Você está na página 1de 6

m(R2509/82/030441~303.

0310
Perslmon Press Lid.

ANALYSIS OF SO, OXIDATION IN


NON-ISOTHERMAL CATALYST PELLETS
USING THE DUSTY-GAS MODEL
MARK E. DAVISt, GRAEME FAIRWEATHERS and JOHN YAMANIS*
Department of Chemical Engineering, University of Kentucky, Lexington, KY 40506,U.S.A.
Abstract-Analysis of the experimental rate data[l3] for SO, oxidation on a commercial V,O, catalyst has yielded
a model that describes the experimental data well while the rate constant follows the Arrhenius law. This model
was subsequently used 10 adequately predict experimentally determined effectiveness factors using the dusty-gas
model flux relations and a single value for the tortuosity factor. The latter analysis validates the excellent
properties of the dusty-gas model to properly describe transport in complex systems.

INTRODUCTION

Since the pioneering publications by Thiele[l] and


Zeldovich[2], on the mathematical modeling of diffusion
and reaction in porous, solid catalysts, an enormous
body of literature has emerged, and this body has been
superbly brought together in the classical reference work
by Aris 131.All but a few of these papers to date have been
based on simple, uncoupled flux relations that are
valid, strictly speaking, either for systems in which
Knudsen diffusion is controlling or the reaction is a
simple isomerization or a depolymerization one. Although these simple flux relations may describe a large
number of real systems, these relations do not vigorously
describe the overwhelming majority of real systems that
involve multicomponent reactions and diffusion processes
other than Knudsen flow. The modeling of the latter
phenomena had to await the development of adequate
flux relations that described the diffusion processes.
Following the development of the dusty-gas model for
multicomponent mixturesIB, Jackson[S] and Jackson et
al.&81
for example, as well as other investigators,
Kehoe and Aris[9], Abed and Rinker [lo, 111,for example, have proposed and analyzed mathematical models
that describe multi-component diffusion and reaction in
isothermal and non-isothermal catalyst pellets. These
models are very appealing from the point of view of
solvability and of numerical analysis, and these desirable
characteristics have been demonstrated in several
papers[CS]. Because these models are comprehensive
and solvable, it is important that they be subjected to
further refinements and to testing in order to ascertain
their adequacy in describing the behavior of real systems. Except for the work by Wong and Denny[lZ],
there is no other publication that deals with the testing or
validation of these models which was one of the main
objectives of the present study.

*Author to whom correspondence should be at: Allied Corporation, P.O. Box 1021-R.Morristown,NJ 07960,U.S.A.
tPresent address: Virginia Polytechnic Institute, Blacksburg,
VA. 24061, U.S.A.
SDepartmentof Mathematics.

In order to test the adequacy of the multicomponent


diffusion and reaction models based on the dusty-gas flux
relations, the experimental effectiveness factors for the
oxidation of sulfur dioxide on an industrial V,O, catalyst
as measured by Livbjerg and Villadsen[13] were used.
Since this reaction was studied in the presence of
nitrogen diluent, the above mathematical models were
extended to include the presence of inert components. In
addition, a new reaction rate model was developed that
adequately described the experimental data[I3] which
were obtained using thin pellets.
REACTIONKIhWICS

Many investigators have studied the oxidation of sulfur dioxide with the objective of establishing the kinetics
of the reaction and a large number of models have been
proposed. However, very few of these models show
similarity of functional form and practically none agree
in the values of the model parameters. Weychert and
Urbanek[ 141in a careful evaluation of rate models proposed up to 1968, pinpointed both the inadequacy of
these models and the shortcomings of the experimental
data used to establish them, and this inadequacy persists
in more recent models [15].
Livbjerg and Villadsen[l3] reinvestigated the reaction
with, presumably, the objective of establishing a rate
model in order to subsequently study and model
diffusion and reaction in large catalyst particles. In view
of the wide disagreement of previously proposed models,
undertaking the former task, that is, establishing a
kinetics mode1 for their catalyst and their experimental
conditions was indeed a very sound approach. However,
Livbjerg and Villadsen did what most other investigators
had done before them, i.e. they restricted their study to a
very narrow range of experimental space by choosing to
use a feed stream of constant composition. This restriction resulted in the correlation between the concentrations of the reaction species, thus leading to substantial loss of discriminating power in evaluating previously proposed models.
Livbjerg and Villadsen tested twelve rate models, four
of which fitted the data obtained in the range of 454447

MARK
E. DAVISet al.

448

484C relatively well, and the best of which was the


following model initially proposed by Traina er a1.[16]

where ,kl= Ps,,/KpPmP

g and 1Is = 2. This model was

considered best in the sense that it gave the minimum


relative error for k based on a weighted sum of squares.
However, the estimates of the rate constant, k, did not
obey the Arrhenius law and its values from two different
runs (Series 2 and 3[13]) at the same temperature
differed by about 7%. These discrepancies prompted the
present authors to take a closer look at the data,
numerical values of which were generated by reading off
the graphs presented by Livbjerg and Vifladsen after
suitable magnification. It must be pointed out that estimates of k obtained from the generated data according to
the articles procedure had an average absolute deviation
of 1.3% from the reported values implying that the data
were not significantly different from the original. The
estimates of k at 453.6Care plotted against conversion
in Fig 1, which clearly demonstrates not only the wide
range of the data but also the very strong dependence of
k on conversion. Since the rate constant of an adequate
model must be independent of conversion, these data
lead to the conclusion that the model of eqn (1) is
inadequate.
The above inadequacy of the best model,eqn (1)
with l/s =2, led to testing other models, with particular
attention given to the driving potential term of eqn (1).
Since the partial pressure data were correlated, no
attempt was made to alter the denominator term, and the

jaJ,
3.7

36

following model,

where Kss and Kso, are the functions of temperature


reported by Traina ef ai.1161,was found (a) to bring the
two estimates of kat 453.6Cto within 1.7% of each
other and (b) to yield rate constant estimates that followed the Arrhenius law to a very good degree as Fig. 2
clearly shows. Though a plot of k versus conversion
showed some trend, the range of k was very much
shorter than that of k, eqn (l), and this behavior is
additional evidence in favor of the adequacy of eqn (2) in
describing the available experimental data. The apparent
activation energy for the sulfur dioxide oxidation according to eqn (2) is 14,185cal/mol and the Arrhenius
equation for kis given by:
In k =

2
$*

The model of eqn (2) is of a form that has most


frequently appeared in the literature for the oxidation of
sulfur dioxide and which corresponds to the stoichiometric reaction although it may not account for the
elementary steps of the reaction which have not been
elucidated as yet. So, in summary, eqn (2) is an empirical
model which, however, satisfies the minimum requirements that a kinetics model should fulfill, i.e. good description of the experimental data and a rate constant that
folIows the Arrhenius law, and in view of Urbanek and
Trelas very recent review[lSl this is the only model
proposed to date that possesses these properties.
The presently proposed rate eqn (2) was used to
evaluate the equations describing diffusion and reaction
in the large (6 X6 mm) pellets, although eqn (I) was also
used for comparison purposes.

x-

MATEEMATICAL

MODEL

Consider a mixture of n chemical species the first m of


which participate in a single chemical reaction

35-

2.965- 7139

2 viAi= 0

3.4

3.3

Y
x
xx

I5

x
,s

32-

8
3.1

xx

_
0

m
o-

0.51
1.31

Fig. 1. Dependenceof therate constantof eqn (1)on conversion.


Data[13]: X, series2; 0, series 3.

133

I35

I 31

fx103

(K-l)

Fig.2. Arrheniusplotof rate constant of eqn (2)

SO, oxidation using the dusty-gas

in a porous catalytic particle, the remaining compounds


being inert. In writing the material and energy balances
in the particle it was assumed that the smooth field
description for the fluxes, the concentrations, and the
temperature is adequate and that the pellet takes one of
three shapes, namely infinite slab with parallel plane
faces (thickness Za), infinite cylinder of circular crosssection, and sphere. It was also assumed that external
heat and mass transport resistances are negligible. Then
the continuity equations are:
f&(zNJ=v,ri=l,....nt

model

449

component flux, (7), (11) and (12) along with the boundary conditions
dT
N, =O,z=O,at.z=O

completely specify the system, and their solution leads to


the non-isothermal effectiveness factor, T, using the following equation

(5)

t = (a + l)N,
av,r

$&zNi)=O
i=m+l,...,n

(13)

(6) The above set of equations is more easily solved in its


dimensionless form which is given by:
(7)

where the mass fluxes are given by the dusty-gas model


of Mason et nl.[4] in one dimension, i.e.

and where thermal diffusion and transpiration as well as


heats of mixing have been neglected. Sensitivity analysis
of the CO, methanation reaction[l7] showed that the
viscous flow term does not have any appreciable effects
on pressure variation or effectiveness factor, which is in
agreement with the findings of other investigators[6,18]
for other reaction systems. Based on these results the
last term on the right hand side of eqn (8) has been
dropped from the following developments.
For the inert components, eqn (6) with the appropriate
boundary conditions yields
Ni=O i=mtl,...,n

(9)

while for the simple geometrical shapes under consideration it has been shown[5] that

for i = 1,.

, m- 1

fori=tntl,...,n
$ $@

= #?R/(l t D,e/D;z)

(15)

rl = (a + l)N(l+ 0,=/D;,)
42
dt
N=O,s=O,al=O
n=l.O,f=l.O,_ri=x,i=l

,..., m-l,
WI+1,...,natJ=l

where the dimensionless quantities are defined in the


notation, the Thiele modulus 4 is based on surface
reaction, the Prater number fi is based on total surface
concentration, and both 4 and B are based on a
P dXi Xi dP _ N,
________
reference diffusivity defined by l/D = (l/D, t l/D;,).
RT dz RT dz v,
For the sulfur dioxide reaction studied by Livbjerg and
j*i
Villadsen[l3], the system comprises four components,
+v,
3
(11) three of which are reactive. All the experiments were
i=m+j
0;
performed at one atmosphere total pressure using an
internal recycle reactor under operating conditions that
where = 1, . . . , m, and
eliminated interphase gradients. Under the latter conditions
the values of mole fractions at the external surmA
P
dxi
Xi do _
h
,
----(12) face of the particles are easily calculated from the
--xxix j-1 0;
RTdz
RTdz
material balances, the conversion, and the feed comII. Equations (5), written for the key position which was set at 7 vol.% SO*, 11vol.% O,, and
where =mtl,...,

Equations (9) and (10) were used to express the flux


relations aS follows in terms of component 1:

450

MARKE. DAVISet al.

82 vol.% NT for all the runs. Intraphase diffusion-limited


experiments using 6 x 6 mm cylindrical pellets were carried out at 453.6and 484.W. The set of eqns (15) were
solved using rate models (1) and (2), for comparison
purposes, and effective diffusivities calculated from

D; =; [97OOr,(?7Mi)]
for the Knudsen diffusivity [ 191,and
10~3T-75(1/M,
+ l/Mj)
P[(Z &)il f

I
u&.);3]2

(17)

for the binary diffusivity[20]. For calculating the


effective diffusivities, physical property data reported in
[13] for fresh catalyst were used, namely, pc =
1.33&m, e=O.44, r,,, =3190x Wcm,
and k=
7.0 X 1OWcal/cm/seclC, while 7, the tortuosity factor,
was treated and an unknown parameter which would be
estimated from the data on experimental effectiveness
factors given in the same paper.
NUMERICALSOLUTIONMETHOD

Solution of the boundary value problem defined by the


set of eqns (15) was achieved by the development of a
generalized software package, named EFFECT, in which
the user can specify reaction rate model, chemical species, and parameters to suit his/her reaction system.
Code EFFECT is available upon request from the
authors.
The program EFFECT achieves solution of eqns (15)
by calling COLSYS[Zl-231 which is a code designed to
handle boundary value problems for mixed order systems of ordinary differential equations. The latter code is
based on spline collocation at Gaussian nodes, the piecewise polynomial approximation being given in terms of
B-splines. Collocations sound theoretical footing has
resulted in the derivation of effective algorithms for error
estimation and adaptive mesh refinements, which are
implemented in COLSYS. The problem is solved on a
sequence of discrete meshes until user-specified error
tolerances are satisfied. For the case of this paper, the
error tolerances were taken to be 10m4.This choice was
based on the results of preliminary experiments [ 171.All
computations reported in this paper were performed in
double precision on an IBM-370-165computer.

latter variables on transport and reaction. It should,


therefore, be expected that a single value of the tortuosity factor, 7, the only unknown parameter, would result
in adequate description of all the experimental data
(Series 2, 3, and 5) for the large particles, provided the
catalyst pellets had not undergone some drastic change
in chemical or physical properties.
The experimentally determined effectiveness factors
for Series 2 and 3 were determined at the same bulk
temperature and overlapping, for most part, conversion
range and found to have significantly different values. In
particular, the factors of Series 3 were significantly lower
than the factors of Series 2, and of the same order of
magnitude as those of Series 5, the latter having been
determined at temperature 30C higher. Since the
effectiveness factor at lower temperature must be higher
than those at higher temperature, all else remaining the
same, the data for the large particles of Series 3 indicate
that they had been affected by unknown factors and thus
were not used in evaluating the adequacy of eqns (15).
Based on the above arguments, it was decided to
estimate the tortuosity factor, 7, from one experimentally
determined effectiveness factor, namely, that of Series 2
at 50% conversion. The estimate of 2.70 was obtained by
requiring the effectiveness factor predicted by eqns (IS)
to match the experimental value. Though this estimate of
7 is not the best in a least-squares sense, it should still be
expected to describe the experimental data well, provided that the model is adequate and the data have been
obtained from catalyst of the same physical and chemical
quality. Figure 3, shows experimentally determined and
predicted, using eqns (2) and (IS), effectiveness factors
for 453.6 and 484C, with the tortuosity factor kept
constant at 2.70. The close agreement between the
experimental points and the predicted curves is rather
striking, especially striking if it is kept in mind that the
average pore size, r,,,,the porosity, @,of the particles, and
the tortuosity factor have been kept constant, although
these properties should, on physical grounds, be expected to vary for different catalyst batches and they do [13].
Figure 4 shows typical component profiles in the
pellet. In all the simulations, the total pressure in the

~r-----l

OB

RFSULTSAND DISCUSSION

The mean-free path of SO, at the experimental conditions explored by Livbjerg and Villadsen (454-48OC)is
of the order of magnitude of the average pore radius of
the catalyst used by the above investigators, and, therefore, transport into the large pellets was almost equally
affected by molecule to wall collisions as well as molecule to molecule collisions, and these processes are
correctly reflected by the model of eqns (15). These
equations account for Knudsen streaming and bulk
diffusion as well as for pressure, temperature and composition variation in the pellets, and the effects of the

o-0

02

0.4

0.6

08

ConversionI

ID

Fig.3. Plot of predicted(solidcurves)effectivenessfactor.Data


[131: A, series2: 0, series5.

SO, oxidation using the dusty-gas model

particles was almost constant, being in the range of 0.97


to 1atm, and the temperature rise was less than 3C.
Figure 5 shows a plot of effectiveness factor vs Thiele
modulus with conversion as parameter at bulk temperature of 453.6C. It should be noted that the low
effectiveness factor at low Thiele modulus is due to the
fact that the modulus has been defined in terms of
reaction rate at the external surface. Approximately
60 sec. of CPU time was spent in generating the data of
Fig. 5, which shows how modest the computer time
requirements in solving the set of eqns (15) is. However,
this is not always true. At high conversions and high
Thiele modulus, the system of equations becomes stiff
and the CPU demand becomes high, and this is the
reason why the predicted effectiveness factor curves in
Fig. 3 based on rate equation (2) were not extended
beyond 90% conversion. Other computer runs using rate
equation (1) predicted the same behavior as in Figs. 3-5,
on the scales of these figures.
In summary, the close agreement between the
experimentally determined and the predicted effectiveness factor is strong evidence that the mathematical
model given by eqns (15) is indeed very adequate in
describing the oxidation of sulfur dioxide in an industrial

451

catalyst, and this evidence, in turn, validates the


excellent properties of the dusty-gas flux relations to
properly describe transport in complex systems.
Acknowledgement-Ackwledgement
is made to Anne Leigh of
the University of Kentucky Computing Center for valuable
assistance in the debugging of EFFECT.

NOTATION

reaction component
radius of the cylindrical pellet, spherical pellet,
half the thickness of the plate shaped pellet, cm
permeability of the pellet, cm2
effective Knudsen diffusivity of i, cmlsec
effective binary diffusivity of i in j, cm/sec
(l/D, + l/D;*)
heat of reaction, cal/gmole
rate constant, eqn (l), gmolelseclg catlatm
rate constant, eqn (2). gmolelseclg cat/atm2
effective thermal conductivity, cal/cm/sec/Y
(D,)lDi

molecular weight of i, g/mole


number of reacting components
total number of components rt 2 m
molar flux of i, gmol/cm/sec
aN,RT/v,DIP dimensionless

20,

key flux

total pressure, partial pressure of ;, atm


universal gas constant
mean pore radius, cm
reaction rate function = pcrso2,gmole/cm3/sec
reaction rate function of SO1, gmolelgcatlsec
rjr dimensionless reaction rate
absolute temperature, K
T/T dimensionless temperature
atomic diffusion volume of i
mole fraction of component i
space coordinate, cm
Greek symbols

IO

05
Dlmenslonless

Radius. z

Plots of predicted component profiles in a catalyst pellet


at 70% conversion, 726.6 K and Thiele modulus equal to 0.4.

Fig. 4.

geometric factor:
sphere = 2

for slab = 0; cylinder = 1;

AH,DP/R (TI)k, Prater Number

DW;,)
effectiveness factor, calculated, experimental
porosity of pellet
viscosity
stoichiometric coefficient of i
P/P dimensionless pressure
density of the catalyst, g/cm
tortuosity factor
a[rR7j PD)], Thiele Modulus
Superscripts
0 at reactor

feed conditions
s at pellet external surface conditions

O,lU

REFERENCES

0.1

Thiele

Modulus, .+

Fig. 5. Plots of predicted (solid curves) effectiveness factors at


different conversions: 1at 70%; 2 at 50% 3 at 30%; 0, data from
1131.

[I] Thiele E. W., Ind. Engng Chem. 1939 31 916.


(21 Zeldovich Ia. B., Zhur. Fiz. Khim. 1931 13 163.
[31 Aris R., The Mathematical Theory of Diffusion and
Reaction in Permeable Catalysts, Voi. 1. Clareidon Press,
Oxford 1975.

MAFXE. DAVISet al.

452

[4] Mason E. A., Malinauskas A. P. and Evans R. B. III, I.


C/rem. Phys. 1967 46 3199 and Refs. therefrom.
[S] Jackson R., Transport in Porous Catalysrs. Elsevier, Amsterdam 1977.
[6] Hite R. H. and Jackson R., Chem. Engng Sci. 1976 32 703.
[7] Kaxa K. R. and Jackson R., Chem. Engng Sci 1980 35 1179.
[S] Kaxa K. R., Villadsen J. and Jackson R., C/rem. Engng Sci.
1980 35 17.
[9] Kehoe J. P. G. and Aris R., C/rem. Engng Sci. 1973 28 2094.
[lo] Abed R. and Rinker R. G., A.LCh.E.I. 1973 19 618.
[ll] Abed R. and Rinker R. G., A.Kh.E.J. 1974 26 391.
[12] Wong R. L. and Denny V. E., Chem. Engng Sci. 197530 709.
[I31 Livbjerg H. and Villadsen I., Chem. Engng Sci. 1972 27 21.
[14] Weychert S. and Urbanek A., Inf. Chem. Engng 1968 23
1257.
[IS] Urbanek A. and Trela M., Coral. Rev.-Sci. Engng 1980 21
13.

I161 Traina F., Cucchetto hf., Cappelli A., Collins A. and Dente
A., Chim. & Ind. 1970 52 329.
I171 Davis hf. E., Ph.D. Dissertation, University of Kentucky,
Lexington 1981.
1181 Haynes H. W. Jr., Can. I. C/rem. Engng 1978 56 582.
(191 Smith I. M., Chemical Engineering Kinetics. McGraw-Hill,
New York 1970.
I201 Sherwood T. K., Reid R., and Prausnitz J., Properties of
Liquids ond Cues. McGraw-Hill, New York 1977.
I211 Ascher U., Christiansen J. and Russell R. D., Math. Camp.
1979 33 659.
1221 Ascher U., Christiansen J., and Russell R. D., Springer
Lecture Nofes Computer Science, Vol. 76, 1979.
1231 Ascher, U., J. Camp. Phys. 1980 34 401.

Você também pode gostar