You are on page 1of 6

Nomenclature , =

density of catalyst particles, lb./cu. ft.


Pstp =
density of gas at 1 atm., 32° F., lb./cu. ft.
a, =
superficial surface of catalyst, surface per unit volume =
time, hr.
of bed, sq. ft./cu. ft.
= heat capacity of gas at constant pressure, B.t.u./lb. ° R.
=
space velocity at 1 atm., 32° F., cu. ft./cu. ft./hr.
Cg
C, =
heat capacity of solid, B.t.u./lb. ° R. Subscripts
G = rate of mass flow of gas, lb./sq. ft. hr. i at axial point i inside packed bed
=
h =
heat transfer coefficient between gas and solid, B.t.u./
°
sq. ft. hr. R. Superscripts
kei =
equivalent thermal conductivity of gas based on total '
cross section of convertor (keg 5kg), B.t.u./ft. hr.
=
=
quantity at time t + At
°

kg
=
thermal conductivity of gas, B.t.u./ft. hr. ° R. Literature Cited
ka =
equivalent point-contact thermal conductivity of solid,
° Anzelius, A., Z. Angew. Math. Mech. 6, 291 (1926).
B.t.u./ft. hr. R. Bird, R. B., Stewart, W. E., Lightfoot, E. N., “Transport Phe-
L =
depth of catalyst bed, ft. nomena,” p. 411, Wiley, New York, 1960.
Pr = Prandtl number for gas, µCg/ka Furnas, C. C., Trans. Am. Inst. Chem. Engrs. 24, 142 (1930).
Qr = heat generation due to a chemical reaction, B.t.u./ Heliums, J. D., Churchill, S. W., “International Development in
cu. ft. hr. Heat Transfer,” p. 985, Am. Soc. Mech. Engrs., New York,
Re =
Reynolds number, G/ , µ). 1961.
S =
cross-sectional area of convertor, sq. ft. Hodgman, C. D., et al., “Handbook of Chemistry and Physics,”
t =
“time” (T —

T/PesCs), hr.
0
F. cu. ft./B.t.u. p. 2267, Chemical Rubber Publishing Co., Cleveland, 1962.
0 Maga, J. A., Kinosian, J. R., “Motor Vehicle Standards, Present
Tav =
average catalyst bed temperature [J'q T(t,z) dz], R. and Future,” SAE Automotive Engineering Congress, Detroit,
°
=
gas temperature, R. 1966.
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

Tg
Mickley, H. S., Sherwood, T. K., Reed, C. E., “Applied Mathe-
°
=
gas temperature at ° z
=
0, R.
T, solid temperature, R. matics in Chemical Engineering,” p. 376, McGraw-Hill, New
Downloaded via INDIAN INST OF SCIENCE on November 19, 2019 at 11:31:10 (UTC).

0 = initial solid temperature, R. 0 York, 1957.


x =
axial distance, ft. Schumann, T. E. W., J. Franklin Inst. 208, 405 (1929).
dimensionless axial distance, z Smith, J. D., “Chemical Engineering Kinetics,” p. 391, McGraw-
z = =
x/L Hill, New York, 1956.
State of California Motor Vehicle Pollution Control Board, Los
Greek Letters Angeles, Calif., “California Procedure for Testing Motor Vehicle
Exhaust Emissions,” 1964.
5 = void fraction of bed Willmott, A. J., Intern. J. Heat Mass Transfer 7, 1291 (1964).
°
At = size of “time” step, hr. F. cu. ft./B.t.u.
=
size of axial interval Received for review July 11, 1966
µ =
viscosity of nitrogen, lb.ma55/ft. hr. Resubmitted July 12, 1967
=
shape factor for catalyst particles Accepted September 21, 1967
peg
=
equivalent density of gas, (peQ =
5pa), lb./cu. ft.
Pe,
=
density of catalyst bed [pes =
(1 5)p,], lb./cu. ft.

Division of Petroleum Chemistry, 152nd Meeting, ACS, New


pa
=
density of gas, lb./cu. ft. York, N. Y., September 1966.

A MODEL OF CATALYTIC CRACKING


CONVERSION IN FIXED, MOVING, AND
FLUID-BED REACTORS
V E R N W .
WEEKMAN, JR.
Applied Research & Development Division, Mobil Research and Development Corp., Paulsboro, N.J.

A deal of exploratory work in catalytic cracking


great (1959) compared the various reactor types. Froment and
**
takes place in fixed-bed reactors, which are cheap and Bischoff (1961) treated catalyst decay in fixed-bed systems
easy to operate on a laboratory scale. Most potential indus- which result from series or parallel fouling reactions. Masa-
trial applications, however, lie in moving or fluid-bed reactors. mune and Smith (1966) described theoretically the interaction
The present work provides a method for relating catalytic between catalyst fouling and diffusion.
cracking conversion from one reactor configuration to another, This paper describes the development of mathematical
based on the principles of reaction kinetics. models that account for the temporary decay of the catalyst
Within the last 20 years various mathematical models have during use. The models are presented in terms of a dimen-
been proposed for catalytic cracking. Voorhies (1945) sionless decay group and a dimensionless reaction group.
presented a model of coking and conversion behavior. Bland- Comparisons among isothermal, fixed, fluid, or moving-bed
ing (1953) gave a model of cracking behavior based on simple reactors reduce simply to comparisons among the appropriate
reaction kinetics. Using the Voorhies expressions, Andrews dimensionless groups. Finally the proposed model is shown

90 l&EC PROCESS DESIGN AND DEVELOPMENT


The conversion of gas oils during catalytic cracking may be represented by a pseudo-second-order reaction
coupled with a first-order decay of the catalyst activity. For plug flow in the gas phase and for vapor
residence times small in relation to the catalyst decay times, it was possible to solve the defining partial
differential equation to yield isothermal models for conversion in fixed, fluid, and moving bed reactors.
Comparisons of these models and cracking results show that they successfully represent the experimental data
over wide ranges of operating conditions. The models are presented in terms of two dimensionless groups:
an extent of reaction group and an extent of catalyst decay group. Comparison of various reactor types
simply becomes a comparison of the appropriate dimensionless groups.

to represent experimental catalytic cracking data very closely. The nonlinear coefficient, B, represents the ratio of oil
The conversion of gas oil is defined here as the fraction of the transit time to the time of catalyst decay, tm. For many
original charge material which has been cracked to the gasoline gaseous reactions, the time of vapor transit will be negligible
molecular weight range or lighter. Since most catalytic compared to the time of the fixed bed experiment—i.e., the
cracking feed stocks have had the gasoline constituents dis- time of catalyst decay—and thus B = 0. If the characteristic
tilled off, conversion reflects fairly well the actual “work” B is essentially zero, we may integrate the following ordinary
of the catalyst. differential equation along the characteristic a constant,
=

provided that the catalyst decay is not extremely rapid.


Discussion
fPv
Defining Equation. An isothermal reaction taking place
dy_
dx PiS
RM (6)
in piston flow through a reactor containing a time-decaying,
non-diffusion-limited catalyst may be described in terms of the In less formal terms we may say that the members of any
weight fraction unconverted, y, by the following continuity given group of oil molecules traverse the bed so fast that they
equation: see catalyst of essentially the same age. Thus decay of the
catalyst is slow relative to the vapor residence time. Equation
l^y) + Uv
dy
(1) 6, then, describes the conversion of the reactant in terms of
Pv & dz both time and axial distance.
The reaction rate-decay term R(y,t) will be a function of the For very low vapor velocities and/or very short catalyst
contact times, the characteristic direction B may not be ap-
amount remaining unconverted and the time during which
decay has taken place. If there is molar expansion, as during proximated by zero and Equation 6 can then be integrated
catalytic cracking, both the vapor density, p„ and the velocity, along the nonlinear characteristic given by Equation 4.
UV) will be functions of the fraction of unconverted charge, y. Fortunately, most gas-phase cracking reactions are run under
For steady-state moving-bed and fluid-bed reactors, the time conditions for which the characteristic direction B is well
partial derivative is, of course, zero. For fixed beds one must, approximated by zero.
Reaction Rate Expression. As the catalyst activity
however, treat the full partial differential equation.
Fixed Beds. Equation 1 may be recast in dimensionless changes with time, the kinetic rate expression R(y, ) will
form by defining a normalized time, based on the catalyst de- be a function of both normalized time of exposure to oil, ,
and the fraction of unconverted reactant, y. The rate ex-
cay time, and a normalized distance, based on the over-all
reactor length. If the molecular weight of the reactants al- pression is complicated further because the gas oil reactivity is
itself a function of conversion. This depth-of-cracking effect
ways bears a fixed relation to that of the products, then (in
terms of the ratio of reactant to product molecular weights, a, is caused by the multiplicity of molecular types present in
and the initial vapor density, p0) the actual vapor density is the original charge. Thus, the first molecules to crack have a
much higher reaction velocity than molecules that crack
simply
subsequently. The net effect of such a phenomenon on a
Po pseudo-first-order reaction is to increase the apparent order of
(2) the reaction, since the average reactivity of the unconverted
y + a(\ -

y)
feed will decrease asy decreases. If the vapor density is treated
The normalized version of Equation 1, including Equation 2, as a constant, the apparent order also increases, because the
may be written in terms of liquid hourly space velocity, actual density decreases as the fraction unconverted gas oil,
S =
Fa/(Vrpi) y, decreases.
A pseudo-second-order reaction could satisfactorily account
^
B^ + d*T =
fPv
Riyfi) (3) for these nonlinear effects; Blanding (1953) also found that
PiS
similar kinetics described his data.
where It was found that the rate of decay of catalyst activity was
proportional to the remaining activity—i.e., simple first-
_hoa_ (4) order decay in time. It was also the simplest decay function
PiStm\y + a( 1

y)]2 that gave a good representation of experimental data. Figure


1 compares the first-order decay function to some hexadecane
Inspection of Equation 3 shows that the characteristic
direction in the x, plane is cracking data obtained by Nace (1965) in a fixed-bed reactor.
After approximately 20 seconds the decay is essentially first-
order. Defining an intrinsic reaction velocity constant as
(5)
k0 and a decay velocity constant as a gives the following form

VOL. 7 NO. 1 JANUARY 19 68 91


O 20 40 60 80 100 120
TIME-ON-STREAM, SECS
Figure 2. Time-averaged conversion for fixed beds

Figure 1. Rate of catalyst decay


no replacement of solid phase but with continuous piston
Data obtained from Nace (1 965) flow—can also be characterized by Equation 10.
gas
As the instantaneous reaction rate term, Equation 7, is in-
of the reaction rate term: dependent of reactor configuration, this term applies equally
well to moving or fluid beds. Of course, the time of decay
R(yfi) =
k0y*e-™ (7) now becomes the catalyst residence time in moving beds and

= the mean catalyst residence time in fluid beds.


cttm
Moving Beds. For moving beds the time of catalyst decay
The reaction velocity constant, k0, and the decay velocity con- at any position is the catalyst residence time for flow through
stant, a, depend on the charge stock-catalyst combination the entire reactor multiplied by the fractional distance, x,
but are independent of the conversion level. The initial traversed. Thus, for a steady-state moving bed piston flow
charge concentration has been adsorbed into the reaction reactor Equation 3 becomes:
velocity constant, k0, since for any given charge material k0
must be determined experimentally. =
(11)
~ -

Ay*e~Xx
The dimensionless decay group, , is the product of the dx
decay velocity, a, and the total time of decay, tm. Thus it As before, the decay group, is the decay
X, velocity, a,
represents the “length” of decay. Equation 6 may now be
written as:
multiplied by the catalyst residence time, tm, for transit through
the entire reactor. Under the boundary condition, y(0) 1, =

the solution of Equation 11 in terms of the instantaneous con-


%
ax
=
-Ayh-™ (8) version (e
= 1 —

y) is

where A( 1
-

e~u)
(12)
X + A(\ -

e-**)
.
_fppkp _
Ko
At the bed exit (x
~ ~

P,S S

1) this expression becomes

The dimensionless reaction velocity group, A, is the reaction


4
A( 1
-

)
~
(13)
velocity multiplied by the vapor phase residence time and X + A(1 -

<·" )
represents the “length” of reaction.
Solution of Equation 8 along the characteristic (dd/dx) = 0, Figure 3 presents a working plot of Equation 13.
Fluid Beds. By assuming one has piston gas phase flow
with boundary conditiony (0) 1 for all ,
gives: =

and perfectiy mixed solids flow, an analytic solution of Equation


1

y ~

(9)
1 + Axe~M

The time-averaged conversion obtained by collecting and


mixing the product flowing from the reactor (x 1) is: =

e (9A)

Evaluation of this integral yields the desired solution for


time-averaged conversion in fixed beds in terms of the dimen-
sionless decay and reaction groups.

1 + A
6 (10)
Ji°g. 1 + Ae~x_ 5 10 15 20 25 30 35 40
EXTENT OF REACTION, A =
K0/S
Figure presents the time-averaged conversion as a function
2
of the decay and reaction groups. A fixed fluidized bed—i.e., Figure 3. Conversion in moving beds

92 l&EC PROCESS DESIGN AND DEVELOPMENT


3 may also be obtained for fluid beds. If 779 represents the This is also, of course, the steady-state solution of Equation 3.
fraction of catalyst particles with ages between 9 and 9 + 79 Equations 10, 13, and 18 have been used to predict the ratio
(internal age distribution), then an average reaction velocity of conversion for the different reactors as a function of the length
constant may be defined as follows: of decay and length of reaction groups. Figure 5 compares
fixed and moving beds and reveals that for first-order decay
k =
(14) at the same A and X, moving beds will always give more con-
version than fixed beds. As the decay group becomes larger,
For perfect mixing the internal age distribution, 7(9), is fixed beds become increasingly less attractive.
simple, e~e. Evaluation of the above integral for perfect Figure 6 compares fluid and moving beds and shows that,
mixing gives the following expression for the average reaction for first-order decay at the same A and X, moving beds always
velocity constant: give more conversion than fluid beds. For high rates of decay
they become almost identical. Likewise, as the reaction group
K A increases, they give almost identical conversion. Although
(15)
1 +
moving beds give more conversion at the same A and X, the
For steady-state fluid beds, Equation 3 now becomes fluid bed generally can be run at much higher catalyst-oil
ratios than moving beds—i.e., lower X—and they are also less
A
dy 2 subject to diffusion limitations—i.e., higher effective A.
jr~v+iy (16)
Figure 7 compares fixed and fluid beds and shows that at
low values of X and A fixed beds with first-order decay give
For the boundary condition, > (0) 1
Equation 16 gives the
=

slightly more conversion, but at higher values of the decay


instantaneous conversion (e 1
y) in terms of decay and
= —

reaction groups group, fluid beds give much higher conversions. As reaction
group A gets larger, the curves begin to cross over, since as
Ax A — oo the conversion in all three reactor types approaches
(17)
1 -f- -f- Ax 1.0. This effect can also be seen in Figures 5, 6, and 7.

At the bed exit {x = this expression becomes simply Rapid decay by metals or high molecular weight aromatics
1) found in heavy gas oils may change the first-order nature of the
A decay.
(18) The generalized dimensionless plots given by Figures 2,
1 + + A
3, and 4 are also convenient for comparing processing condi-
Figure 4 provides a working plot of Equation 18 in terms tions required to achieve a given constant conversion level.
of A and X. The catalyst decay occurring is assumed to pro-
ceed uniformly over the length of the various reactors; how-
ever, with high metal content feeds or highly asphaltic feeds,
the decay may not be uniform, requiring a modification in the
model.
Comparison of Fixed, Fluid, and Moving-Bed Reactors.
If there were no decay of catalyst activity, any vapor plug flow
reactor would give the same conversion, regardless of catalyst
contact time. For the fixed-bed reactor, time-averaged and
instantaneous conversion would be identical for a nondecaying
catalyst.
Application of l’Hospital’s rule to Equations 10 and 13
and inspection of Equation 18 reveal that the fixed, moving,
and fluid-bed models all have the following limit as the catalyst
decay goes to zero—i.e., X —* 0.
0 2 4 6 8 10 12 14 16
A EXTENT OF REACTION, A
Lim Eq. 10, 13, 18 =
-—;—- (I) Time-averaged conversion
x—o 1 + A
Figure 5. Ratio of fixed-bed to moving-bed conversion

Figure 4. Conversion in fluid beds

VOL. 7 NO. 1 JANUARY 1 9 68 93


Comparisons of reactor variables reduce to comparisons of the
appropriate dimensionless decay and reaction groups. The
following four examples show just how simple these comparisons
are.

Compare reactor volumes required for the same conversion,


catalyst residence time (constant ), and oil throughput.
(Vol. reactor), _
A,
(Vol. reactor)2 A2

Compare catalyst/oil ratios required for the same conversion


and space velocity (constant A).
(Cat./oil), _
\2
(Cat./oil)2 ,
Compare space velocities required for the same conversion
and catalyst residence time (constant X).
O 5 10 15 20
EXTENT OF DECAY, X (Space velocity), _
A2
(l) TIME-AVERAGED CONVERSION (Space velocity)2 A,

Figure 7. Ratio of fixed-bed to fluid-bed conversion Compare catalyst residence times required for the same
conversion and space velocity (constant A).
(Cat. res. time), _
X,
(Cat. res. time)2 X2

As the reaction velocity constant, K0, and the decay velocity


constant, a, are intrinsic parts of Figures 2, 3, and 4, they may
also be used to compare catalysts and charge stocks. Both
constants reflect the combined effect of charge stock and
catalyst. For example, the amount of deactivation a catalyst
has undergone during operation could be determined by mak-
ing two identical runs: one on the fresh and one on the aged
catalyst. The fractional activity loss would simply be the
ratio of the activity groups, A, read from the appropriate
reactor chart.

-J—.............—1-
Rating of catalysts based only on their activity can be
o
hazardous, particularly when the test is made in a fixed bed
' '

0 10 20 30 40 50 60 70
S, Liquid Hourly Space Velocity, V/(V)(hr) unit and the application will be in a moving or fluid bed system.
Table I shows how two hypothetical catalysts with realistic
Figure 8. Comparison of fixed-bed model and fixed-
values of a and K0 rank differently in terms of conversion de-
bed zeolite catalyst data
pending on the type of reactor used. A more complete de-
Charge. Mid-Continent gas oil
Catalyst temp. 900° F.
scription of cracking catalyst must include its decay as well
1
Equation 10 as its reaction velocity.

Figure 9. Comparison of fixed-bed model and fixed-bed zeolite catalyst


data
Charge. Mid-Continent gas oil
Catalyst temp. 900° F.
————
Equation 10

94 I & E C PROCESS DESIGN AND DEVELOPMENT


The moving-bed form of the model was checked against
Table I. Hypothetical Comparison of Two Catalysts in data from a laboratory moving-bed unit. Table II compares
Fixed, Moving, and Fluid Bed Reactors the model with the experimental data. Again the two re-
Catalyst 1. K„ 30 hr.-1 a 20 hr.-1 =

quired constants were obtained with a fitting technique.


Catalyst 2. Ka 5 hr. 1
a = 2 hr. 1

The fixed-bed data were obtained by operating the moving-


Fraction Conversion
Catalyst bed reactor as a fixed-bed reactor.
No. S A Fixeda Fluid Using only the constants
tm Moving
obtained from the moving-bed runs, the fixed-bed model pre-
1 1 0.5 30 10 0.34 0.75 0.73
2 1 0.5 2. 1 0.59 0.61 0.55 dicted the actual time-averaged experimental fixed bed reason-
a
Time-averaged. ably well; the maximum deviation was only 4% conversion.

Nomenclature
a = ratio of reactant to product molecular wt., Mr/Mp
A dimensionless reaction group, ratio of reaction to
Comparison of Models to Experimental Data. Compari-
=

son of the fixed bed model with a large number of fixed-bed space velocity, p0fk0/p¡S
=
0/ß / \ + a(\ y)]2 —

catalytic cracking runs revealed that the analytic model was / =


fraction voids in catalyst bed
capable of correlating a wide range of conditions. The fixed- F0 = lb. reactant/hr.
bed reactors employed were the same as those used for routine kc =
coking constants
k0 =
intrinsic reaction velocity constant at 0, hr.-1
catalyst evaluations in this laboratory and have been described
=

(includes initial charge concentration because of


(Plank et al., 1964). Only two fitted constants were necessary second-order reaction)
—namely, those for the reaction and the decay velocity. F-o Pofko/p i

Figure 8 compares the experimental fixed-bed gas oil cracking R(y, ) =


instantaneous kinetic rate expression
data (Wojciechowski, 1964) with the fixed-bed model and a S =
liquid hourly space velocity, vol./(vol.) (hr.)
t =
clock time, hr.
zeolite catalyst. The model successfully correlates the data
tm =
catalyst residence time, hr.
over a 16-fold range in space velocity and a 64-fold range in T 0
=
temperature, R.
run time. The two required constants were fitted by a non- Uv =
velocity of oil vapor, ft./hr.
linear optimization technique using a least squares criterion Vr =
volume of reactor, cu. ft.
x =
normalized axial distance, z/z0
(Wheeling and Kelly, 1962). instantaneous weight fraction charge unconverted
y =

Figure 9 presents fixed-bed gas oil cracking data (Coonradt, y =


time-averaged weight fraction charge unconverted
1964) plotted as space velocity vs. the catalyst-oil ratio for a z =
axial distance in reactor, ft.
slightly different zeolite catalyst. The catalyst-oil ratio is the z„ =
total reactor length, ft.
total catalyst volume divided by the total volume of oil pumped
Greek Letters
during the run time—i.e., tm l/ß S. Again, the model
=

successfully correlates the data over a wide range of conditions. a =


decay velocity constant, hr.-1
It also predicts a maximum in conversion as space velocity is ß =
catalyst to oil ratio, vol. of cat./vol. of total oil for
fixed beds, (vol. of cat./hr.)/(vol. of oil/hr.) for
increased at constant catalyst-oil ratio. This phenomenon is
moving and fluid beds
typical of highly active yet rapidly decaying catalysts. Com- e =
instantaneous weight fraction converted to gasoline
parison of the two catalysts shows that the second one (Figure 9) and lighter
is slightly less active and decays at a slower rate.
=
time-averaged weight fraction converted to gasoline
and lighter
=
normalized time-on-stream, t/tm
=
dimensionless decay group, a/ß S or atm
p0 =
initial charge density at reactor conditions, lb/cu. ft.
Table II. Comparison of Models to Experimental Data
=
density of reactor vapor phase, lb./cu. ft.
(Mid-continent gas oil, commercial TCC catalyst)
Pi =
density of liquid charge at room temperature,
lb./cu. ft.
a =2.96 hr.-1
K0 = 1069 e~e'RT 2.9 at 900° F.
=

Q
=
9000 cal./g.-mole Literature Cited
LHSV, Andrews, J. M., Ind. Eng. Chem. 51, 507 (1959).
V./(V.) Cat.f Oil. Cat. Res. Temp.. Blanding, F. H., Ind. Eng. Chem. 45, 1186 (1953).
Coonradt, H. L., Research Department, Mobil Oil Corp., Pauls-
°
(Hr.) v./v. Time, Min. F. Model Exptl.
Moving Bed boro, N. J., private communication, 1964.
Froment, G. F., Bischoff, K. B., Chem. Eng. Sci. 16, 189 (1961).
2 4 7.5 900 0.53 0.55 Masamune, S., Smith, J. M., A.I.Ch.E.J. 12, No. 2, 384 (1966).
4 7.5 950 0.58 0.58 Nace, D. M., Research Department, Mobil Oil Corp., Paulsboro,
2 15 950 0.53 0.50 N. J., private communication, 1965.
1 30 950 0.47 0.50 Plank, C. J., Rosinski, E. J., Hawthorne, W. P., End. Eng. Chem.
1 30 905 0.42 0.41 Prod. Res. Develop. 3, 165 (1964).
2 15 903 0.48 0.45 Voorhies, A., Jr., Ind. Eng. Chem. 37, 318 (1945).
4 7.5 850 0.47 0.49 Wheeling, R. F., Kelly, R. J., Research Department, Mobil
1 30 950 0.47 0.47 Oil Corp., Princeton, N. J., private communication, 1962.
Fixed Bed Wojciechowski, B. W., Research Department, Mobil Oil Corp.,
Paulsboro, N. J., private communication, 1964.
(Time-Averaged)
4 0.75 20 900 0.32 0.33 Received for review May 1, 1967
0.375 40 900 0.25 0.27 Accepted September 29, 1967
1.5 10 900 0.36 0.39
2 1.5 20 900 0.47 0.43 Division of Petroleum Chemistry, 154th Meeting, ACS, Chicago,
111., September 1967.

VOL 7 NO. 1 JANUARY 1968 95

You might also like