structures was apparently Mond et al. [7], in 1888.
6rst analysed quantitatively by several workers [1–3]. The
earlier works were further developed by other workers [4 –
1.1. Thermodynamics of water–gas shift reaction
6]. The important result of these analyses is the quantitive
The water–gas shift reaction is moderately exothermic
(GH = −41:1 kJ=mol) and hence its equilibrium con-
∗Tel.: +90-0442-231-2298; fax: +90-0442-233-1133. stant decreases with temperature, and high conversions are
E-mail address: levent62@yahoo.com (M. Levent). favoured by low temperatures. The position of equilibrium

0360-3199/01/$ 20.00 ? 2001 International Association for Hydrogen Energy. Published by Elsevier Science Ltd. All rights reserved.
PII: S 0 3 6 0 - 3 1 9 9 ( 0 0 ) 0 0 1 1 6 - 6
552 M. Levent / International Journal of Hydrogen Energy 26 (2001) 551–558

Nomenclature
ri radial coordinate in the centre of the
A; B constants of integration
catalyst (m)
C concentration of reacting component
r=R normalised distance (dimensionless)
(kmol=m3 )
R radius of the catalyst at the surface (m)
Ci concentration at the catalyst centre
R gas constant (kJ=kmol K)
(kmol=m3 )
Sh Sherwood number (dimensionless)
C0 concentration in the 9uid bulk (kmol=m3 )
T reaction temperature (K)
Cs concentration at the catalyst surface
T0 temperature at the bulk of the catalyst (K)
(kmol=m3 )
Ti temperature at the centre of the catalyst (K)
Ci =C0 concentration gradient (dimensionless)
Ts temperature at the surface of the catalyst
C1 integration constant
(K)
C1 ; C 2 adsorption equilibrium constants of compo-
V spherical-shell volume (m3 )
nents (1=bar)
D; De: e:ective di:usivity of the catalyst (m2 =s)
Greek letters
E i ; Et activation energies of reaction (kJ=mol)
GH heat of reaction (kJ=mol)
f; f; f 1st, 2nd and 3rd derivatives of a function
GHi enthalpy change of component i (kJ=mol)
F(r) function
Gr spherical-shell thickness (m)
hm mass transfer coeJcient (m=s)
GHR enthalpy change of a reaction (kJ=mol)
hT heat transfer coeJcient (W=m2 K)
; e: e:ective thermal conductivity (W=m K)
k reaction rate constant (1=s)
 Thiele modulus for the porous catalyst
k 1 ; k2 adsorption equilibrium constants for reac-
(dimensionless)
tants (1=bar)
 e:ectiveness factor (dimensionless)
k−1 ; k−2 adsorption equilibrium constants for prod-
ucts (1=bar)
Subscripts
Ki adsorption equilibrium constant of compo-
e: e:ective
nent i (1=bar)
i centre conditions of the catalyst
k1 rate coeJcient for reaction (1) (kmol bar
m mass
kg cat. h)
0 bulk conditions
Pi partial pressure of component i (bar)
s surface conditions
Q heat 9ux (kJ=s)
r rate of reaction (kmol=kg cat. h)
Superscripts
r radial coordinate or radius of the catalyst at
n iteration number
any position (m)
m; n; p; q stoichiometric coeJcients for reacting sub-
rs radial coordinate at the surface of the cata-
stances
lyst (m)


is virtually una:ected by pressure. Under adiabatic condi- inlet temperature, which in practice was about 200 –250 C.
tions conversion in a single bed of catalyst is thermody- The conventional iron-based catalysts are not suJciently
namically limited as; the reaction proceeds and the heat active for such low temperature operation, but low temper-
of reaction increases the operating temperature, and so re- ature(LT) water–gas shift catalysts are, and when they are
stricts the conversion possible. Typical carbon monoxide used in the second bed, carbon monoxide exit concentra-
levels that are achieved [8] at the exit with a single-adiabatic tions as low as 0.1– 0.3% can be achieved, an acceptable
bed of chromia-supported iron-based high-temperature(HT) economic level for subsequent methanation.
shift catalyst in an ammonia plant are in the range 2– 4%,

within the temperature range 310 –450 C. The thermody- 1.2. Kinetic model for the water–gas shift reaction
namic equilibrium limitation on the reaction can be reduced
by using two or more beds of high temperature(HT) shift 1.2.1. Kinetcis over high temperature(HT) shift catalyst
catalyst with inter-bed cooling and, perhaps, removal of car- Numerous kinetic studies of the water–gas shift reaction
bon dioxide between the stages. over iron oxide=chromia catalysts have been reported, and
A signi6cant improvement in the conversion of carbon more than 20 di:erent kinetic equations have been proposed.
monoxide could then be obtained with a two-bed opera- Di:erences among authors have been especially marked in
tion, with the second bed operating at the lowest possible their opinion of the reaction mechanism, and the e:ect of
M. Levent / International Journal of Hydrogen Energy 26 (2001) 551–558 553

pressure on the reaction rate. The reasons for this con9ict value of 121.8 kJ=mol deduced from semitechnical experi-
have been attributed to the presence of impurities [9] in the ments with the same catalyst. Their results indicate that sim-
gases used, to varying degrees of mass-transfer limitation ple 6rst-order reaction can be deduced from the Arrhenius
[10] and to the fact that kinetic measurements have been plot.
mostly obtained with integral rather than di:erential reac-
tors which were often operating only at or near atmospheric 1.2.2. Kinetics over low temperature(LT) shift catalyst
pressure. Fott et al. [11] asserted that the e:ect of pressure Much less publications have appeared on the mech-
on the reaction rate is the main criterion which enables the anism and kinetics of the water–gas shift reaction
suitability of a given kinetic equation to be evaluated. over copper-based catalysts than over high temperature
The 6rst systematic attempt to combine rate equations and iron-based catalysts, but similar types of kinetic expres-
reaction mechanisms was made by Kulkova and Temkin [14] sions have been proposed. Indeed, the pore di:usion lim-
who used a pure iron oxide catalyst prepared by precipitation ited version of the Langmuir–Hinshelwood Eq. (7) for
of iron nitrate with ammonia. copper-based shift catalysts is consistent with plant data
Experiments on the continuous catalytic water–gas shift and semi-technical scale results [19].
reaction have been carried out by Hulburt and Vasan [15] k[CO][H2 O]1=2 (1 − [CO2 ][H2 ]=K[CO][H2 O])
in which they make full use of the principles of statistical Rate = : (7)
P 1=2 (1 + C1 [CO] + C2 [H2 O] + · · ·)
design to elucidate the reaction mechanism
k[H2 O] 1.3. Derivation of concentration pro8les with mass
r= : (2)
1 + K[H2 O]=[H2 ] transfer resistance
Kodama et al. [16] used an iron oxide catalyst containing
A 6rst-order irreversible reaction in a spherical catalyst
10% Cr 2 O3 . They arrived at an equation of the type
particle can be represented when the reaction A → B occurs.
k{[CO][H2 O] − [CO2 ][H2 ]=K} A material balance for the reacting component over an ele-
r= : (3)
1 + K1 [CO] + K2 [H2 O] + K3 [CO2 ] + K4 [H2 ] ment of catalyst volume leads to a second-order di:erential
equation. For spherical catalyst particles, when steady state
This expression was put in relation to a steady-state mech-
conditions have been reached, the equation may be derived.
anism(without a rate-determining step) in which reaction
Mass di:usion into the spherical particle at radius r is
occurs between the gaseous components and reduced and
oxidized surface sites. @C
F(r) = −4r 2 D : (8)
Glavachek et al. [14]; Kul’kova and Temkin [17]; Shchi- @r
brya et al. [18] derived the following rate expression on the The expansion of the function F(r) by using Taylor’s theo-
basis of oxidation–reduction model: rem is
k1 k2 {[CO][H2 O] − [CO2 ][H2 ]=K} h2 
r= ; (4) F(r + h) = f(r) + hf (r) + f (r)
k1 [CO] + k2 [H2 O] + k−1 [CO2 ] + k−2 [H2 ] 2!
Podolski and Kim [11]; Fott et al. [12] and Chinchen et h3 
+ f (r) + · · · : (9)
al. [13] carried out experiments in recycle reactors and used 3!
statistical techniques to discriminate between rival models. On applying Taylor’s theorem to Eq. (8),
They also re-examined some of the previously published @C
data, and concluded that only Langmuir–Hinshelwood and f(r + Gr) = 4r 2 D
@r
the power-law model could adequately accommodate all of  
the experimental results. @ @C
+4DGr r2 + 0((Gr)2 ): (10)
Langmuir–Hinshelwood: @r @r
kKCO KH2 O {[CO][H2 O] − [CO2 ][H2 ]=K} A mass balance over the shell of catalyst Gr is
r= :
{1 + KCO [CO] + KH2 O [H2 O] + KCO2 [CO2 ] + KH2 [H2 ]}2
Rate of di:usion into element
(5)
=rate of di:usion out of element
Power law [9,12,17]: +rate of chemical reaction
 
r = ak[CO]m [H2 O]n [CO2 ]p [H2 ]q ; (6) @C @ @C
4r 2 D + 4DGr r2 + 0((Gr)2 )
@r @r @r
Chinchen et al. [13] studied the kinetics of high-tempera-
@C
ture shift catalyst from experiments with a semi-technical =4r 2 D + 4r 2 GrKC: (11)
unit and a microreactor system under relevant industrial @r
conditions combined to give consistent kinetics parameters. After simplifying, we obtain
 
They estimate the activation energy to be 129:4±2:1 kJ=mol 1 @ 2 @C k
r = C; (12)
from microreactor experiments, in good agreement with the r 2 @r @r D
554 M. Levent / International Journal of Hydrogen Energy 26 (2001) 551–558

where r is the radial co-ordinate, C is the concentration of 1.4. Estimation of temperature pro8le in catalyst pellets
reacting component, k is the reaction velocity constant and
D is the e:ective di:usivity. On di:erentiating Reaction with mass and heat transfer occurs simulta-
2 neously at any position within the pellet. The resulting
@C 2 @C k
+ = C: (13) intrapellet concentration and temperature gradients cause
@r 2 r @r D
the rate to vary with position.
The di:erential equation describing 6rst-order reaction and The energy balance for a spherical catalyst is
di:usion in the spherical particle, puts
@2 T 2 @T kC
+ = GH; (17)
V = Cr; @r 2 r @r e:

then where GH is the heat of reaction, k the reaction rate


constant, e: the e:ective thermal conductivity, and r the
V particle radius.
C= :
r The solution of Eqs. (17) and (13) with these boundary
By substituting in Eq. (13), we get conditions gives the temperature pro6le within the pellet.
Eliminating kC from Eqs. (13) and (17) yields
@2 V k 2

− V = 0: @C 2 @C e: @2 T 2 @T
@r 2 D De: + = + ;
@r 2 r @r GH @r 2 r @r
This di:erential equation has the solution

    @ 2 @C e: @ 2 @T
k k De: r = r : (18)
V = Cr = A exp r + B exp − r ; (14) @r @r GH @r @r
D D
By integrating Eq. (18) once using the following boundary
where A and B are constants of integration. Eq. (10) is solved conditions:
for the boundary conditions:
@C
(1) the concentration is 6nite at the centre of the sphere = 0; r = 0;
@r
(r = 0, CA = 6nite); (2) at the external surface of the catalyst
pellet, @T
  = 0; r = 0;
@C @r
r = R h(C0 − Cs ) = D ;    
@r r=R @C e: @T
De: r 2 = r2 + C1
because the mass 9ux from the bulk to the catalyst surface @r GH @r
is equal to the di:usion 9ux within the pellet. Applying the
and then integrating a second time using the following
boundary conditions on Eq. (14) to get the concentration
boundary conditions:
pro6le in the catalysed reaction for the spherical catalyst
particle, C = Cs ; r = rs ;
C
=
C0 T = Ts ; r = rs ;

k
sinh D
r
    : (15) one obtains
[ Rr sinh k
D
R + hDr
mR
k
D
cosh k
D
R − hmDrR2 sinh k
D
R] GH
De: (Cs − Ci ) = (Ts − Ti ): (19)
e:
Eq. (15) shows that the concentration pro6le depends only
upon dimensionless parameters, the Thiele modulus , the This result, originally derived by Damkoehler [2], is not
Sherwood form of hm R=D, and the normalised distance r=R. restricted to 6rst-order kinetics, but is valid for any form of
Here, C0 is the concentration in the 9uid bulk, kmol=m3 , hm rate expression and it shows that the maximum temperature
is the mass transfer coeJcient, m=s, De: the e:ective dif- rise depends on the heat of reaction, transport properties of
fusivity of the catalyst, m2 =s, k the reaction velocity con- the pellet, and the surface concentration of the reactant.
stant, 1=s, R the radius of the catalyst at the surface, m, and The boundary conditions in the presence of surface resis-
r = radius of the catalyst at any position, m. tance to heat of mass 9ow are
The e:ectiveness factor in terms of Thiele modulus and
Sherwood number, Sh is r = rs ; hT (T0 − Ts ) = GHhm (C0 − Cs );
 

3  coth  − 1 @T GHDe: @C
= 2 1
: (16) r = ri :
 1 + Sh ( coth  − 1) @r e: @r
M. Levent / International Journal of Hydrogen Energy 26 (2001) 551–558 555

Therefore, the general solution is


GHDe: GHhm
T(x; y; z) = C(x; y; z) + T0 − C0
e: hT

hm De:
+Cs(x; y; z) GH − : (20)
hT e:
This equation was used in a computer program to estimate
the temperature pro6le at di:erent values of the Thiele mod-
ulus.

1.5. Estimation of extraparticle concentration and


temperature di:erences
Fig. 1. Internal temperature di:erences at various bulk temperatures
At steady state the reaction rate expressed per unit mass (Ts − Ti ) K, at  = 0:21 W=(mk).
of pellet, may be written either in terms of di:usion rate
of bulk gas to the surface or in terms of the product of the
rate that would have been determined if the concentration 2. Results and discussion
in the particle were uniform and e:ectiveness factor would
be taken into account: In this study the reaction conditions have been taken as:
total pressure of the process is 31 bar, carbon monoxide
r = Ahm (C0 − Cs ) = VkCs ; (21) partial pressure is 3 bar, steam partial pressure is 16 bar,
where C0 and Cs are the concentration in the bulk gas hydrogen partial pressure is 8 bar, and carbon dioxide partial
and surface, respectively, hm is the mass transfer coeJcient pressure is 4 bar. The temperature range was taken to be

between bulk gas and solid surface, and A is the external 250 –450 C.
surface area per unit mass of the pellet. V is the catalyst We have applied in this study kinetic information
volume,  is the e:ectiveness factor, and k is the reaction on the water–gas shift reaction over commercial iron
rate constant. oxide=chromia catalyst mainly taken from Chinchen et
The temperature di:erence between 9uid, T0 , and surface, al. [13,20]; they found that kinetics is best described by
Ts , can be represented in terms of heat transfer rate from the Langmuir–Hinshelwood model. On the basis of the
the bulk to the pellet in the case of an endothermic reaction, experimental work carried out on a semi-technical plant
or from the pellet to 9uid bulk in the case of exothermic at pressures up to 30 bar they found that the reaction is
reaction, signi6cantly pore-di:usion limited in 8.5 mm pellets at

temperatures above 350 C. The reaction of water–gas shift
Q = AhT (T0 − Ts ) = VkCs GHR ; (22) over a spherical catalyst has been assumed initially to
where Q is the heat 9ux, hT is the heat transfer coeJcient be an isothermal reaction. This hypothesis may be tested
and GHR is the enthalpy of a reaction. for temperature pro6les, since signi6cant deviations from
The formulae below are solved in a sequential estimation isothermality in temperature pro6les will mean that the
procedure in which the estimations are continually updated. hypothesis is not correct.
The formulae are for concentration distribution Fig. 1 shows the variation of the temperature di:erences
between the surface and the interior of the catalyst particles
Ahm (C0 − Csn ) = Vn−1 k n−1 Csn ; (23) with bulk temperature for the thermal conductivity value of
where the e:ectiveness factor  is given by Eq. (16) and 0.21 W=(mK) for iron oxide=chromia catalyst. Fig. 1 shows
the reaction velocity constant k is estimated at the surface that at that value of thermal conductivity, the temperature
◦ ◦
temperature Ts . It is supposed that the particle is isothermal. di:erence is a minimum of −4:5 C at 440 C. It also shows
The equation for the surface temperature is that temperature di:erences become less at the industrial

range of temperature between 365 and 540 C. Hence, the
AhT (T0 − Tsn ) = Vn−1 · k n−1 Csn GHR ; (24)
reaction is essentially isothermal within the range of indus-
where the superscripts in Eqs. (23) and (24) identify the trial application. It is clear that the hypothesis of an isother-
iteration number. mal pellet is satisfactory. The value of thermal conductivity
These formulae are iterated to convergence using suc- is reported by Chinchen et al. [13,20].
cessive substitution. Once convergence has been obtained, The external temperature di:erences are shown in Fig. 2,

the surface temperature corresponding to the converged val- small at temperatures below 340 C, but increasing with bulk
◦ ◦
ues is given by Eq. (24) and surface concentration by Eq. temperature to reach 55 C at 540 C. Fig. 3 shows temper-
(23). Even at high temperatures the pro6le of temperature is ature pro6le with bulk temperatures which has been found
essentially isothermal, and therefore the method of calculat- from Eq. (20). The lines marked (1 ), (2 ) and (3 ) show
ing temperature and composition pro6les is valid. that the temperature within the catalyst is constant from the
556 M. Levent / International Journal of Hydrogen Energy 26 (2001) 551–558

Fig. 5. The overall temperature di:erences at various bulk temper-


atures, at  = 0:21 W=(mk).
Fig. 2. External temperature di:erences at various bulk tempera-
tures, at  = 0:21 W=(mk).

Fig. 3. Temperature pro6les in a porous spherical catalyst, at di:er-


ent values of the Thiele modulus (1 = 3:5; 2 = 9:64; 3 = 26:89)
and  = 0:21 W=(mk). Fig. 6. Reaction rate constant versus experimental bulk tempera-
tures, at  = 0:21 W=(mk).

The logarithm of reaction rate constant versus experi-


mental bulk temperatures at thermal conductivity value of
0.21 W=(mK) is shown in Fig. 6. As seen in this 6gure, the
logarithm values of reaction rate constant increase linearly
with bulk temperatures. The reaction rate constant has been
extrapolated from the experimental results of Chinchen et
al. [13,20].
Arrhenius plots for the reaction water–gas shift are shown
in Fig. 7 in which the upper curve(1) shows the dependence
of the intrinsic reaction velocity constant upon exp(1=T ).
The gradient of the line is related to the activation energy for
Fig. 4. Temperature ratio in a porous spherical catalyst, at the reaction, and it is a matter of considerable interest that
(1 = 0:53; 2 = 1:48; 3 = 3:98; 4 = 8:87; 5 = 19:75) and at at low temperatures to the right of the graph, the activation
 = 0:21 W=(mk). energy is 129:4 kJ mol, a value that is clearly identi6ed with
chemical reaction being the rate-limiting step. However, to
centre to a region close to the catalyst surface followed by the left of the graph, there is a fall in activation energy
small di:erences within the surface region. because the rate-limiting step is that of adsorption of reactant
The dimensionless temperature distributions T=T0 are on the catalyst surface.
shown in Fig. 4 at di:erent values of Thiele modulus and The variation of the Thiele modulus for water–gas
e:ectiveness factor as indicated in the curves of this 6gure. shift with bulk temperature is shown in Fig. 8, where the
Fig. 5 shows the overall temperature di:erences at intra- Thiele modulus is seen to increase exponentially with bulk
particle thermal conductivity value of 0.21 W=(mK). It is temperature.
clear that the overall temperature di:erences in dimension- The e:ectiveness factor is expressed by Eq. (16) and
less form are decreased with 9uid temperature. is plotted against the Thiele modulus in Fig. 9. Values of
M. Levent / International Journal of Hydrogen Energy 26 (2001) 551–558 557

Fig. 7. Arrhenius plots for a porous spherical catalyst, the upper


curve(1): shows dependence of the intrinsic reaction rate constant
upon exp(1=T ), the lower curve(2): shows adsorption of reactant
Fig. 10. Concentration distribution in a porous spherical cata-
on the catalyst surface.
lyst at various Thiele modulus values and e:ectiveness factors
(1 = 0:36; 1 = 0:99; 2 = 1:17; 2 = 0:92; 3 = 3:23; 3 = 0:64;
4 =7:98; 4 =0:32; 5 =18:45; 5 =0:15) and at =0:21 W=(mk).

Fig. 8. Thiele modulus versus bulk temperatures, at


 = 0:21 W=(mk).

Fig. 11. External concentration di:erences at various bulk temper-


atures, at  = 0:21 W=(mk).

Pore di:usion limitation becomes more pronounced at



temperatures above 400 C. Lines (4) and (5) in Fig. 10

correspond to 400 and 450 C for which the values of the
Thiele modulus are 7.98 and 18.45, respectively, and the
values of e:ectiveness factor are 0.32 and 0.15, respectively.
The result of this work agrees with the studies of Chinchen
et al. and Twigg [13,20,21]. Fig. 11 shows the external con-
centration di:erences against various bulk temperatures. As
Fig. 9. E:ectiveness factor for a porous spherical catalyst, at seen on this 6gure, the external concentration di:erences
 = 0:21 W=(mk).
are increasing exponentially to a value of 8 with increases
in bulk temperatures. The external concentration di:erences
e:ectiveness factor have been reported in the Catalyst Hand- are not signi6cant at low temperatures since they are less

book [21] for a relatively large catalyst particle size used in than 9:646 × 10−6 kmol=m3 below 250 C.
industrial water–gas shift. Di:usion limitation is more pronounced with an increase
The concentration pro6le within the catalyst body has in temperature as shown in Fig. 12, which clearly shows
been found from Eq. (15) as described previously and Fig. the fall in concentration at temperatures in the range 350 –

10 shows the concentration pro6les against dimensionless 540 C. The intraparticle concentration di:erences increase
◦ ◦
distance r=rs . The second curve at a temperature of 300 C at a bulk temperature of 390 C as shown in Fig. 12, but
with a value of Thiele modulus = 1:17, and e:ectiveness after that, it decreases because of the reduction in surface

factor = 0:92, shows that there is no di:usion limitation at concentration to reach 36.97 at 540 C, when concentration
this temperature. at the centre is 3:1451 × 10−30 .
558 M. Levent / International Journal of Hydrogen Energy 26 (2001) 551–558

Even when GH is low the centre and surface temperature


may di:er appreciably, because catalyst pellet may have a
low thermal conductivity.
The maximum temperature di:erence between surface
and interior is about 4 –9 K at the highest reaction rate, while
the external temperature di:erences between the 9uid and
surface temperatures are of the order of 100 K.
Intraparticle concentration di:erences have been esti-
mated in this work, by 6nding the concentration at the cen-
tre of the catalyst from the Eq. (15) as shown previously.
At low bulk temperatures, the reaction rate constant will be
small, the Thiele modulus will be low, and the e:ectiveness
factor will be high, therefore the intraparticle concentration

Fig. 12. Internal concentration di:erences at various bulk temper- di:erences is small and can be neglected below 265 C.
atures, at  = 0:21 W=(mk).

References

[1] Thiele EW. Ind Engng Chem 1939;31:916.


[2] Damkoehler G. Chem Ing 1939;3:430.
[3] Zeldovich YB. Acta Phys USSR 1939;10:583.
[4] Wheeler A. Adv Catal 1951;III:249.
[5] Weisz PB. Chem Engng Progr Symp Ser 1959;557(25):29.
[6] Wicke E. Chem Ing Tech 1957;29:305.
[7] Mond L, Langer C. British Patent 1888;12:608.
[8] Twigg MV. In: Pearce R, Patterson WR, editors. Catalysis
and chemical processes. London: Leonard Hill, 1981. p.124.
[9] Bohlbro H. An investigation on the kinetics of the conversion
of carbon monoxide with water vapour over iron oxide based
catalysts, 2nd ed. Copenhagen: Gellerup, 1969.
[10] Ruthven DM. Can J Chem Engng 1969;47:327.
Fig. 13. The overall concentration di:erences at various bulk tem- [11] Fott P, Vosolsobe J, Glaser V. Coll Czech Chem Commun
peratures, at  = 0:21 W=(mk). 1979;44:652.
[12] Podolski WF, Kim YG. Ind Engng Chem Proc Des Dev
1974;13:415.
Fig. 13 shows the overall concentration di:erences at var- [13] Chinchen GC, Logan RH, Spencer MS. Appl Catal
ious bulk temperatures above. It is clear that the overall 1984;12:97.

concentration di:erences are becoming constant 390 C bulk [14] Kul’kova NV, Temkin MI. Zh Fiz Chim 1949;23:695.
temperatures above. [15] Hulburt HM, Vasan CDS. AIChE J 1961;7:143.
[16] Kodama S, Fukui K, Tame T, Kinoshita M. Shokubi
1952;8:50.
[17] Glavachek V, Morek M, Korzhinkova M. Kinet Katal
3. Conclusions
1968;9:1107.
[18] Shchibrya GG, Morozov NM, Temkin MI. Kinet Katal
From this result the external temperature di:erences are 1965;6:1057.
large compared with intraparticle temperature di:erences at [19] Campbell JS, Metcalfe SR. Catalyst handbook. London: Wolfe
high 9uid temperature. Heats of reaction have considerable Publishing, 1970, p. 115.
e:ect on the internal temperature di:erences. When the heat [20] Chinchen GC. British Patent 1976, 1,500,089.
of reaction is large, intraparticle temperature gradient may [21] Twigg MV. Catalyst handbook, 2nd ed. London: Wolfe
have larger e:ect on the reaction rate than the concentration. Publishing Ltd., 1989.

You might also like