You are on page 1of 12

Chemical Engineering and Processing, 32 (1993) 233-244 233

Dynamic modelling of industrial SO, oxidation reactors.


Part II. Model of a reverse-flow reactor

K. Gosiewski
Institute of Inorganic Chemistry, ul. Sowiliskiego 11, 44-101 GIiwice (Poland)

(Received January 27, 1993)

Abstract

This paper presents the results of a computer simulation of a reverse-flow reactor. The results are compared with
experimental data quoted by other workers. The necessity to take into account gas-to-catalyst heat transfer is
stressed, so that quantitatively correct predictions may he obtained using the model. A brief description is given of
the way flow reversal has been incorporated in the model, and the possibility of the accumulation of errors occurring
during the calculations is mentioned. It is shown that the one-dimensional heterogeneous model of the bed makes
it possible to analyze the process with an accuracy sufficient for the preliminary selection of the parameters of the
reactor, whereas the homogeneous model leads to discrepancies which are too high. An analysis of the parametric
sensitivity of the model enables the formulation of practical conclusions concerning the design of such reactors.

Introduction 1. Reactors with periodically forced changes in the inlet


stream parameters (most often the reactant concen-
In Part I [la] of this series, equations have been trations, but also pressures and temperatures).
presented for a model involving a single pass of the 2. Reactors with periodical changes in the direction of
catalytic SOZ oxidation reactor with a fixed bed. Next, flow through the bed, the so-called ‘reverse-flow
verification of the model was given on the basis of a reactors’.
comparison between the simulated and actual transients,
both for a single pass and for a whole SO, oxidation unit. The literature concerning this problem is extensive.
In reactors for stationary SO1 oxidation, the dynamic Thus, even in 1984, a review paper by Renken [ lb]
states only occur in certain special cases and are there- provided a long list of references. Apparently, the
fore only of marginal importance in process engineer- above two groups of reactors involve similar phenom-
ing. Such states include the various cases involved in ena, the former being mainly concerned with non-
the start-up of these installations (cf. Part I). Other stationary catalysis. To describe this properly, it is
dynamic states which may be of interest in an analysis necessary to analyze the phenomena occurring on a
of the process result from the reaction of industrial microscale in a single pellet. However, as shown in the
metallurgical installations to the time-dependent oper- literature, the improvement in the extent of conversion
ating conditions. This concerns, in particular, the reac- achieved by the whole reactor in this group does not
tion of the process to the drop in concentration of SOZ usually exceed a few per cent. Some cases are also
in the process gas which frequently occurs when the mentioned where the overall effect may even be nega-
oxidation unit cooperates with copper smelter convert- tive. Despite a number of references (see review article
ers. These cases are, however, similar to the ‘hot’ start- [lb]), this mode of operation has not led to any serious
ups discussed in detail in Part I. industrial applications.
A separate problem is that of a reactor operating under The other group (reverse-flow reactors with numer-
a continuously forced non-stationary state. Analyses of ous industrial applications - cf. Matros [2,3c, 4,5],
such set-ups, and especially their design, require some Wojciechowski et al. [6, 71) includes apparatus in which
reliable tools to describe their dynamics. In this case, the the direction of flow is periodically reversed. Such
dynamic states are no longer just an episode in the arrangements involve the combination of a reactor and
operation of the installation, but constitute the normal a heat regenerator. A ‘cycle’ in such reverse-flow reac-
operating regime. Reactors operating under conditions tors corresponds to the time during which the reaction
of forced non-stationary state conditions may be roughly medium flows through the reactor in both directions.
sub-divided into the following groups: The time of flow in one direction is called a ‘half-cycle’.

02552701/93/$6.00 0 1993 - Elsevier Sequoia. All rights reserved


234

Hence, it is a reactor involving an internal autothermal reactor, by an appropriate choice of model parameters
recovery of heat along its regeneration path. This is the (and also, of the reactor parameters) will be dealt
classification which has been most frequently used in with when the parametric sensitivity of the model is
recent publications (see, for example, Eigenberger and analyzed.
Nieken [ 81, Matros [ 3c] and Froment [3b]). In a similar fashion to the simulation of the dynamic
Froment [3b] claims that dynamic surface phenom- states involved in the start-ups of SO, oxidation units,
ena are not of crucial importance if the reversal cycle is the intention was to compare the present simulation
longer than 20 min. From a number of theoretical results with experimental data obtained in an industrial
studies and experiments performed in actual reactors by reactor. However, a principal difficulty was the lack of
Matros et al. [2, 3c, 4, 51, it has been shown that this any experimental data of our own. A large number of
condition is usually fulfilled in the oxidation of SOa, such data are given, however, in articles based on work
even with low SO2 concentrations in the inlet gas. The undertaken in the Institute of Catalysis at Novosibirsk,
important results obtained with such reactors arise e.g. refs. 2, 3a and 15, and concern both laboratory
from the heat of reaction generated, coupled with heat experiments, pilot-plant tests (at a flow rate of c.
accumulation inside and heat transport within the bed. 3000 SCM h-l) and several large-scale industrial instal-
At low reactant concentration, the reactor may be even lations with flow rates of 30 000-40 000 SCM h-’ [3a].
treated roughly as an ordinary heat regenerator Hence, it was decided to use the data cited in the above
(Gosiewski and Sztaba [ 3d]). Thus, the model described publications in testing our model. Unfortunately, for
in Part I may be useful in analyzing phenomena taking the industrial reactors, it turned out that the number of
place in a reverse-flow reactor. Similar models, al- numerical parameters (e.g. the geometry of the reactors,
though sometimes differing in important details, have the volume of catalyst) was too small to justify any
been employed by authors dealing with the mathemati- serious model calculations. However, the laboratory [2]
cal description of such reactors (see, for example, refs. and pilot-plant [3a, 151 experiments were described in
2, 3b, c, e, 4, 5 and S-13). much greater detail and could therefore be employed in
Indeed the problem has been discussed in a large simulated testings of the model.
number of papers, far outnumbering those concerned
with the start-up simulation investigated earlier. Thus,
the question arises whether it is worthwhile to carry out Discussion of the model of the reverse-flow
an analysis of this problem. It appears that of the reactor employed
numerous authors cited, only Eigenberger and Nieken
[8] have made any serious attempt to investigate the General description of the model
effect of the model pafametet-s on the temperature and For the purpose of numerical analysis, a mathemati-
conversion profiles along the bed in a systematic man- cal model has been formulated which would make it
ner. Their studies were restricted, however, to a homo- possible to simulate all of the variants of such a reactor
geneous model. In what follows it will be shown that which are employed practically. The block diagram of a
the effect of the gas-to-catalyst heat transfer is most ‘generalized’ reverse-flow reactor is shown in Fig. 1.
important and should not be nbglected. The diagram contains four fixed-bed passes; heat ex-
The object of the simuJations described in subsequent change occurs in the middle of the reaction zone. Two
sections of this paper is to determine the extent to passes (the first and the last) may be packed with the
which the description, of the bed may be simplified inert bed and act as heat regenerators. The middle
in modelling reverse-flow reactors. In the design of passes (2 and 3) are packed with the catalyst and form
such reactors for strongly exothermal reactions (e.g the active zone of the reactor.
the oxidation of SOZ), it is crucial to select reactor Heat transfer may be produced by the supply of heat
parameters in such a way that values of the maximum (heating the reacting medium) for very low concentra-
temperature, r,, , which are too high are avoided. tions of reactants, or by the withdrawal of heat (cooling)
This problem has been discussed at length by Matros for moderately high concentrations. For intermediate
(see, for example, ref. 2). In the interpretation of the concentration values, when there is no heat exchanger,
effect of the model parameters on the value of T,,,,, the model assumes zero heat exchange in the middle of
insufficient attention appears to have been paid to the the bed.
intensity of the gas-to-pellet heat transfer. Matros [2] The block diagram of the model presented in Fig. 1
has pointed frequently to the effect of the axial conduc- provides a particular generalization which makes it
tion of heat in the bed on. the value of T,,;, but has possible to evaluate a reverse-flow reactor for all possi-
neglected the even more important dependence on gas- ble variants, i.e. both ‘with’ and ‘without’ the inert bed
to-pellet heat transfer. The ‘possibility of influencing at the reactor inlet and outlet, and also ‘with’ or
T max and, consequently, the average conversion in the ‘without’ cooling or heating in the middle of the bed.
235

4 For t = 0:

&
v
1
Tk(X, 0) = T,(x, 0) = f(x) (4)
PASS 1
and to the simplified boundary conditions:
For x=0:

+=(I-&).a.[T,(O,t)-T,.] (5)

T.(O, t) = T@(t) and ~(0, t) = co(t) (6)


For x = L:
82-k
-_=O \ ,
ax
Some authors (see, for example, Eigenberger and
Nieken [Xl) use a homogeneous model to describe the
reverse-flow reactor. Thus, in order to assess the effect
of this simplification the following simplified form of
Fig. 1. Generalized block diagram of the model of a reverse-flow the differential heat balance was tested.
reactor. It was assumed that:
T(x, t) = T*(x, t) = Tk(X, t)
Model of the bed (catalyst or inert)
and consequently eqn. (8) was substituted for eqns. (1)
As a dynamic description of the bed, the heteroge-
and (2) in the model:
neous model has been employed (cf. Part I). The model
has been supplemented, however, with an additional
term allowing for the heat loss from the catalyst to the
surroundings. The introduction of this term arises from
The boundary conditions for this simplification as-
the fact that the simulations were also of concern to a
sume the following form:
small-scale laboratory reactor. Whereas in industrial or
even pilot reactors with diameters of several meters this For x = 0:
loss may be neglected, for the laboratory reactor de- and ~(0, t) = c,,(t)
T(0, t) = T,,(t) (9)
scribed by Matros [2] reliable results could not be
obtained unless the heat losses through the wall were For x = L:
taken into account. dT
For the sake of simplicity, it was assumed that heat -_=O (10)
ax
transfer to the environment occurs only from the cata-
lyst. The inclusion of such heat losses in the gas-phase with the initial condition:
heat balance has no appreciable effect on the numerical For t = 0:
results. The term which relates to heat transfer with the
surroundings [bold type in eqn. (2)] is the only modifi- T(x) = f(x) (11)
cation made with respect to the model discussed in Part The model for an inert bed differs from that for an
I of this series.
active (catalytic) bed only in the value of the reaction
rate, r, which is set equal to zero in the former case.
n,.C~.~+a.S.(T,-_T,)=O (1) The discussion of an expression describing the reac-
tion rate has been given in Part I of this series.
3%
e.---P,.C,.~+~.S.(7,-T,)+H.r.p,+
a

ax*
Model of interpass heat exchange
-2 .(Tk_ *.> =o
The interpass heat exchange is modelled as the with-
.ac+r.p,=o
ngax
drawal of a given amount of heat in a counter-current
heat exchanger disregarding the dynamics (this ap-
Equations (1) -( 3) are subject to the following initial proach is employed in the simulations described in later
conditions: sections of this paper). This amount can also be given
236

a priori in the model (options) as the temperature or 2. Rewriting the parameter matrices (PAR) describing
enthalpy difference between the stream leaving pass 2 the passes according to the following formula:
and that entering pass 3.
PAR( 5 - i) = PAR(i) (13)
The reactor bed has a much larger heat capacity than
a (usually small) heat exchanger or even the electric 3. Rewriting the matrices of the parameters of the
heater used by Wojciechowski [7] in his reactors for the streams (STREAM 1) for streams 2-6 at the moment
catalytic combustion of organic substances. This may of reversal (new boundary conditions), according to:
provide the justification for neglecting the dynamics of STREAM l(m) = STREAM l( 7 - m) (14)
the heat exchanger in the model of the overall reactor,
in contrast to the approach employed in discussing the where:
start-ups of the SO, oxidation unit. An additional m = number of stream (2 to 6)
argument for assuming cooling without any analysis of
Following the “turning” of the reactor, the number-
the thermal dynamics of the heat exchanger is the lack
ing of the passes and streams is changed, while the
in the references mentioned (e.g. refs. 2 and 15) of any
direction of integration is apparently unchanged.
data which could serve as a basis for evaluating the
dynamics of the heat exchanger. Particular numerical properties of the reverse--ow
reactor
Numerical method of solution of the overall model; The “turning” of the reactor described in the foregoing
incorporation of flow reversal section introduces a particular positive feedback into the
For the simultaneous solution of the equations de- numerical calculations. The feedback arises from the fact
scribing the overall model, the same numerical method that the parameters of the streams, as calculated at the
was employed as that used in the dynamic flowsheeting end of the preceding half-cycle, are always burdened with
system DYNAM outlined briefly in the previous part of a certain calculational error 6. This error is a sum of a
this series. numerical error a,,, and a model error dmodwhich result,
The essential modification in the model for the whole for example, from all the necessary simplifications in-
reactor compared with the original DYNAM simulator volved in the mathematical description. The error may
consists in taking the flow reversal into account. This be expressed by the following sum:
modification has already been introduced at the level of
@,i=6i,i num + G&i (IS)
the numerical model, which is based on difference equa-
tions and on a numerical formulation of the initial where
conditions [eqn. (4)].
j = generalized number of parameter (in stream or at
From the numerical point of view, flow reversal
a mesh point in the bed)
constitutes a special case of discontinuity in the initial
i = number of mesh point along the bed
and boundary conditions and requires inversion in the
direction of their integration. To simplify the numerical In models without flow reversal, such as those for
procedure the reversal is modelled by “turning the unsteady states during start-ups of the SO, oxidation
reactor upside down”, with the directions of flow and plant, these errors would simply add to the final error
integration kept unchanged. Also, the numbering of of simulation. The situation for models for reverse-flow
passes in the model is reversed, as they are always reactors is, however, different. At the moment of rever-
numbered along the direction of flow. sal, these errors distort both the initial conditions [eqn.
Such modelling of flow reversal requires: (12)] and the boundary conditions [eqn. (14)] for the
next half-cycle of calculations.
1. Inversion of the ‘states’ of the passes at the moment
This phenomenon is a numerical positive feedback,
of reversal (matrix STATE 1). If the states of the passes
and although it does not necessarily lead to the instabil-
are determined unequivocally by the temperature
ity of the method, it may result in much higher para-
profile along a pass, this operation rewrites the temper-
metric sensitivity of the model, since the errors in the
atures calculated at the end of the preceding half-cycle
subsequent computational cycles accumulate, in a simi-
(new initial conditions) according to the following for-
lar fashion to the accumulation of heat in the reactor,
mula:
producing a so called ‘heat wave’.
T,(5 - i,j) = Tk(i, n -j + 1) (12) These phenomena may be responsible for (amongst
other things) differences in the numerical results depend-
where:
ing on the initial temperature profile along the reactor,
i = number of pass and thus, the number of cycles necessary to reach a state
i = actual number of mesh point along pass in which the profiles for two consecutive cycles do not
n = total number of mesh points. differ from each other (this state will subsequently
237

be called a ‘quasi-stationary’ state). Consequently, the where


numerical accuracy of the mathematical models of the c. -H
bed employed in describing a reverse-flow reactor has to AT,,=“- * J&v (17)
be higher than in modelling other unsteady states where c,
the above feedback does not occur. A similar accumula- with the average conversion described as follows:
tion of numerical errors has been observed in simulating t+r,/2

unsteady states for metallurgical installations containing


x,, = f - X(t) - dt (18)
two external loops of autothermal heat recovery, and c s
thus positive feedback loops as well. t
The need to fulfill this condition has been stressed by
‘Quasi-stationary’ state and a test for the consistency Matros [2]; however, he does not seem to have taken it
of heat balance into account too seriously. In other papers dealing with
Despite the fact that the reactor operates in a contin- reverse-flow reactors, no mention has been made of
uously forced non-stationary state, the arbitrarily as- the necessity to meet such a requirement by a model
sumed initial condition of eqn. (4) leads, after some describing a reverse-flow reactor operating in the quasi-
time, to the reactor operating in a state which may be stationary state. The author’s own experience with a
called ‘quasi-stationary’. This state arises when the large number of numerical models for these reactors
transients in consecutive cycles do not differ from one suggests that the rigorous fulfillment of this condition is
another. Such a definition has been assumed by Eigen- extremely difficult due to the accumulation of the nu-
berger and Nieken [8], for example, who called this merical errors mentioned earlier. Computational prac-
state ‘stationary’. Although the definition seems quite tice shows that if S,, is less than 10% of AT,,, the
obvious, the name ‘quasi-stationary’ appears to be numerical results provided by the model may form a
more appropriate as the system is permanently dis- sound basis for a discussion concerning the heat bal-
turbed by the reversal of flow. ance of the reactor. If, however, this value increases to
The approach to the quasi-stationary state from an several tens of per cent, any model-based conclusions
arbitrarily assumed initial state requires a large number have to be treated with caution. It must be pointed out
of cycles. This number depends on how much the that a,,,, is, in a sense, a measure of the amount of ‘false’
assumed initial temperature profile differs from that at heat which accumulates in the model due to the accumu-
the beginning of a cycle in the quasi-stationary state. In lation of errors inherent in the model and/or numerical
the simulations performed, this number varied from errors [see eqn. ( 15)] rather than because of the physical
several to c. 20 cycles. The stability of the method may regeneration of heat within the bed. If the mathematical
be tested by checking whether substantially different model accumulates more or less heat than the physical
initial states lead to the same quasi-stationary state. This heat balance for the reactor would suggest, then any
is a basic test for the numerical method employed. If a attempt at reaching practical conclusions may be risky.
considerable accumulation of the processing errors actu- In the simulations discussed below, the error 6,,, in
ally occurred, the result for the final quasi-stationary state most cases, did not exceed a few per cent of ATad.
would depend strongly on the number of the cycles neces-
sary to attain this state. After the details of the model
had been perfected, no such phenomenon was found in Simulation results and discussion of the parametric
the simulations performed. This, however, is simply a test sensitivity of the model
of the numerical accuracy of the model and does not
constitute a proof of its physico-chemical correctness. A detailed discussion of the effect of the various
Another important test for both physico-chemical parameters employed in the model on the temperature
and numerical correctness is a procedure which will be profiles in the quasi-stationary state has been presented
called the ‘heat balance consistency test’. An alternative by Eigenberger and Nieken [8]. This effect was investi-
definition of the quasi-stationary state is that it is a gated using a mathematical model and then compared
state in which the first principle of thermodynamics is with experimental results obtained in a laboratory reac-
fulfilled or, in other words,’ when the overall heat tor in which propane and methane underwent combus-
balance for the reactor is satisfied during a half-cycle. If tion on a commercial catalyst to allow purification of
there is no heat loss and/or heat exchange in the middle automotive exhaust gases. The reactor was 345 mm in
of the reactor, this may be written as follows: length and had an internal diameter of 50mm. The
* f &I2 article stressed the problem of the considerable losses of
heat through the walls of such a reactor. In the case
Bhb=Arti- ; T,,,(t) - dt - Ti, =0 (16)
( C s > discussed by Eigenberger and Nieken, this problem was
I overcome by applying additional compensation heating
238

through the external wall of the reactor. Good qualita- Poland. For k in Boreskov’s equation in the form
tive agreement was obtained between the actual profiles presented in Part I of this series, the following experi-
and those predicted by the model. With respect to the mental formula (see ref. 20) was applied:
quantitative agreement, the authors themselves admit-
ted that it was largely dependent on the correct level for
k = K, x exp (-4.95 + 15.0 x tin)
the compensation heating. The reactor operated at rela- I
tively short reversal cycles (60-90 s).
As has already been mentioned, a wealth of experi-
mental data obtained in the Institute of Catalysis at
Novosibirsk concerning the oxidation of SO, has been
presented in refs. 2, 3a and 15, to mention only a few.
A particularly large number of the experimental results
obtained in a pilot plant have been reported in the
paper of Bunimovich et af. [3a]; the experimental data
have been compared with theoretical predictions. The
monograph by Matros [2] quotes, however, several
vastly differing models, and it is difficult to guess which In the case discussed, the value of & in formula (19)
of them was actually employed for comparative pur- was taken as being that for moderately used Polish
poses in the calculations used by the Novosibirsk catalysts; luckily, sufficient agreement was obtained de-
group. From personal discussions with the authors, and spite the fact that the experiments cited in ref. 3a were
also from some comments presented in ref. 2, it may be carried out using an unknown Russian catalyst.
inferred that the majority of the calculations concerning The numerical values of the model parameters are
Tnlax and the degree of conversion were undertaken given in Table 1. The values of GLwere estimated using
using a simplified model assuming a so-called ‘sliding the formulae given in refs. 16 and 17, whereas those of
regime’ in the form of ordinary differential equations. I, were based on refs. 18 and 19.
Such a model, even if it leads to a satisfactory agree-
ment between the parameters mentioned, is incapable An analysis of the parametric sensitivity of the model;
of predicting either the transients or the temperature a comparison between the heterogeneous and
and conversion profiles along the bed. homog?eous models
In ref. 3a, values have been given of the average
Numerical values of the parameters used in conversion in a cycle X,, and of the maximum temper-
the simulation ature T_, for a number of the operating regimes of the
To determine the numerical values of the parameters pilot plant. The parameters of this plant have been
employed in the mathematical modelling, it is necessary described both in refs. 3a and 15. The number of data
to have a certain amount of information concerning presented in these papers is sufficient for the purpose of
both the experimental reactor and the catalyst used. mathematical modelling of the process.
The data for the laboratory reactor were taken from To test the model, the operation of the plant was
ref. 2, whereas those on the pilot reactor were from refs. simulated for an inlet SO, concentration of 3.2%. The
3a and 15. The author’s experience with the static model parameters used in calculations for this variant
mathematical models for industrial reactors shows that are those in Table 1 corresponding to a flow rate of
a kinetic equation describing the reaction rate over a 3000 mz h-’ the variant itself being termed a ‘basic
given type of commercial vanadium catalyst requires a variant’.
certain identifuzation of the equation to kinetic experi- Calculations were undertaken to test the parametric
ments on this particular catalyst [20]. The problem of sensitivity of the model and were carried out neglecting
the kinetic equation has been briefly discussed in Part I the heat losses to the surroundings. A comparison with
of this series. For the catalyst described in ref. 3a, no the experimental values presented in ref. 3a shows,
such data were available. This is certainly a weak point however, that such losses must have actually occurred
in the simulations performed. As follows from the (TM, predicted was higher than the experimental
studies, of Eigenberger (cf. ref. 8) and also from the value). Calculations for the ‘basic variant’ data, but
results presented below, the influence of the catalyst with heat losses of c. 5% of the overall amount of heat
kinetic parameters is most important; hence, a correct generated, led to good agreement (h, = 1 W m-i K-l;
determination of the reaction rate is essential. In the T, = 10 “C).
calculations, an equation has been assumed and fitted Next a number of parameters which could have a
(identification according to the method given in ref. 20) considerable effect on X,, and T,,,,, were varied sub-
to a vanadium catalyst produced by PZChem Lubon in stantially in order to study the parametric sensitivity of
239

TABLE 1. Numerical values of the basic parameters of the model used in the simulation

Parameter Unit Simulation variable

1.3% so, 3.2% SO, 9.4% so,

Pilot act. [ 151 Laboratory act. [2] Pilot act. [3a] Pilot act. .[ 151

Flow rate m3 h-’ 2500 26 3000 2750


(under standard
conditions)
Inlet gas vol. % SO, 1.3 SO, 3.2 SO, 3.2 SO, 9.4
composition 02 10.0 02 12.5 02 12.5 0, 11.6
N, 88.7 N, 84.3 N* 84.3 N* 79:o
Reactor diameter m 1.55 0.175 2.0 2.8
Total bed height m 3.7 1.3 4.4 2.0
Catalyst pellet size mm cylinders cylinders cylinders cylinders
25 x 50 x 6 8x8x3 25 x 50 x 6 25 x 50 x 6
Heat transfer Wm-2K-’ 78 96 53 78
coefficient, a
Specific surface, S m* m-3 380 500 380 380
Effective conductivity, 1, W mm’ K-’ 1.0 0.5 1.0 0.833
Effective specific heat kJ kg-’ K-’ 3.5 2.5 3.5 3.5
of catalysis, Ck,cff
Bulk density of kg mm3 500 500 500 500
catalyst, p,
Heat of reaction, H kJ kmol-’ K-i 8.89 x IO” 8.89 x 104 8.89 x 104 8.89 x 104
Cycle time, 2, min 30 50 50 60

the model. In addition, simulation was carried out two-fold increase in this coefficient produces a visible
employing the homogeneous model, i.e. that which rise in the maximum temperature T,,,,, and a strong
completely neglects the gas-to-catalyst heat transfer re- drop in the average conversion X,,.
sistance. The results are listed in Table 2. In principle, The omission of the resistance to heat transfer from
they support the conclusions presented by Eigenberger the homogeneous model leads to a considerable dis-
and Nieken [S], especially regarding the effect of the crepancy between the values calculated and those found
reaction rate constant, which is difficult to understand experimentally. Additionally, it has been found that the
on purely intuitive grounds (the lower the value of K,,, homogeneous model is more. sensitive to AC than the
the higher the maximum temperature and the lower the heterogeneous model; thus, I the importance of this
average conversion X,,). An analogous conclusion con- parameter is sometimes over-estimated in cases where
cerns the effect of substituting the inert bed for the inlet the homogeneous model is employed. The predicted
and outlet sections of the active catalytic bed. Substitu- values of the temperature profile for the three different
tion of a certain volume of catalyst with an inert lengths of bed are shown in Fig. 2. For a bed of
material also leads to a substantial increase in T,,,,, and sufficient length, a flat ‘crest’ to the heat wave is formed
a decrease in X,,. Similarly as in ref. 8, a moderate along which no reaction occurs. A change in length
effect has been found for the axial dispersion which, in over a relatively wide range affects only the width of the
the case discussed, is represented by the effective heat heat wave, without any modification in the parameter
conductivity of the bed, 2,. A small influence has also values calculated.
been found for the following parameters: (i) inlet gas In the case studied, it was possible to shorten the bed
temperature, Tin; (ii) heat capacity of the bed (it affects down to c. 25% of the ‘basic’ value, virtually without any
only the velocity of movement of the heat wave; thus, effect on T,,,, and X,,. Further contraction of the bed
for a sufficiently long bed, it does not influence T,, or to 20% of the initial value (without changing the dura-
X,,); (iii) length of the bed (provided it is sufficiently tion of the cycle) led to the reaction being extinguished.
long); and (iv) duration of the cycle (provided it is not The computed conversion profiles along the bed for
too long, otherwise the reaction could be extinguished). the ‘basic variant’ as well as for the bed shortened to
An important conclusion of the analysis is the strong 50% and 25% are shown in Fig. 3. The first increase in
effect of the heat-transfer coefficient into the pellets, ~1. conversion is visible in that part of the bed correspond-
This effect has not been discussed either in ref. 8, where ing to the rising segment of the ‘heat wave’, while the
only the homogeneous model was used, or in ref. 2. A second (final) increase corresponds to the declining
240

TABLE 2. Values of X,, and T,, predicted by the model for


various values of the parameters relative to the ‘basic variant’

No. Description of the variant X” (-) T,, ( “C)


Heterogeneous model eqns. (I), (2) and (3)
1 Basic variant 0.92 575 s
‘ii 0.6
2 Basic variant, 0.93 540-560 ;
:
heat losses -5%
8
3 1, = 0 0.91 581 0.4
4 1, = 5 x n,, 0.94 567
=5Wm-‘K-1
0.2
5 Ti:,.= 250 “C 0.92 577
6 K” = 1/2K, 0.86 620
7 C, = 1.0 kJ kg-’ K-’ 0.92 590 0.0
8 L = 1/4L, 0.92 570 xc-1
= 1.1 In 2.2 4.4 L cm

9 a=2xa, 0.88 612 Fig. 3. Calculated conversion profiles along the bed for the ‘basic
=106Wm-2K-’ case’ and for the bed shortened to 50% (L = 2.2 m) and to 25%
10 t, = l/2&, 0.92 575 (L = 1.1 m), at the end of a ‘half cycle’.
= 25 min
11 t, = 2 x t,, 0.92 578
= 100 min Comparison between the predicted temperature profiles
12 inert = 2 x 0.55 m 0.47 737 along the bed and those measured in the laboratory
catal. = 2 x 0.55 m
reactor described in ref. 2
Homogeneous model eqns. (3) and (8) Figure 4 presents a comparison between the tempera-
13 Homogeneous mode1 with 0.84 658 ture profiles along the catalytic bed predicted by the
,&=,I,=1 Wm-‘K-r
14 Homogeneous model with 0.82 687
model [eqns. (l)-(7)] and the experimental profiles
I.= = 0 taken from the monograph of Matros [2]. The profiles
Data after Bunimovich et al. [3a]
shown correspond to three instants during a 50 min
15 Experimental data from 0.93 SO-550 quasi-stationary cycle, i.e. at the beginning of the cycle,
pilot plant at 8 min and at 15 min (cf. the description of the
16 Predicted by the 0.91 580 experiments in ref. 2). The points in Fig. 4 correspond
mathematical model to the measured values redrawn from the graph cited in
ref. 2, the broken line representing the profile calculated
without any losses of heat from the reactor while the
segment. The decreasing temperature allows one to solid line denotes the profiles estimated assuming heat
make use of the advantageous equilibrium conditions losses at an ambient temperature of 20 “C and with
for lower temperatures at the end of the bed, thus
assuring a very high final conversion.

x r-1
0.65 1.3 L un,

Fig. 4. Calculated (solid lines -with heat losses; broken


2.2 4.4 lines - without heat losses) and measured (points after ref. 2)
Fig. 2. Calculated temperature profiles along the bed for the temperature profiles along the bed of a laboratory reactor for
‘basic case’ (L = 4.4 m) and for the bed shortened to 50% 3.2% SO* and t, = 50 min, at the following instants: t = 0 (begin-
(L = 2.2 m) and to 25% (L = 1.1 m). at the end of a ‘half cycle’. ning of a cycle); t = 8 min; and t = 15 min.
241

TABLE 3. Comparison of the measured conversions with those predicted by the mathematical model

Average Unit Simulation variable


conversion
1.3% so, 3.2% SO, 9.4% so2

Pilot act. [ 151 Laboratory act. [2] Pilot act. [3a] Pilot act. [ 151

Calculated _ 0.968 0.923 0.917 0.918


Measured - 0.95 0.94 0.91 0.896

the reactor-surroundings heat-transfer coefficient, hS,


equal to 3.5 W mm2 K-‘.
In his monograph, Matros [2] quotes heat losses in a
laboratory installation of up to 50%. In the simulation
described in the present paper, the value of the heat-
transfer coefficient, hS, was chosen in such a way that
the value of the maximum temperature, T,,, would be
close to that obtained in a laboratory installation. For
such a value of hS, based on a deviation of the average
increase in temperature in the reactor from AT,,, as
calculated from eqn. (16), it is possible to estimate the
heat losses relative to the overall amount of heat gener-
ated in the system. Such an estimation yields a value of Ol,,.,,.,,,,,,~,,,~,,,,,,,,!,.,. ,,,.0.5 ,,I,, ,,,,0.b ,,,,, -I
0.0 0.1 0.2 0.3 0.4 0.5 oh oh .o x C-J
up to 70% thus exceeding that given in ref. 2 (up to 1.85 3.7 L cm)

50%). This seems to be too high a value, although the Fig. 5. Calculated (lines) and measured (points after ref. 15)
experience with laboratory reactors shows that at small temperature profiles along the bed of a pilot reactor at the instant
reactor diameters, high temperatures and in the pres- of flow reversal (for 1.3% SO2 and t, = 30 min).
ence of a large number of metal elements in the bed
(e.g. thermocouples), these losses may indeed assume I” cl
high values. In addition, these elements strongly distort 800
the ‘theoretical’ temperature profile along the bed. 700
A comparison of the measured conversions with those
simulated in the present work is shown in Table 3. 600

s 500
s
Comparison between the predicted temperature pro$les b
f400
along the bed and those measured in a pilot reactor
I?!
described in refs. 2, 3 and 15 300

As in ref. 2, refs. 3 and 15 also contain sufficient 200


information to enable a comparison to be made be-
tween the simulation results and measurements pcr-
formed in a pilot plant operating at the gas flow rates ‘“r
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 x C-J
of 2500-3000 SCM h-l. In ref. 15, the experimental 1.0 2.0 L cm1

profiles have been given for a reactor operating at SO1 Fig. 6. Calculated (lines) and measured (points after ref. 15)
concentrations of 1.3 and 9.4% (in the latter case, with temperature profiles along the bed of a pilot reactor, with cooling
cooling in the middle of the bed). The data used in the in the middle of the bed, at the instant of flow reversal (for 9.4%
simulations are again presented in Table 1. A compari- SO, and t, = 60 min).
son of the simulation results with the measured values
(taken from the graphs presented in ref. 15) is shown in gas whose temperature is equal to the inlet temperature
Figs. 5 and 6. For a concentration of 9.4%, an addi- to the reactor, i.e. 80 “C. The size of the exchanger was
tional problem arises in estimating the amount of heat selected in such a way that the cooling obtained might
withdrawn from the middle of the bed. It has been be similar to that observed in the pilot plant (for a
assumed, therefore, that in accord with the flow sheet of heat-transfer coefficient of c. 15 W mm2 K-l, this condi-
the installation discussed in ref. 15 cooling is carried tion would lead to an exchanger with an area of some
out in an indirect heat exchanger using a cold process 20 m’). The temperature drop in the middle of the bed
varies during the cycle; using the above assumptions, size (cf. the values for the SO* concentration of 3.2%).
however, its average value was equal to c. 90 “C. From an analysis of the parametric sensitivity, it follows
that a twofold change in c1produces a substantial change
in T,,, and, consequently, in the average conversion
Summary of the results against a background of the X,,. It may be concluded, therefore that the size of the
literature information catalyst pellet influences the parameters of the process to
a greater extent through the change in the product
The simulation calculations performed have been (U x S) than in any change in A,.
based pn values of the coefficients characterizing the During simulation, a strong effect has been found
gas-to-catalyst heat transfer, CI, and the effective heat between the numerical structure of the model for the
conductivity of the bed, I,, estimated according to known reverse-flow reactor (positive information feedback) and
literature correlations [ 16- 191calculated as a mean value the accuracy of the calculations. Thus, it is essential that
of two different formulae. Strictly speaking, the intensity a check of the accuracy should be carried out by testing
of gas-to-catalyst heat transfer is determined by the the consistency of the overall heat balance of the reactor.
product of the coefficient CIand the specific surface area The agreement reported in this paper between the
of the catalyst, S, i.e. (CXx S). Both a and & may be temperature profiles predicted and those measured is not
evaluated using various literature formulae which, how- perfect. However, taking into account that only pub-
ever, do not always strictly correspond to the conditions lished experimental data have been used, it is reasonable
prevailing in a given reactor. Their calculation may to assume that the model employed is sufficiently reliable
therefore a priori contain considerable inaccuracies. for the preliminary design of the reverse-flow reactors.
To date, attention has been paid mainly to the effect It should be pointed out that a one-dimensional model
of heat conductivity in the bed. In a number of publica- cannot be expected to yield ideal agreement with the
tions (cf. for example, ref. 2), Matros has given approx- experimental results which themselves are usually bur-
imate relationships between the maximum temperature, dened with considerable measurement errors. The other
Tnlax2 and the parameters which describe the kinetics of source of discrepancy may have been the exclusion from
the reaction, AT,, and the effective heat conductivity in the model of heat transfer through the elements of the
the bed, il,. These relationships do not include the reactor body.
parameters which characterize the intensity of the gas- A number of publications have appeared which treat
catalyst heat transfer, i.e. (CI x S). The simulations per- the question just discussed as a purely numerical prob-
formed revealed, however, that within the range of lem (cf. refs. 9-12). In this event, any physical conclu-
values employed for these parameters their influence in sions reached would seem to be doubtful. The
a reverse-flow reactor is quite considerable. information given in ref. 12, that the ‘quasi-stationary’
Simulations of various types of plant start-ups de- state is reached after only 150-300 cycles suggests that,
scribed in Part I of this series did not demonstrate such when account is taken of the strong dependence of the
an important influence for (a x S) as in the case dis- results on the initial conditions, the model has simply
cussed, hence this effect must be connected with the produced a drift resulting from numerical errors.
special nature of flow reversals in the reactor. A reverse-flow reactor for the oxidation of SOa has
In ref. 2 as well as in other papers, it has been found also been simulated by Rakowski [ 141. The model was
that a decrease in the maximum temperature may be investigated at an inlet temperature, TO, of c. 500 “C;
achieved by decreasing the linear velocity of the gas and the value of T,,, obtained in the simulations exceeded
increasing & through increasing the pellet size. This 1000 “C. This suggests that a physical objective was
increase in size does indeed lead to a radical decrease in simulated under conditions of little practical impor-
Tmax. This, however, results more from the associated tance. The lack of any physical verification of the
decrease in both x and S than from an increase in 1,. For model, or at least an application of the ‘heat balance
the pellet sizes employed and values of the Re and Pr consistency test’ proposed earlier, indicates little practi-
numbers which actually occur in SO* oxidation reactors, cal utility for the computer analysis performed.
the same value of A, could well fall within the range of Example: Most authors have found an essential inter-
0.5-1.5 W m-’ K-l. An increase in 1, up to the unreal- relationship between the parameters describing the re-
istically high value of 5 W m-’ K-’ led to an effect in the action rate and the average conversion. On the other
simulations performed which, although noticeable, was hand, Gawdzik and Rakowski [9] have suggested that
too small to justify any definite conclusion about a the DamkGhler number (Da) does not affect the,
decisive influence of 1, on T_,. It may easily be seen efficiency of the reactor, whereas Gupta and Bhatia [ 121
from the data presented in Table 1, that the product have found a strong effect for Da using the same model.
(m x S) for realistically assumed values may change by Such widely differing conclusions are not unusual when
a factor of more than twofold with a change in the pellet the numerical analysis is not related to the actual
243

characteristics of a given process. Over the range of taken in the application of inert sections of bed;
temperatures for which the process was analyzed in ref. they should be employed only in the case of reac-
14, the reaction rate was indeed rather small and, tions with a small heat effect (small ATad). In the
hence, the effect of Da may have been somewhat weak. oxidation of SO,, this concerns cases of lower
Such conclusions do not have, however, any practical reactant concentration.
significance. This may be illustrated by the fact that, 4. The heterogeneous model presented in this paper
according to ref. 14, the average conversion obtained [eqns. ( 1) -(7)] yields results which are sufficiently
for the process was 0.3, although from the numerous reliable for the preliminary design of reverse-flow
papers of Matros it follows that in this case we are reactors for SO2 oxidation.
dealing with reactors operating at conversions exceed-
ing 0.9. On the other hand, refs. 11 and 12 suggest that
an increase in the gas-catalyst heat transfer increases Acknowledgement
T,,, (thus agreeing with the conclusions of the present
paper), but that this also leads to a substantial increase The present analysis of the model of a reverse-flow
in conversion. Such a situation is possible when the reactor would have been impossible but for the papers
reaction in the bulk of the reactor occurs under condi- published by Professor Matros’ group from the Insti-
tions far removed from equilibrium - a condition tute of Catalysis in Novosibirsk [2, 3a, 151. The author
rarely met in this type of reactor. Such conclusions wishes, therefore, to acknowledge this fact and thank
should, therefore, never be regarded as ‘general’ the authors of these publications. He also hopes that
they will not object to the use made of their results in
Practical design conclusions the present article.
On the basis of the literature quoted and the tests
performed for the heterogeneous mathematical model,
some important practical conclusions may be formu- Nomenclature
lated:
c reactant ( SOZ) concentration, kmol
1. When the catalytic bed is sufhciently long, i.e. when kmol-’
a flat ‘heat wave’ crest is clearly formed, the final CiC gas specific heat, kJ kmol-’ K-’
conversion and the maximum temperature in the G catalyst specific heat, kJ kg-’ K-’
bed depend only weakly on a number of the model D reactor diameter, m
parameters. If, therefore, a judicious excess of cata- Da dimensionless Damkohler number
lyst is employed, no deterioration in the final conver- h, catalyst-surroundings heat-transfer co-
sion nor superheating may be feared; on the contrary, efficient, W me2 K-i
this leads to a stable operation irrespective of the H heat of reaction, kJ kmol-’
presence of possible disturbances in the process. K, reaction rate constant, h-’ Pa-’
2. Reaction occurs solely during the increasing and L reactor length, m
decreasing parts of the heat wave (Fig. 3) with 53 molar gas flux, kmol h-’ mm2
equilibrium prevailing along its crest. This produces PAR(i) matrix describing the parameters of the
a clear dependence of conversion on T,,,. Apart ith reactor pass
from the effect of linear velocity (cf. Matros [2]), r effective reaction rate, kmol kg-’ h-l
the basic factors affecting T,,, are the pellet size s specific catalyst surface area, m2 me3
and the catalyst activity. Thus, for higher SO, STREAM 1 matrix describing streams in the system
concentrations, a catalyst should be chosen which STATE 1 matrix describing the state of the reactor
has a greater granular size and good activity. These passes
two conditions may, to some extent, be contrary to t time, h
each other, although they do provide guidelines for T temperature, K or “C
the preparation of a catalyst for reverse-flow reac- Tmax maximum catalyst temperature, K or “C
tors. For very low concentrations, the catalyst used X axial coordinate, m
should be of a rather small pellet size. 2 =x/L dimensionless axial coordinate
3. The use of an inert bed results in the heat wave x conversion
slope occurring in a part where no reaction occurs. -&” average conversion
This leads to a substantial increase in T,,,,, and to
the contraction of that part of the bed in which Greek letters
reaction occurs and, in turn, to a decrease in the a gas-solid heat transfer coefficient,
average conversion. Hence, great care should be kJ m-* h-‘K-’
244

6 simulation error 5 Yu. S. Matros, A. S. Noskov, V. A. Chumachenko and 0. V.


adiabatic temperature rise, K Goldman, Theory and application of unsteady catalytic detox-
ATad
ication of effluent gases from sulfur dioxide, nitrogen oxides
P” catalyst bulk density, kg me3
and organic compounds, Chem. Eng. Sci., 43 (1988) 2061-
& effective heat conductivity,
2066.
kJ m-i h-’ K-i 6 J. Wojciechowski and J. Haber, .Z. Appl. Catal., 4 (1982)
E void fraction coefficient 275.
7 J. Wojciechowski, Swingtherm - a new method of catalytic
gas purification, Przem. Chem., 62 (1982) 502-504 (in Polish).
Subscripts 8 G. Eigenberger and V. Nieken, Catalytic combustion with
b basic variant periodic flow reversal, Chem. Eng. Sci. 43 (1988) 2109-
g gas 2115.
in reactor inlet 9 A. Gawdzik and L. Rakowski, Dynamic properties of the
k catalyst (solid phase) adiabatic tubular reactor with switch flow, Chem. Eng. Sci., 43
(1988) 3023-3030.
0 pass inlet (x = 0)
10 A. Gawdzik and L. Rakowski, The methods of analysis of
out reactor outlet the dynamic properties of the adiabatic tubular reactor
s surroundings with switch flow, Comput. Chem. Eng., 13 (1989) 1165-
1173.
11 S. K. Bhatia, Analysis of catalytic reactor operation
with periodic flow reversal, Chem. Eng. Sci., 46 (1991) 361-
References 367.
12 V. K. Gupta and S. K. Bhatia, Solution of cyclic profiles in
(a) K. Gosiewski, Dynamic modelling of industrial SO2 oxida- catalytic reactor operation with periodic flow reversal, Com-
tion reactors. Part I. Model of ‘hot’ and ‘cold’ start-ups of the put. Chem. Eng., 15 (1991) 2299237.
plant, Chem. Eng. Process., 32 (1993) 1 1 l- 129; (b) A. Renken, 13 J. Thullie and A. Burghardt, Methanol synthesis in a flow
Unsteady state operations of continuous reactors, Znt. Chem. reversal reactor, Znz. Chem. Proc., 10 (1989) 175-188 (in
Eng., 24 (1984) 202-213. Polish).
Yu. S. Matros, Catalytic Processes under Steady State Condi- 14 L. Rakowski, Investigation of the dynamic properties of the
tions, Elsevier, Amsterdam/Oxford/New York/Tokyo, 1989. adiabatic tubular reactor with switch flow, PhD thesis, Silesian
(a) G. A. Bunimovich, V. 0. Strots and 0. V. Goldman, Technical University, Ghwice, 1986 (in Polish).
Theory and industrial applications of SO, oxidation reverse 15 A. A. Balashov, Yu. W. Filatov, W. P. Kozlov, W. C.
process for sulfuric acid production, Proc. Znt. Conf. Unsteady Epifanov, A. E. Popov, 0. N. Smimova and T. Yu.
State Processes in Catalysis, June 1990, Novosibirsk, USSR, Glukhova, A nonstationary method of sulphur dioxide oxida-
VSP, Utrecht/Tokyo, pp. 7-24; (b) G. F. Froment, Reversed tion in sulphuric acid production, Khim. Prom. (Moscow),
flow operation of fixed bed catalytic reactors, ibid., pp. 58-87; (1987) 730-733 (in Russian).
(c) Yu. S. Matros, Performance of catalytic processes under 16 B. W. Gamson, Chem. Eng. Prog., 47(1951) 19.
unsteady state conditions, ibid., pp. 131-163; (d) K. Gosiewski 17 C. R. Wilke, Trans. Am. Inst. Chem. Eng., 41 (1945) 445.
and R. Sztaba, A simplified design of reverse flow nonstation- 18 D. G. Bunnel, H. B. Irvin, R. W. Olson and J. M. Smith, Znd.
ary reactor for low reactant concentrations, ibid., pp. 629-635; Eng. Chem., 41 (1949) 1977.
(e) J. Thullie and A. Burghardt, Application of the flow 19 C. A. Coberly and W. R. Marshall, Chem. Eng. Prog., 47
reversal reactor to the methanol synthesis, ibid., pp. 687-692. (1951) 141.
G. K. Boreskov and Yu. S. Matros, Unsteady state perfor- 20 R. Sztaba and K. Gosiewski, A method of matching kinetic
mance of heterogeneous catalytic reactions, Catal. Rev. Sci. equations of the SO, oxidation on vanadium catalyst, Znz.
Eng., 25 (1983) 551-590. Apar. Gem., (1981) 3-7.

You might also like