You are on page 1of 9

Ind. Eng. Chem. Prod. Res. Dev.

1982, 21, 29-37 29

CATALYST SECTION

Transients of Monolithic Catalytic Converters: Response to Step


Changes in Feedstream Temperature as Related to Controlling
Automobile Emissions

Se H, Oh* and James C. Cavendish

General Motors Research Laboratories, Warren, Michigan 48090


See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.
Downloaded via HOGSKOLAN BORAS on February 23, 2022 at 13:25:59 (UTC).

The transient behavior of automobile monolithic converters has been studied using a mathematical model which
accounts for the simultaneous processes of heat transfer, mass transfer, and chemical reaction. The monolith
response to a step increase in the feedstream temperature was examined as a function of catalyst design
parameters and operating conditions In order to analyze their effects on the lightoff behavior of monolithic catalysts.
Also, our simulation results showed that a step decrease In the feedstream temperature can lead to a somewhat
unexpected transient temperature rise in the solid phase (that is, the catalytic wall) above its initial temperature,
provided that the reactant concentrations in the feedstream are sufficiently high. This phenomenon is similar to
the “wrong-way" behavior In packed-bed reactors previously reported in the literature.

Introduction response of a catalytic monolith after a hot monolith has


Catalytic monoliths used for controlling automobile been subjected to a cool feedstream containing high con-
emissions operate under highly transient conditions. One centrations of combustible species. Of particular interest
important mode of transients is a rapidly varying inlet is the prediction that a step decrease in the feedstream
exhaust temperature. At the time of a cold start, for ex- temperature can lead to a temporary temperature increase
ample, a cold monolith is suddenly exposed to exhaust gas in the solid phase (i.e., catalytic wall) of the monolith well
at an elevated temperature. above its initial temperature, even though the feedstream
The temperature transients of catalytic monoliths have temperature assumed after the step change is low enough
direct implications in automobile emission control. For to quench the reaction completely at steady state.
instance, converter performance during its warm-up period The prediction of such temperature excursions is in-
is an important factor in the design of catalytic converters, teresting in that the system’s variable (wall temperature)
because the Federal Test Procedure (FTP) driving cycle moves temporarily in the opposite direction to the per-
requires a cold start of a vehicle at the beginning of the turbation in the input variable (feedstream temperature).
test. Furthermore, automobile monolithic catalysts often The occurrence of this “wrong-way” phenomenon in
encounter unexplained temperature excursions which can packed-bed reactors has been observed experimentally by
result in thermal degradation or even melting of the many investigators (e.g., Hoiberg et al., 1971; Hansen and
monolithic converters (e.g., Morgan et al., 1973). There- Jorgensen, 1974; Van Doesburg and De Jong, 1976a, 1976b;
fore, it is of practical importance to understand how heat Sharma and Hughes, 1979) and has recently been analyzed
is generated and distributed in catalytic monoliths during in detail by Mehta et al. (1981). Our simulation results
exothermic chemical reactions under transient feedstream suggest that similar phenomena can also occur in mono-
conditions. It is hoped that such an understanding will lithic reactors when the reactant concentrations in the feed
aid in the development of monolithic converters which give are sufficiently high and the feedstream temperature is
improved warm-up performance and are less prone to rapidly decreased. It has been observed (Mondt, 1976,
overtemperature problems. 1981) that similar operating conditions (high unburned
In this paper a transient mathematical model of catalytic hydrocarbon concentration and decreasing exhaust tem-
monoliths is developed which accounts for the simulta- perature) are encountered during sudden vehicle decel-
neous processes of heat transfer, mass transfer, and eration or prolonged high-speed, closed-throttle coasts.
chemical reaction. The model is then used to simulate the Our model predicts, for a step decrease in the exhaust
dynamic behavior of a catalytic monolith following a step temperature, a substantial transient temperature rise in
change in the feedstream temperature. The Pt-catalyzed the solid phase of the monolithic converter under certain
oxidation reactions of CO, hydrocarbons, and H2 are con- conditions; however, the predicted temperature rise was
sidered. The monolith response to a step increase in the found to be too small to cause monolith melting, in ac-
feedstream temperature was examined as a function of cordance with a recent experimental observation (Mondt,
catalyst design parameters and operating conditions in 1981).
order to analyze their effects on the lightoff behavior of In succeeding sections a transient, one-dimensional
monolithic catalysts. We also simulated the transient monolith model is developed and used for parametric
0196-4321 /82/1221-0029$01.25/0 © 1982 American Chemical Society
30 Ind. Eng. Chem. Prod. Res. Dev., Vol. 21, No. 1, 1982

calculations. Emphasis will be given to the effects of Basic Equations and Assumptions. In this paper we
converter design parameters and operating conditions on adopt a transient, one-dimensional model to study the
both the lightoff behavior and temperature excursions of dynamic response of a catalytic monolith following a step
monolith converters. change in the feed temperature. The model neglects the
radial variations of the gas-phase temperature, concen-
Development of the Mathematical Model
In our model we consider the following oxidation reac- tration, and velocity within the individual channels so that
these variables are to be interpreted as cross-sectional
tions over Pt
averages. The material and energy balances for the gas
CO + y,o2 —
C02 phase are

C3H6 + 9/202 —
3C02 + 3H20 dcg,i dcg-i
= -

c8i¡) í =
1, ..., 5 (9)
CH4 + 202 —
C02 4- 2H20
dTg dTg
H2 + V2O2 -
H20 = + hS{T* ~

(10)
(PtCn~óf ~vPtCw~^ Ts>
Here propylene is assumed to be representative of “fast-
oxidizing hydrocarbons” in automobile exhaust and The corresponding model equations for the solid phase are
methane of “slow-oxidizing hydrocarbons” (Kuo et al., Ptot
1971). Also, hydrogen is included as a separate species a(x)R¡(ca,Ta) = -

i =
1.....5 (11)
because it significantly influences the behavior of a mon- jppknfiicgj ca4)

olith during the warm-up process, as will be shown later.


The specific reaction rate expressions (i.e., rates per unit
Pt surface area) for the oxidation of CO, C3He, and CH4
were obtained by calibrating the rate equations of Voltz d2Ts 4

et al. (1973) against the recycle reactor data obtained in 8 (1


-

«)—7 + hS{Tg -

Ta) + a(x)Z(-miRi(ca,Ta)
dx¿ i=1
our laboratory (Schlatter and Chou, 1978). In this cali- (12)
bration, only the preexponential factors of the rate con-
stants were adjusted, without changing the activation en- The axial dependence of a(x) is explicitly shown in eq 11
ergies and adsorption equilibrium constants given by Voltz and 12 to stress the fact that the noble metal surface area
et al. The rate expressions used in the computations are is permitted to vary along the reactor length. The sub-
given below. script i in eq 9 and 11 refers to the species of interest: i
=
1, CO; i = 2, C3H6; i = 3, CH4; i = 4, H2; and i = 5, 02.
Rco kiccoco2/G
-
mol CO/(cm2 Pt-s) (1) Other limiting assumptions invoked in formulating the
above equations include (1) negligible temperature gra-
Rca H6
=
^2cC3Hec02/G mol C3H6/(cm2 Pt-s) (2)
dients in the solid phase in the transverse direction, (2)
Rcnt
=
^3CCH4C02/G mol CH4/(cm2 Pt-s) (3) negligible axial diffusion of mass and heat in the gas phase,
and (3) the occurrence of chemical reactions only on the
where external surface of the catalytic wall.
G = We remark that axial heat conduction in the solid phase
is included in our model because, as we shall see, it has
T{ 1 + KlCco + K2Cc3h6)2(1 + K3Cco2Cc3He2)(l + tf4cNO°·7) an important effect on the wall temperature profile.
(4) For our calculations, the accumulation of mass and heat
¿Ü!
= 6.699 X 109 exp(-12556/T) in the gas phase was neglected (that is, dcg¿/dt = dTJdt
=
0 in eq 9 and 10), since their time constants are typically
k2 = 1.392 X 10n exp(-14556/71 much smaller than that of the solid thermal response
= 7.326 X 106 exp(-19 000/T)
(Young and Finlayson, 1976a). With this quasi-static ap-
k3 (5)
proximation, then, the appropriate boundary and initial
and conditions are
=
65.5 exp(961/7! cg,.(o,t)
=
cgir (13)

K2 =
2,08 X 103 exp(361/D rg(0,t)
=
Tg“ (14)

K3 = 3.98 exp(ll 611/71 dTa dTa

exp(-3733/7!
K4 = 4.79 X 105
(6)
^ , ) -r-W)
ox Ox
= = 0 (15)

The rate of hydrogen oxidation in the exhaust gas as-


is Ta(x, 0) =
Ta0(x) (16)
sumed to be the same as that for carbon monoxide (Kuo The boundary conditions given by indicate that no eq 15
et al., 1971; Young and Finlayson, 1976b). That is to say heat exchange is assumed between the solid and the sur-
~

^ich2co2/G mol H2/(cm2 Pt-s) (7) roundings both at the inlet and the outlet of the monolith.
#h2 The values of gas/solid heat and mass transfer coeffi-
This assumption is compatible with earlier experimental cients were estimated based on the Nusselt and Sherwood
observations (Dabill et al., 1978; Stetter and Blurton, 1980) numbers obtained from analytical solutions for fully de-
that in the presence of CO, H2 oxidation rate is inhibited veloped laminar flow with constant wall heat flux (Shah
by CO to approximately the same extent as the CO oxi- and London, 1971). These Nusselt and Sherwood numbers
dation rate. From stoichiometry, the reaction rate for are the asymptotic values reached downstream of the de-
oxygen is given by veloping laminar flow regime located near the monolith
0.5fiCo + 4.5Rc3h6
inlet; however, the length of this hydrodynamic entrance
4-
2Rch< +
~

Ro,
region is only a small fraction of the monolith length under
0.5Rh2 mol 02/(cm2 Pt-s) (8) typical converter operating conditions (Sherony and Sol-
Ind. Eng. Chem. Prod. Res. Dev., Vol. 21, No. 1, 1982 31

brig, 1970), and thus its contribution was neglected in the Table I. Standard Set of Parameter Values
calculations. The heat and mass transfer coefficients were =
101.3 kPa (1 atm) =
2.5 g/cm3
Ptot p3
calculated from Wg
=
40 gls a 268.95 cm2 Pt/cm3 reactor
=

A =
60 cm2 Xs= 1.675 X 10"2 J/cm-s-K
h =
Nu„Xg/(2fih) (17) L =
10 cm Cpg 1.089 J/g-K
=

Rh = 0.06062 cm Cps= 1.071 + 1.56 X 10"4 T3-


km¿
=
ShJ)t/( 2Bh) (18)
(square channel) 3.4 3 5 X 104/TS2 J/g-K
e =
0.6836
where the molecular diffusivity of species i, D¡, was esti-
mated using the Slattery-Bird formula (Bird et al., 1960), case 1. step increase in step decrease in
case 2.
and the thermal conductivity of the reaction mixture, Xg, feed temperature feed temperature
was approximated by that of N2 (=2.269 X lO^Tg0·832 J/ Tso 300 K
=
T30 750 K =

cm-s-K). T„in 600 K =


T„in 300 K =

Numerical Solution of the Equations. The purpose 2% CO / 2% CO


of this section is to provide a brief description of the nu- 450 ppm CbH6 | 6000 ppm C3H6
.in 50 ppm CH4 in 1 650 ppm CH4
merical methods used to approximate the variables TB, Tg, ,» - . -

0.667% H, g-‘ ( 0.667% H2


(cg-1, cg>2,..., Cg>5)T and c„ whose transient behavior is
=
cg 5% 02 J
3.5% O,
governed by eq 12,10,9, and 11, respectively. To that end 500 ppm NO I 100 ppm NO
we express these equations in the following compact form

ST S2T Note that eq 24 is an initial-value problem which re-


( ß)1
=
M1" e)^i + (19) quires evaluation of Tg and Ca. Approximations to these
values are provided by replacing eq 20, 21, and 22 by
dTg spatially discrete analogues.
-
F2(T8,Tg) (20) =
Tg(x;,f))/Ax
-

(Tg(x,·+!,£)
T2((Ta‘ + Tai+1)/2, (Tg(xi+1,t) + Tg(x,.,i))/2) (25)
=
(21)
F3(Tg,Cg,cB) T3((Tg(x¡+1,í) + Tg(x¡,í))/2,
=
-¡^ Cg(x¡,í))/
-

(Cg(x¡+1,t)
(cg(x,+1,í) + Cg(Xi,í))/2, (ca(xi+1,í) + ca(x¡,í))/2) (26)
0 =
F4(Tg,TB,cg,cB) (22)
T4((Tg(x,·+!,£) + Tg(x,-,í))/2, (T/ + Ta‘+1)/2,
where ( ) and F¿ (1 < i < 4) are defined in the No-
(cg(xí+1,t) + Cg(x,-,í))/2, (ca(xí+1,í) + ca(x¡,í))/2) 0 (27)
=
menclature section.
Notice that eq 19-21 represent a set of coupled nonlinear Equation 24 is an ordinary matrix differential equation
differential equations while eq 22 is an algebraic equation while eq 25-27 represent a nonlinear set of algebraic
relating Ta, Tg, cg, and c„. Also, eq 19 is a transient partial equations in which t is a parameter. If we regard Tg, cg,
differential equation while eq 20 and 21 are ordinary and c8 in eq 25-27 as implicit functions of Ta, then eq 24
differential equations in the independent variable x with is of the form
the time variable, t, regarded as a parameter. If the solid
temperature Ta(x,t) is known at a given value of time, then =
F(Ts,t) t > 0 (28)
eq 20 (with appropriate inlet boundary conditions) can be B(Ta)^
solved for Tg and the coupled system of eq 21/22 can then In order to evaluate the right-hand side of eq 28 for a given
be solved for concentrations cg and ca. These simple ob- vector Ta and a given value of t, the nonlinear system of
servations provide the basis for the numerical methods eq 25 must be solved for Tg(x„í), 1 < i < N followed by
used to approximate eq 19-22. the solution of the coupled nonlinear system of eq 26/27
Our approach to solving eq 19-22 was to first partition for Cg(x¡,t) and ca(x„t). Hence, eq 28 represents a nonlinear
the spatial domain (0 ) with a uniform spatial grid initial-value problem in which evaluation of the forcing
where : 0 = xl< x2< < xN = L. We remark that for
...
term requires the solution of a nonlinear set of equations.
the calculations reported here, N = 81. Next, we ap- To integrate eq 28 we used a banded version of the
proximate Ta in eq 19 by piecewise linear trial functions GEAR codes (Hindmarsh, 1976), a carefully engineered
4>j{x) associated with (see Cavendish and Oh, 1979, for Fortran subroutine for the integration of stiff ordinary
details) matrix differential equations. High, controllable numerical
accuracy (in the time variable, t) is achieved at minimum
Ta(x,t) =
77'( )0;·( ) (23) computer cost by the dynamic variation of both time steps
;'»i and multistep integration methods used in the code. To
Galerkin’s method is then used to replace eq 19 by a solve the nonlinear systems of eq 25 and 26/27 for
semi-discrete space, continuous-time matrix differential Tg(x¡+i,í), and cg(xi+1,£), cB(xi+1,t), respectively, Newton’s
equation of the form method was used.
It should be noted that strictly speaking, the Jacobian
of F in eq 28 is not a banded matrix. However, because
=
ETa + Fx{fa,Tt,Ca) (24)
B(fe)^ of the high, local spatial dependence of Tg(x,t), ca(x,t), and
where Cg(x,t) on Ta(x,t), the Jacobian of F is well approximated
by a banded tridiagonal matrix.
fa =
(TaHt), TaHt).....7VW Results and Discussion
=
Table I shows a standard set of parameter values used
Tg (Tg(xut), Tg(x2,t).....Tg(xjv,t))T in the computations for both a step increase (case 1) and
Ca
=
(, ), ca(x2,i), .... ca(xN>t))T
a step decrease (case 2) in the feedstream temperature. For
each of these two cases, to be reported here separately, we
The initial condition [TB(0)]; =
[T^x,)] is imposed on Ta(£) will first examine the standard case in detail to gain insight
to complete the system. into the system’s behavior. Then, the parametric sensi-
32 Ind. Eng. Chem. Prod. Res. Dev., Vol. 21, No. 1, 1982

Figure 1. Time variation of the wall temperature profile following


a step increase in the feedstream temperature, at the standard con- Figure 2. Normalized axial position of wall temperature peak as a
ditions listed in Table I (case 1). function of time for three different values of solid thermal conduc-
tivity. Other parameter values are listed in Table I (case 1).
tivity of the system will be investigated by systematically
perturbing the key design and operating parameters
around the standard values.
The monolith properties listed in Table I correspond to
a square-channel ceramic monolith (cordierite) having a
cell density of 46.5 openings/cm2 (300 openings/in.2) and
a wall thickness of 0.0254 cm. The exhaust gas compo-
sition and flow rate for case 1 are similar to the test con-
ditions of an engine dynamometer system developed for
the evaluation of converter lightoff (Herod et al., 1973).
The exhaust compositions used for case 2 were obtained
by measuring the engine-out concentrations of a 1979
California Buick 3.8-L V6 engine when the vehicle sud-
Figure 3. Conversions of the individual species as a function of time
denly decelerated after stable engine operation at 96 km/h at the standard conditions listed in Table I (case 1).
(Mondt, 1981). The initial converter temperature for case
2 is assumed to be uniform, because the exothermic tem- location of the temperature peak at early times (for t <
perature rise in the converter after the stable engine op- 60 s) is independent of the thermal conductivity of the
eration was observed to be minimal as a result of the at- solid phase, it significantly influences the wall temperature
tendant low CO and hydrocarbon concentrations in the profile in the later stages of the warm-up process. This
exhaust (Mondt, 1981). As will become apparent from the demonstrates the importance of including the axial heat
discussion of the results for case 2, the combination of the conduction in the solid phase in the modeling of monolith
high reactant (primarily hydrocarbon) concentrations and lightoff.
the high initial monolith temperature listed in Table I It is worth mentioning that a monolithic converter ex-
results in a temporary temperature rise in the monolith hibits a different mode of lightoff behavior depending on
during the transients. During vehicle deceleration, exhaust the temperature level of the exhaust gas. At sufficiently
flow rate also decreases with time; its effect on the tem- high exhaust temperature (e.g., 700 K), the lightoff occurs
perature excursions will be examined later. at the monolith entrance after the converter is heated up
Case 1. A Step Increase in the Feed Temperature. by convective heat transfer, so that the solid temperature
In this case we simulate the transient response of a cata- remains highest at the inlet of the monolith throughout
lytic monolith, initially at room temperature, following a the transient period. This is in contrast to the case of Tgin
step flow of stabilized exhaust gas at an elevated tem- = 600
K, where the location of the solid temperature peak
perature. Figure 1 shows the time variation of the wall varies with time along the reactor length, as was illustrated
temperature profile in the monolith for the standard in Figure 2.
conditions listed in Table I. At early times, the hot exhaust Figure 3 shows the conversions of the individual species
heats up the upstream portion of the monolith primarily as a function of time for the standard conditions listed in
by convective heat transfer (see t = 15.7 s). As time Table I (Tgm = 600 K). Although the conversions at a given
elapses, however, the downstream section of the monolith time are somewhat different among the species of interest
becomes hotter than the upstream section (see t = 47.2 s) here, their lightoff times (say, the time required for 50%
because the reaction exotherm generated in the upstream conversion) are similar except for CH4, because of the
section is constantly carried downstream by the exhaust kinetic coupling between the species. For clarity, then, we
flow. This leads to the development of a temperature peak will use a plot of CO conversion vs. time only in subsequent
in the downstream section as a result of a vigorous reaction figures in order to illustrate the warm-up performance of
in that region, and then this temperature peak moves the monolithic catalytic converter.
rather slowly toward the inlet of the monolith (see t = 71.8 Figure 4 shows the effects of the CO concentration in
s, 124.1 s, and 300 s). the exhaust on the lightoff behavior of a monolith. The
Such movement of the temperature peak in the solid catalyst lightoff is delayed substantially upon increasing
phase during the converter warmup process is clearly il- the CO concentration, as can be anticipated from the
lustrated by Curve B in Figure 2. Also shown in Figure negative-order kinetics of the CO oxidation reaction over
2 are the results obtained when the thermal conductivity Pt. The variation of the C3H6 concentration similarly
of the solid is perturbed about its standard value ( , = affects the catalyst lightoff, because the oxidation rates
1.675 X 10"2 J/cm-s-K) by a factor of 10 (see curves A and are also inhibited by C3He as a result of its relatively strong
C). The thermal conductivity value assumed in curve C chemisorption on Pt. Increasing the 02 concentration in
is representative of a typical metal-substrate monolith. Of the exhaust, on the other hand, improves both the lightoff
particular interest is the observation that athough the and steady state oxidation performance.
Ind. Eng. Chem. Prod. Res. Dev., Vol. 21, No. 1, 1982 33

Table II. Channel Size and Void Fraction as a Function


of Cell Density
cell density,
openings/cm2 Rh, cm e

3LÓ 0.07710 0.7371


46.5 (standard) 0.06062 0.6836
93.0 0.03915 0.5696

Figure 4. Effects of CO concentration in the exhaust on catalyst


lightoff. Other parameter values are listed in Table I (case 1).

Figure 7. Effects of cell density on catalyst lightoff at T,™ = 600


K. Other parameter values are listed in Table I (case 1).

Figure 5. Catalyst lightoff behavior with and without H2 in the


exhaust. Other parameter values are listed in Table I (case 1).

Time (s)

Figure of cell density on catalyst lightoff at T„™ = 700


8. Effects
K. Other parameter values are listed in Table I (case 1). (Note the
change in time scale.)

shortens its lightoff time (Oh et al., 1980). This difference


in lightoff behavior can be explained on the basis of a
difference in the heat transfer characteristics between the
Figure 6. Effects of exhaust gas flow rate on catalyst lightoff. Other
two converter configurations; in contrast to monoliths
parameter values are listed in Table I (case 1). (where the heat transfer coefficient is virtually independent
of gas flow rate as can be seen from eq 17), the gas/solid
The concentration of H2 in the exhaust gas is usually heat transfer coefficient for packed-bed converters in-
assumed to be determined by the equilibrium of the creases with increasing gas flow rate (De Acetis and
water-gas shift reaction, which gives a mole ratio of ap- Thodos, 1960). As the exhaust flow rate through a
proximately 1:3 between H2 and CO (Wei, 1975). As shown packed-bed converter increases, this improvement in the
in Figure 5, the model predicts considerably faster catalyst heat transfer characteristics dominates the competitive
lightoff when the presence of H2 in the exhaust is ac- effect of the decreased residence time of gas within the
counted for. It is also interesting to note that H2 in the reactor, resulting in the enhancement of the converter
exhaust leads to a solid temperature exceeding the adia- lightoff performance.
batic reaction temperature even at steady-state conditions The cell density (that is, the number of openings per unit
(Heck et al., 1976; Young and Finlayson, 1976b), as illus- cross-sectional area) is an important design parameter in
trated in Figure 1 (compare the wall temperature at x = developing a monolithic catalyst with improved perform-
0 and x = 1 when t = 300 s). Such an overshoot in the solid ance. hi actual catalyst design, the wall thickness is usually
temperature profile was observed experimentally by He- held constant when perturbing the cell density. This re-
gedus (1975) during steady-state oxidation of H2. This is sults in the simultaneous variation of both the channel size
so because, as indicated by its high Lewis number (= and the void fraction of the monolith, as the cell density
DjPgCpg/ , 4), H2 can diffuse to the catalytic wall faster is varied (see Table II; square channels with a fixed wall
than the neat generated as a result of reaction can be thickness of 0.0254 cm). As shown in Figure 7 (Tg“ = 600
conducted away. K), the lightoff characteristics of the catalyst are improved
The effect of exhaust gas flow rate on the lightoff be- as the cell density is decreased. The improved lightoff
havior of a monolith is depicted in Figure 6. The monolith performance predicted for monoliths having lower cell
lightoff is delayed upon increasing the flow rate, primarily densities can be attributed to their lower solid fraction (1
due to the attendant decrease in the contact time between -

<), and thus the reduced thermal capacity of the mono-


the catalyst and exhaust gas. It is interesting to note that lith.
the opposite effect is predicted for a packed-bed converter It is important, however, to emphasize that decreasing
as the flow rate is varied; that is, increasing the flow rate the cell density does not always improve catalyst lightoff
34 Ind. Eng. Chem. Prod. Res. Dev., Vol. 21, No. 1, 1982

Figure 11. Time variation of wall temperature profile following a


Figure of monolith length on catalyst lightoff. Other
9. Effects step decrease in the feedstream temperature, at the standard con-
parameter values are listed in Table I (case 1). ditions listed in Table I (case 2).

and the linearly increasing (curve C) activity profiles are


considered, and the computational result for the standard
case (uniform activity profile, curve B) is also reproduced
in the same figure for a direct comparison. Note that the
Pt content in the monolith is kept constant in all three
cases that are considered here. It can be seen that the
catalyst’s lightoff performance as well as the overall per-
formance is improved substantially when the noble metal
is concentrated in the upstream section of the monolith
(compare curves A and B). In the case of the linearly
increasing Pt profile, on the other hand, the lightoff time
Figure 10. Effects of noble metal activity profile on catalyst lightoff. (say, the time required for 50% CO conversion) is pre-
Other parameter values are listed in Table I (case 1). dicted to be slightly shorter than the standard case, but
the catalyst’s activity before and after the lightoff is lower,
performance. When the inlet gas temperature is increased resulting in the deterioration of the overall performance
from 600 K to 700 K, for example, the order of catalyst (compare curves B and C).
lightoff is reversed; that is, faster lightoff occurs with higher Case 2. A Step Decrease in the Feed Temperature.
cell density, as illustrated in Figure 8. This can be at- In this case we examine temperature excursions encoun-
tributed to the fact that with increasing exhaust temper- tered in a catalytic monolith after a hot monolith has been
ature, the catalyst’s lightoff behavior is increasingly dom- exposed to a step flow of cool exhaust gas containing high
inated by the convective gas/solid heat transfer (hS(Tf -

concentrations of reactants. The possibility that such


T8) in eq 12) rather than by the heat capacity of the solid exhaust temperature perturbations can lead to a transient
((1 e)p8Cpa in eq 12). In this case, then, rapid catalyst
-

temperature rise in the solid phase above its initial tem-


lightoff is favored by high cell density as a result of the perature (that is, an overtemperature) has been pointed
attendant decrease in the channel size. (Note that both out by Young and Finlayson (1976b). The objective here
h and S are inversely proportional to the channel size.) is to identify conditions under which the overtemperature
Figure 9 compares the transient responses of catalytic phenomenon occurs and to determine the magnitude and
monoliths having different lengths, following a step in- location of the temperature peak as a function of converter
crease in the feedstream temperature (300 K to 600 K). design and operating parameters.
Here the frontal area of the monolith is kept constant (A As mentioned earlier, case 2 approximates a situation
= 60
cm2), and a fixed amount of Pt (i.e., constant a-A-L) which may occur in a fully warmed-up monolith when a
is distributed in the monoliths of three different lengths. vehicle suddenly decelerates. During deceleration of a
It can be seen that the catalyst lightoff is enhanced upon vehicle, both exhaust flow rate and temperature decrease.
decreasing the reactor length. Notice, however, that if the However, since the overtemperature phenomenon is caused
reactor length is too short, the catalyst’s activity after primarily by a reduction in the exhaust temperature, we
lightoff (i.e., the steady-state performance) is considerably will focus on its effect first. As will be shown later, a
degraded, as illustrated by the curve for L = 2.5 cm. simultaneous reduction of the exhaust flow rate and tem-
The computational results also show that the variation perature does not change the essential features of the
of the frontal area of the monolith (while keeping its length overtemperature phenomenon; that is, a reduction in the
and Pt content constant) similarly affects the monolith exhaust flow rate simply shifts the location of the wall
performance; that is, as the frontal area decreases, a tra- temperature peak toward the monolith inlet without sig-
deoff again exists between improving the lightoff behavior nificantly changing its magnitude.
and impairing the steady-state performance. Figure 11 shows how the wall temperature profile varies
It should be noted, however, that the simultaneous with time, following a step decrease in the feedstream
variation of the frontal area and length of the monolith temperature. The parameter values were taken to be the
while keeping its volume constant does not significantly same as those listed in Table I. As we expect, the upstream
change the lightoff characteristics of the monolith. This section of the monolith is constantly cooled by the cold
is not surprising, because the basic model equations, when feed exhaust. In the downstream portion of the monolith,
appropriately nondimensionalized, indicate that the however, a temperature peak develops above the initial
transient behavior of the monolith depends primarily on wall temperature (t = 2.1 s), grows in magnitude, and
the product of A and L, and not their individual values. moves toward the inlet of the monolith (t = 4.7 s). At t
Figure 10 shows how the lightoff behavior of a mono- = 10.1 s, the highest wall temperature ( 8- = 991
K) is
lithic catalyst is influenced by the Pt activity profile along attained at a normalized axial location of 0.675. This “hot
the reactor length. Both the linearly decreasing (curve A) spot”, of course, cannot sustain itself indefinitely; hence
Ind. Eng. Chem. Prod. Res. Dev., Vol. 21, No. 1, 1982 35

Figure 12. Time variation of gas-phase CO concentration profile


following a step decrease in the feedstream temperature, at the
standard conditions listed in Table I (case 2). Figure 13. Maximum wall temperature as a function of the local
Pt surface area. Other parameter values are listed in Table I (case
2).
the temperature peak moves downstream with decreasing
268.95 cm2 pt/cm3 Reactor
magnitude (t = 25.6 s and t = 64.7 s) and eventually leaves (A) a =

the reactor (t = 200 s).


The quantity Ta-max, shown in Figure 11, is the highest
wall temperature reached during the transient process.
This will be referred to as the maximum wall temperature,
and its magnitude and axial location will be a focal point
of our discussion in the sequel.
Notice in Figure 11 that the exhaust temperature after
the step change was chosen to be sufficiently low {Ttm =
300 K), so that the reaction is quenched over the entire
length of the monolith at steady state (see t = 200 s in
Figure 11). In this paper we will limit our attention to such
cases. The simulation results for Tg™ = 400 K showed
Figure 14. Effects of initial wall temperature on maximum wall
similar temperature excursions in the monolith following
temperature for two different local Pt surface areas. Other param-
the exhaust temperature perturbation. eter values are listed in Table I (case 2).
It is important to emphasize that the presence of H2 in
the feed is not a necessary condition for the occurrence of the gas/solid heat transfer process.
the temperature rise reported here; the model predicts It can be deduced from the above discussion that in
similar overtemperature phenomena even when the pres- order for the overtemperature phenomenon to occur, the
ence of H2 in the exhaust is not accounted for. As will be following two conditions must be satisfied simultaneously:
explained in the following paragraph, the transient tem- (1) sufficient cooling of the upstream section of the reactor
perature rise results from the combination of high exhaust (so that very little reactant is consumed by reactions taking
concentrations of combustible species and high initial place there), and (2) sufficient reactivity in the downstream
monolith temperature. section (so that the unconverted reactants carried from the
In order to understand this transient temperature rise, upstream section can be readily reacted to generate heat).
it is instructive to examine the gas-phase concentration The delicate nature of the process is illustrated in Figure
profile of the reactant (e.g., CO) as a function of time, for 13, where the maximum wall temperature encountered
the standard case considered in Figure 11. As Figure 12 during the transient is plotted as a function of the local
indicates, the gas-phase CO concentration in the upstream Pt surface area. Interestingly enough, the largest tem-
section of the monolith initially increases with time (com- perature rise is attained at an intermediate value of Pt
pare t = 0 and t = 2.1 s), because the reaction rate (and surface area. If the Pt surface area is too low, the rate of
thus the conversion of the reactant) in that region de- heat generation in the downstream portion of the reactor
creases as a result of the cooling of the catalytic wall by is not high enough to cause a significant temperature rise.
the cold exhaust. Because of the relatively large heat If the Pt surface area is too high, on the other hand,
capacity of the solid, however, the wall temperature in the considerable reaction occurs in the upstream section de-
downstream section of the reactor continues to remain spite the cooling in that region, so that only a relatively
high. As a result, the gas stream containing high concen- small amount of unconverted reactants contacts the hot
tration of the unconverted reactants eventually contacts catalytic wall in the downstream section of the monolith.
the hot catalytic wall in the downstream section of the For this reason, the resulting temperature rise is smaller
monolith. This leads to vigorous reactions in the down- than that corresponding to the intermediate Pt surface
stream section of the reactor, as evidenced by a rapid area, as illustrated in Figure 13. The sharp change in the
decrease in the gas-phase CO concentration profile at t = maximum wall temperature that occurs just on the left-
10.1 s. The attendant reaction exotherm, then, causes a hand side of the peak in Figure 13 can be attributed to the
transient temperature rise above the initial solid tem- fact that for this range of Pt surface areas, the maximum
perature. This condition persists until the downstream solid temperature is encountered at the reactor exit and
portion of the monolith is cooled by the exhaust far enough thus its magnitude is limited by the monolith length. Also,
so that a decrease in the rate of heat generation due to the we remark that besides its magnitude variation, the loca-
temperature decrease overshadows the competitive effect tion of the maximum wall temperature shifts toward the
of the increased reactant concentration. The appearance monolith inlet as the local Pt surface area is increased.
of the transient temperature rise is limited to only some Figure 14 shows how the initial wall temperature influ-
parts of the monolith as a result of the spatial variation ences the transient temperature rise following a step flow
of the competing effects between the heat generation and of cool exhaust gas at 300 K, for two different local Pt
36 Ind. Eng. Chem. Prod. Res. Dev., Vol. 21, No. 1, 1982

Figure 17. Time variation of wall temperature profile for a linearly


decreasing Pt activity profile. Other parameter values are listed in
Figure 15. Effects of monolith length on maximum wall tempera- Table I (case 2).
ture. Other parameter values are listed in Table I (case 2).
21050 -

1100 r

a 950 -

64'
| 850 -

!
65
s 750 -1-1 1 1_I_I_I_I_i I
0 0.2 0.4 0.6 0.8 1.0
Axial Position of Maximum Wall Temperature

Figure 18. Effects of exhaust flow rate on the magnitude and lo-
cation of maximum wall temperature. Other parameter values are
16. Effects of cell density on maximum wall temperature.
listed in Table I (case 2).
Figure
Other parameter values are listed in Table I (case 2).
fraction, the monolith with circular channels leads to a
surface areas. For each of the two Pt surface areas con- higher solid overtemperature than the square-channel
sidered, the highest wall temperature (and thus the largest monolith. This can be attributed to the fact that circular
temperature rise, as determined by the vertical distance channels provide more effective cooling of the upstream
between the solid and the dotted curves in Figure 14) is section of the monolith as a result of the higher gas/solid
reached at an intermediate initial wall temperature; both heat transfer coefficient.
higher and lower initial wall temperatures lead to lower The noble metal activity profile along the reactor length
wall temperatures during the transient. The similarity has an important effect on the solid temperature profile
between Figures 13 and 14 is not surprising, because the developed within the monolith (again, after a step decrease
Pt surface area and the specific reaction rate (which de- in the feed temperature). Figure 17 shows how the solid
pends on the wall temperature) appear in the model temperature profile varies with time when the Pt activity
equations as their product only. In addition, the regime profile is a linearly decreasing function of axial distance.
of overtemperature tends to shift toward higher initial wall In this case virtually no overtemperature is predicted, in
temperatures as the Pt surface area is decreased (e.g., by contrast to the considerable temperature rise (241 °C)
sintering). Notice also that the peak temperature of curve observed in the case of the uniform activity profile (see
B (lower Pt surface area) is higher than that of curve A. Figure 11). We remark that the total Pt surface area in
Figure 15 shows the effects of the monolith length (L) the monolith is the same for both cases. The results dis-
on the maximum wall temperature. Here, the length was played in Figure 17 suggest the possibility of minimizing
perturbed while keeping both the frontal area of the the temperature excursions in catalytic monoliths by a
monolith (A) and the Pt content in the reactor ( · ·£) proper choice of the noble metal activity profile.
constant at their standard values. As indicated in Figure During a sudden vehicle deceleration, the exhaust flow
15, a higher wall temperature is encountered following a rate is also rapidly decreased, in addition to a reduction
rapid decrease in the feed temperature, as the monolith in the exhaust temperature. Therefore, it is of practical
length is increased. It is interesting to note that the interest to investigate how the magnitude and location of
variation of the monolith frontal area (with reactor length the maximum wall temperature change with the exhaust
and Pt content fixed) similarly affects the temperature flow rate. Figure 18 summarizes the results of a set of
excursions of the monolith; that is to say, increasing the calculations for a range of exhaust flow rates. The same
frontal area tends to increase the transient temperature step decrease in the exhaust temperature as given in Table
rise. However, the simultaneous variation of frontal área I was considered in the calculations. For flow rates lower
and length while keeping their product (i.e., monolith than 60 g/s, the magnitude of the maximum wall tem-
volume) constant does not significantly change the mag- perature remains more or less constant, independent of the
nitude and normalized axial location of the maximum wall exhaust flow rate, whereas its location tends to shift closer
temperature. to the monolith inlet with decreasing flow rate. For flow
The magnitude of the temperature rise is also affected rates exceeding 60 g/s, the overtemperature phenomenon
by the cell density of the monolith. As shown in Figure becomes limited by the monolith length, so that the
16, the maximum wall temperature increases with in- maximum solid temperature is encountered at the reactor
creasing cell density. In perturbing the cell density, square exit and the magnitude of the temperature rise decreases
channels with a fixed wall thickness of 0.0254 cm were rapidly with increasing exhaust flow rate.
assumed. This results in the simultaneous variation of the The effect of the solid thermal conductivity was also
channel size and void fraction of the monolith, as illus- investigated by perturbing its value about the standard
trated in Table II for three selected cell densities. Also, value (see Table III). As might be expected, increasing
our calculations show that for a given channel size and void the thermal conductivity decreases the magnitude of the
Ind. Eng. Chem. Prod. Res. Dev., Vol. 21, No. 1, 1982 37

Table III. The Magnitude and Location of Maximum h = heat transfer coefficient, J/cm2-s-K
Wall Temperature as a Function of Solid Thermal (-Aff)¡ =
heat of combustion of species i, J/mol
Conductivity k¡ = rate constant for reaction i, mol-K/cm2-s
mass transfer coefficient for species i, cm/s
-

\s, J/cm-s-K ^sjnax) K xmaxlL km¿


K¡ = adsorption equilibrium constant, (mole fraction)'1
1.675 X 10'3 996 0.6625 L = length of monolith, cm
1.675 X 10'2 (standard) 991 0.6750 =
limiting Nusselt number, 2hRb/\g
1.675 X 10'1 975 0.7625 Fu¡t = total pressure, kPa
fig gas constant
=

fih = hydraulic radius of channel, 2(cross-sectional airea/wetted


_ perimeter), cm
fi, = specific reaction rate for species t, mol/cm2 Pt-s
S = geometric surface area per unit reactor volume, cm2/cm3
Sh„ = limiting Sherwood number, 2kmiRb/Di
t = time, s
T = temperature, K
T. = gas temperature, K
Tg™
= inlet gas temperature, K
T, solid temperature, K
=

T,o = initial temperature in the solid phase, K


T,m„ = maximum solid temperature, K
Figure 19. Temperature excursions in a monolith having a nonu- u = linear superficial velocity of exhaust gas, cm/s
niform initial wall temperature profile. Other parameter values are Wg
= mass flow rate of
gas, g/s
listed in Table I (case 2). x = axial coordinate, cm
*mai = axial location corresponding to the maximum solid
temperature rise during the transient. Interestingly, temperature, cm
however, the magnitude and location of the temperature
Greek Letters
peak were found to be rather insensitive to the variation
of the solid thermal conductivity, and this is indicated in c =void fraction of the monolith
Table III. 8 thermal conductivity of the gas, J/cm-s-K
=

We conclude this paper by considering the effects on X,


= thermal
conductivity of the solid, J/cm-s-K
=
pg gas density, g/cm3
transient converter response caused by nonuniformities
p, = solid density, g/cm3
in the initial wall temperature profile. Recall that our 4>¡(x)
=
piecewise linear Galerkin basis function
discussion so far has been limited to the case where the *( 8) = (1 eWCp, + T.idCVdT,)]
-

initial wall temperature profile is uniform over the entire


length of the monolith. Under certain operating condi- Literature Cited
tions, however, the reaction lights off at some point within Bird, R. B.; Stewart, W. E.; Ughtfoot, E. N. "Transport Phenomena"; John
Wiley: New York, 1960; p 505.
the monolith, resulting in a nonuniform axial wall tem- Cavendish, J. C.; Oh, S. H. Chem. Eng. J. 1979, 17, 41.
perature profile (e.g., Heck et al., 1976). Figure 19 shows Dabm, D. W.; Gentry, S. J.; Holland, . B.; Jones, A. J. Catal. 1979, 53, 164.
De Acetis, J.; Thodos, G. Ind. Eng. Chem. 1960, 52, 1003.
how the wall temperature profile varies with time after a Hansen, K. W.; Jorgensen, S. B. Adv. Chem. Ser. 1974, No. 133, 505.
monolith with a nonuniform initial wall temperature Heck, R. H.; Wei, J.; Katzer, J. R. AIChEJ. 1976, 22, 477.
profile (T80 = 550 K, x/L < 0.2; T,0 = 750 K, x/L > 0.2) Hegedus, L. L. AIChEJ. 1976, 21, 899.
Herod, D. M.; Nelson, . V.; Wang, W. M„ SAE Paper 730557, 1973.
has been exposed to a step flow of cold exhaust gas at 300 Hindmarsh, A. C. "GEARIB, Solution of Implicit Systems of Ordinary Differ-
K. As can be seen from the comparison of Figure 11 and ential Equations with Banded Jacobian", Lawrence Livermore Laboratory
Report UCID 30130, Feb 1976.
Figure 19, the nonuniform initial wall temperature profile Holberg, J. A.; Lyche, B. C.; Foss, A. S. AIChEJ. 1971, 17, 1434.
does not alter the essential features of the transient re- Kuo, J. C.; Morgan, C. R.; Lassen, H. G., SAE Paper 710289, 1971.
Mehta, P. S.; Sams, W. N.; Luss, D. AIChEJ. 1981, 27, 234.
sponse, except that the location of the maximum wall Mondt, J. R., SAE Paper 760320, 1976.
temperature is shifted somewhat downstream. Mondt, J. R„ private communication, General Motors Research Laboratories,
Warren, MI, 1981.
Nomenclature Morgan, C. R.; Carlson, D. W.; Vote, S. E., SAE Paper 730569, 1973.
Oh, S. H.; Cavendish, J. C.; Hegedus, L. L. AIChEJ. 1990, 26, 935.
a(x) catalytic surface area per unit reactor volume, cm2/cm3
=
Schlatter, J. C.; Chou, T. S. "Measuring Oxidation Rates In a Recycle
A = frontal
area of the monolith, cm2 Reactor"; presented at the AIChE 71st Annual Meeting, Miami Beach, FL,
B(Ta) = mass matrix in Galerkin method, eq 28 Nov 1978.
c,
= concentration of species i, mole fraction Shah, R. K.; London, A. L. "Laminar Flow Forced Convection Heat Transfer
and Flow Friction In a Straight and Curved Ducts—A Summary of Analyt-
cg-,
= concentration of species i in bulk gas stream, mole ical Solutions"; Technical Report No. 75, Department of Mechanical Engl·
fraction neerlng Stanford University, Stanford, CA, 1971.
c-/” = inlet concentration of species i, mole fraction
=
Sharma, C. S.; Hughes, R. Chem. Eng. Sol. 1979, 34, 625.
specific heat of gas, J/g-K Sherony, D. F.; Solbrlg, C. W. Int. J. Heat Mass Transfer 1970, 13, 145.
Cpg Stetter, J. R.; Blurton, K. F. Ind. Eng. Chem. Prod. Res. Dev. 1980, 19,
Cp¡
=
specific heat of solid, J/g-K 214.
Cy
= concentration of
species i in the solid phase, mole fraction Van Doesburg, H.; De Jong, W. A. Chem. Eng. Sel. 1976a, 31, 45.
c8
= vector with entries c,t¡, mole fraction Van Doesburg, H.; De Jong, W. A. Chem. Eng. Sol. 1976b, 31, 53.
Voltz, S. E.; Morgan, C. R.; Uederman, D.; Jacob, S. M. Ind. Eng. Chem.
D; =
diffusivity of species in the reactive mixture, cm2/s Prod. Res. Dev. 1973, 12, 294.
E= stiffness matrix in Galerkin method, eq 24 Wei, J. Adv. Catal. 1975, 24, 57.
F, = hS(T\ Ta) + a(x)E4i=i(-m¡Ri(ct,T,) Young, L. C.; Finlayson, B. A. AIChEJ. 1976a, 22, 331.
-

f2 =
hscra Tg)/(o„gCpg)
-
Young, L. C.; Finlayson, B. A. AIChEJ. 1976b, 22, 343.
[F,], =
-Am,¡S(Cg,,·
-

cjfi Received for review August 28, 1981


[F/h a(x)Ri(c„Tt) Pu>tkm,iS(cgj
= - -

c6ii) /(RgTg)
G =
quantity defined by eq 4 Accepted November 2,1981

You might also like