Sie sind auf Seite 1von 7

differ in molecular composition, particularly with respect The Voorhies (1945) relationship between catalyst coke and

to olefins, from virgin gas oils. From the linearity of the data catalyst residence time can be written as,
for the virgin gas oils, the following relationships can be de-
rived : C = atcb
where C = catalyst coke, t, = catalyst residence time, aiid a
and b are constants. The coefficient a is to a significant extent
a measure of the coking tendency of a charge stock. I t is
reasonable to expect some correlations e u t between a and the
catalyst decay constants. Figures 7 and 8 are, respectively,
plots of a vs. a and n vs. a. Both of the catalyst decay con-
From Figures 2 and 3, thevalues of a. and bo are 25 and -0.42, stants inciease n i t h incieasing a, which agrees with the trend
respectively, and the corresponding ones for a1 and bl are 20 shown 011 Figure 4.
and -0.46 for the range of data shown. These values were
calculated by least-squares regression of the data for virgin Summary and Conclusions
gas oils. Relationships of K S to aromatic/iiapht.hene ratio did
The kinetic rate constants for the catalytic cracking of gas
not correlate as well as the other rate constants. However,
oils and subsequent gasoline formation can be correlated with
K z does not change drastically wit,h composition (Nace e t al.,
the aromatic/naphthene ratio for a wide range of virgin
1971), and it typically varies only from 1.5 to 2.5 at 900F.
charge stocks. The catalyst decay constant also correlates
The relationships between a , K O ,and K 1 vs. the ratio of
directly Kith the aromatic/naphthene ratio. Finally, gas oils
aromat,icsjiiaphthenesas represented in Figures 1-3 appear to
whose composition tends to produce more coke also give lower
be independent of the level of the paraffin concentrations in
cracking and gasoline formation rates and higher catalyst
the charge stocks. The data in these figures cover reasonably
decay rates.
wide ranges of paraffin conceiit,rations.
Since catalyst decay in catalytic cracking is considered to literature Cited
be intimately associated wit,h coke formation, it is interesting
Akhmetshina, M. N., Levinter, ?*I.E., Tanatarou, 11.A., hlocha-
to examine the relationship between the catalyst decay con- lov, Y. D., Aeftepererab.Aeftekhim. (hloscow), 1969 (8), 10.
stant, a . aiid cat.alyst coke. Figures 4-6 are plots of a, &, Crocoll, J . F., Jaquay, R. D., Petro/Chem Eng., C-24, November
aiid K l vs. catalyst coke and show a increases and K o and K I 1960.
Fitzgerald, M. E., RIoirano, J. L., Morgan, H., Cirillo, V. A., App.
decrease with increased catalyst coke. Spectrosc., 24, 106 (1970).
The relationship of the coke deposited to the decay constant Nace. D. AI., Voltz. S. E.. Weekman. Jr.. V. W.. Ind. Eno. Chem.
is to be c.spectmedsince coke is the major contributor to catalyst Process Des. Develop., 16, 330 (1971).
Reif, H. E., Kress, R. F., Smith, J. S., Petrol. Refiner, 40 ( 5 ) ,
decay. The decay constant, a , however, reflects all causes of 237 (1961).
decay such as the adsorption of aromatics and nitrogen com- Service, W. J., PetrolChem Eng., C-32, November 1960.
Shnaider, G. S.,llukhin, I. I., Chueva, R I . A , , Kogan, Y. S.,
pounds as well as t.hat coke which remairis after st.ripping. Kham. Tekhnol. Topl. Masel, 1969 (I), 10.
From Figure 5 and 6, it is clear that thosecharge stocks which Voorhies, A., Ind Eng. Chem., 37,318 (1945).
produce more coke also have a decreased cracking and gasoline Weekman, V. W., Ind. Eng. Chem. Process Des. Dezelop., 7, 90
11968)
format,ion tendency. Iiit,erestingly, the slopes of KO and KI Weekman, 1.W., ibid., 8 , 385 (1969).
with coke are approsimately equal indicating the init,ial Weekman, V.W., Nace, D. hl., AIChE J., 16, 397 (1970).
selectivity ( K l j K o ) is largely independent of coke make. White. P. J.. PreDrint No. 24-68. API Division of Refining. 33rd
Y

3lid-Yearhlee&ng,Philadelphia, Pa., May 1968.


,lverage values of t,he nonstrippable catalyst coke in experi-
ments wit8hcatalyst residence times of 5.0 minutes were used for review October 29, 1970
RECEIVED
i n these plots (Kace e t al., 1971). ACCEPTEDMay 3, 1971

Tests for Transport limitations in Experimental


Catalytic Reactors
David E. Mears
Union Oil Go. of California, Union Research Center, Brea, Calif. 9662i

I n esperiment,al studies of heterogeneously catalyzed in an experimental reactor reflect only chemical events,
reactions, one of the first objectives should be to determine gradients must be virtually eliminated from three domains:
whether the investigation is coriceriied with catalytic kinetics Intraparticle within individual catalyst particle.3
or with interactions between the kinetics and transport Interphase between the external surface of the particles
pheiiomena. T h e intrusion of temperature aiid concentration and fluid adjacent to them
gradient,s can lead t o severe deviations in a catalysts per- Interparticle between the local fluid regions or catalyst
formance, in many cases completely disguising the true particles
kinetics of the reaction. To ensure that kinetic data obtained The latter doniaiii equally well could be called intrareactor

Ind. Eng. Chem. Process Des. Develop., Vol. 10, No. 4, 1971 541
since it applies to gradients within the reactor as a whole. where aisand Rots are, respectively, the rate expression and
While the intraparticle domain has been extensively analyzed its first derivative with respect to reactant concentration
for several decades, the latter two domains only recently evaluated a t the external surface. A numerical value of 0.75
have received much attention. appearing on the right side of the equation was rounded to 1
Criteria are n o x available to estimate whether transport in this derivation. This criterion can be extended to negative-
effects in any domain are significantly altering the experi- order reactions by taking the absolute value of the derivative.
mental results. I n deriving such criteria, perturbation analysis For simple power-law kinetics (as= kc,"), the perturba-
has proved particularly useful for detecting when deviations tion criterion becomes:
of the rate start to become appreciable. The method involves
expanding the reaction rate expression in Taylor series in
(3)
concentration, temperature, or both. Criteria involving
various dimensionless Damkohler numbers (Damkohler, where n is the reaction order. A nearly identical result was
1936) are then obtained which will hold deviations to a n recently obtained by Stewart and Villadsen (1969) as an
acceptable tolerance. This paper compares perturbation illustration of their orthogonal collocation procedure, which
criteria with those derived b y other approaches, and evaluates is also applicable to other geometries and kinetics. Equations
means of minimizing transport limitations in experimental 2 and 3 fail for zero-order kinetics because the concentration
reactors. deviations are beyond the scope of the perturbation approach.
Intraparticle Transport For this case, the value of 6 on the right obtained by Weisz
is recommended.
Intraparticle transport has been analyzed for a wide For reactions involving a change in the number of moles,
variety of reaction kinetics, particle geometries, and thermal the effectiveness factor relationships are shifted as a function
behavior, as well-reviewed elsewhere (Petersen, 196513; Satter- of the transport mechanism. When the mode of transport
field and Sherwood, 1963). Generally the objective has been is Knudsen diffusion, Otani et al. (1964) showed that a
t'o calculate an effectiveness factor, 7 , defined as the ratio pressure gradient develops in the catalyst, but the shift
of the actual rate to that which would occur if the temperature in effectiveness factors is generally negligible. When molecular
and concentration were constant throughout the catalyst diffusion prevails, a decrease i n moles enhances the transport
particle. The solutions show that TJ becomes inversely pro- while an increase hinders it. Weekman and Gorring (1965)
portional to the characteristic dimension of the particle in showed that the influence of a volume expansion or contrac-
the regime of strong diffusion influence. Cnfortunately, tion can be expressed in terms of the volume-change modulus:
direct application of such solutions requires knowledge of the
true kinetic behavior and the intrinsic rate constant. e = (. - i ) ~ , (4)
Jf7eisz and Prater (1954) presented a criterion for the
absence of significant diffusion effects with irreversible in which Y , is the mole fraction of reactant a t the catalyst
reactions which is independent of the rate constant. I t is surface and v is the stoichiometric coefficient in the reaction
assumed that Fick's first law governs diffusion in the porous A + vB.Following Kubota e t al. (1969), the criterion for an
media, that the effective diffusivity, D e , remains independent isothermal particle can be modified to:
of the nature of the reaction, and that the intrinsic catalytic
activity is distributed uniformly throughout. T o ensure 7 2 (5)
0.95 in an isothermal spherical particle with a first-order
reaction, the criterion requires:
I n dilute cases ( Y s<< l ) , the volume-change effect becomes
negligible.
Anderson (1963) applied the perturbation approach t o
derive a criterion for the lack of importance of temperature
where is the observed reaction rate per unit particle volume, gradients in catalyst particles. The reaction is assumed to
C, is the reactant concentration a t the external surface of the follow an Arrhenius temperature dependence. For quasi-
particle, and r p is the radius of the particle. The dimensionless isothermal behavior, the observed rate must not differ
group, called Damkohler's number for diffusive transport, from the rate that would prevail a t constant temperature
expresses the ratio of chemical reaction rate to diffusive b y more than an acceptable amount, say 5%. The resulting
flus. The merit of the criterion is that it is stated in terms of criterion, in terms of Damkohler's number involving heat
observables, and a parameter, D e , which is independently transport by conduction, is
measurable. It also applies t o transport of a gas in liquid-filled
pores if the effective diffusivity is correctly evaluated (Satter-
field et al., 1968). Weisz (1957) later showed by means of a
linear approximation of the concentration gradient a t the
surface that the numerical value on the right is conservatively where \ A H ] is the absolute value of the heat of reaction, X
0.3 for second-order reactions and 6 for zero order. The is the thermal conductivity of the particle, E is the true
corresponding value for first-order reactions is 0.6 by this activation energy, R is the gas constant, and Tsis the absolute
approsirnation. temperature a t the catalyst surface. Again a numerical
When strong inhibition by the reaction product occurs, value of 0.75 on the right has been rounded to 1. This criterion
Ailustin and 'CTTalker (1963) and Petersen (1965a) demon- is valid whether diffusional limitations exist in the particle
strated that the Weisz-Prater criterion can fail. Hudgins or not.
(1968), using a perturbation method, derived a design Kubota and Yamanaka (1969), extending the work of
criterion to ensure TJ > 0.95: Akehata et al. (1961), derived a criterion involving pertur-
bations in both temperature and concentration. The criterion,
expressed here in our nomenclature, ensures 9 > 0.95 pro-
vided that:

542 Ind. Eng. Chern. Process Des. Develop., Vol. 10, No. 4, 1971
Diagnostic tests for detecting heat and mass-transfer effects in experimental catalytic reactors are reviewed
and updated. New perturbation criteria are presented for intraparticle, interphase, and interparticle trans-
port in both single- and mixed-phase flow in fixed beds. Emphasis i s placed on the proper choice of reactor
parameters to minimize deviations from isothermal plug-flow performance.

of Equation 9, which also covers diverse kinetics, appears


(7) more generally applicable.
h criterion for isothermal operation of a catalyst particle
in which 3rd R f t Sare. respectively, the derivatives of the is now apparent from inspection of Equation 9:
rate expre-4oii< \vith respect t,o concentration and tempera-
ture, eraliint,ed a t the surface. While applicable to both endo- (13)
thermic :iiid exothermic reactions, the criterion becomes Petersen (1965b), by means of an asymptotic solution, derived
imagiiiai.y for rare exothermic cases in which t'he second a similar criterion with a value of 0.3 on the right for first-
term iii the brackets becomes larger than the first'. order reactions.
The criterion can he extended to these cases for which For most cases encountered in practice, Akehata et al.
7 > 1 by squariiig both sides. For power-law kinetics (except (1961) note that the term in Equation 7 accounting for the
zero order). 7 = I i 0.05 if: intraparticle temperature effect amounts t o less than 10%
of the term for diffusion. Consequently the effect of tempera-
ture distribution within a catalyst particle is usually much
less important than the internal diffusion problem. Transport
or i n t e r m - of tlic o1)servcd rate and dimensionlef;sparameters: limitations occurring outside the catalyst particle must
nom be considered.
(9)
Interphase Transport

jii which: For vapor-phase systems, the greater part of the resistance
to heat transfer is often in the boundary layer or "film"
E around the particle, rather than within (Carberry, 1966).
Y =
RT, This happens because the effective thermal conductivity
of the solid is usually much larger than the conductivity
exprrsscs tlic scn-itivity of the reaction rate to temperat~ure
i n the gas phase. I n the case of a highly exothermic reaction,
and :
the particle temperature can become considerably greater
than that of the bulk stream.
A1 P
A criterion for detecting the onset of a heat transport
limitation in the film has been derived b y the author (Mears,
represents the maximum temperature difference that can 1971b) with the perturbation approach. The heat transfer
exist iii the particle relatire to the surface temperature, resistance of the film i s assumed to be lumped a t the surface.
( T - T,?)!T,. Typical values are 10-40 for the former, If the obqerved rate is t o deviate by less than 5o/c, the cri-
and -0.1-0.1 for the latter parameter (McGreavy and terion requires:
C r e s w d l , 1969). The last criterion can 'also readily be
derived by the orthogonal collocat,ion procedure of Stewart
and Villadl;en (1969).
T h e special case for r h i c h 112 - ?PI is close t'o or equal to
zero requires further consideration. When this occurs, the where Ta is the temperature of the bulk fluid, and h is the
heat effect compensates for t,he diffusion effect so that the heat transfer coefficient given b y correlations (Gliddon and
rate stay-: nearly constant until the concentration within Cranfield, 1970; Littmaii e t al., 1969). This dimensionless
the catalyst has dropl)ed t o lehs than 80y0 of the surface criterion is valid whet,her transport limitations exist, in the
value. Such deviations exceed t,he raiige of applicabilit,y particle or not. It shows t8hat interphase heat transport
of ihr perturbatioii approach, so this peculiar case is better limitations can be expected when reaction heats and rates
of Petersen (1965b). are high, aiid flow rates low (small h).
described by the asyniptotic anal
The latter method give5 a value of 13 for the right side of Note the similarity in form to Equation 6, with h replacing
Equation 9 when 71 aiid rp both equal 1.O. A / r P . Comparison reveals t h a t the interphase heat' t'ransfer
It i-: interesting to compare the criterion of Equation 9 resistance becomes limiting before the intraparticle resistance
n-ith the more widely known one of Weiqz and Hicks (1962) as long as:
for exothermic firqt-order reactioiis: Bip = hd,/X < 10
i n which the Biot number expresses the ratio of thermal
resistance of the particle to that of the film. Condition 15
is usually met in laboratory reactors because of the low
This relationshii) ivas derived by analogy t,o the isothermal flow rates.
case, with a conservatire simplification of the differential For isothermal cases with first-order kinetics, Carberry
equation. The criterion safely predicts positive deviations (7 (1961) showed that the effect of bulk mass transport across
> 1.05) for esotlierrnic reaction.;, but fails to predict nega- the film can be aswmed negligible relative to surface-reaction
tive deviations for endothermic reactions. Thus t,he criterion kinetics provided t h a t :

Ind. Eng. Chem. Process Des. Develop., Vol. 10, No. 4, 1971 543
qk/k,a < 0 .I (16) This equation is applicable to both endothermic and exo-
thermic reactions, predicting the occurrence of positive devia-
in which k , is the gas-particle mass transfer coefficient, and
tions for the latter.
a is the external surface-to-volume ratio of the particle.
This expression is now readily generalized to other reaction
Interparticle Transport
orders by the perturbation approach, allowing a 5oj, deviation
as before. The criterion can be expressed in terms of another Trausport effects, occurring both radially and axially,
Damkohler number: within the reactor as a whole are particularly difficult to
evaluate and control. If neglected, radial temperature
cd = &,/CbkC < 0.15/n (17) gradients can lead to reaction rates several thousandfold
where c b is the bulk fluid concentration. Comparison with greater a t the axis than at the wall (Denbigh, 1965). Less
Equation 14 for typical cases shows that heat traiisfer frequently, bypassing or axial dispersion of mass may seriously
across the film causes excessive deviations long before mass distort the reactor perforrrance.
transfer becomes limiting. The criterion also shows that One commoii approach to minimixiiig these effects is to
zero-order reactions are not affected by iuterphase transport. adopt the differeritial reactor, in which low coiiversioiis are
One case in which severe interphase mass transport liniita- taken. But high precision of chemical analysis is required,
tions might be expected is the trickle-flow reactor, i n which a and interphase gradients are not always eliminated. T h e
liquid reactant trickles down the catalyst pellets while a recycle reactor provides a means of operating differentially
reacting gas flows either in the same or opposite direction. while maintaining a finite overall conversion. Flow rates
This type of reactor is extensively used in the petroleum past the catalyst can then be independently varied to elim-
i!idustry for reactions such as hydrotreating and hydro- inate interphase and interpart,icle gradients. The well-stirred
cracking of heavy oils. heterogeneous catalytic reactor, in which pellets are suspended
Satterfield et al. (1969) examined the transfer of hydrogen in wire cages attached to a rotating shaft, offers another
across the liquid YiIm for a first-order hydrogenation reac- alternative. The merits and demerits of these reactors have
tion. I n close parallel to the last criterion, they concluded been ably discussed in recent reviews (Carberry, 1969;
that mass transfer from the gas phase to the outer surface Smith, 1968).
of the pellets will not be a significant resistance provided: Integral reactors are sonietimes necessary, however, for
bot,h practical and theoretical reasons. For comples reaction
Girp/C*kLs < 0 . 1 5 (18) networks, it may prove difficult to prepare synthetic feeds
where kLs, the overall liquid side mass transfer coefficient which match the composition of the actual fluid a t various
can be estimated by procedures outlined in their paper, stages of conversion. High conversion data are usually prefer-
and C* is the equilibrium concentration of hydrogen in the able in statistical discrimination between rate espressions of
bulk liquid. the Larigmuir-Ilinshel~~oodtype (Dariiard and ?\Iitchell,
An important conclusion was tjhat mass transport through 1968; Alezaki and Happel, 1969). 13arnard and Xitchell
the liquid film cannot b e a significant limitation unless the (1968) cont,end that more accurate results can be obt,ained
catalyst particles themselves are operating a t a very low by measuring integral data over the whole of the conversion
effectiveness factor. T h i s results because the effective diffu- vs. reciprocal flow rate graph than in an analysis using only
sivity within the catalyst does not usually exceed a value of a part of the graph. Moreover, integral data are more selective
about of that in the boundary layer or film adjacent because a prospective rate equation is required to satisfy
to the pellet, aiid the film thickness is always substantially two conditions: effectively, rate vs. contact time and rate
less than the catalyst radius. T h e same conclusion applies, vs. initial concentration. Differeutial data have only to fit
of course, to the vapor-phase case discussed previously. the observed variation of rate with initial concentration.
It is now obvious that a criterion can be derived for the But to obtain reliable data from integral reactors, the investi-
combined intraparticle-interphase case by the perturbation gator must take extra precautions to ensure that heat and
approach. I n effect, one merely replaces Dirichlet boundary mass transfer effects do not interfere.
conditions: I n the radial direction, this means enhancing both heat
aiid mass transport. Dilution of the catalyst with inert
T = T,, C = Cs a t r = rp (19) solids is often useful because it effectively decreases the ratio
of catalyst volume to wall area for heat transfer. Caldwell
with mixed boundary conditions: and Calderbank (1969) described strategies for optimizing
reactor performance by varying the dilution ratio with asial
x distance. I n experimental reactors, our objective is to elim-
inate the usual rise of the centerline temperature to a hot
spot,. Their analysis, based on a one-dimensional model,
shows that a zero temperature gradient can be imposed
along the reactor length by using a dilution ratio which
Excluding n - Y b P b close to zero, the resulting criterion for decreases linearly with conversion. Since this distributes the
q = 1 f 0.05 becomes: reaction rate uniformly down the bed, the dilution ratio
also decreases linearly with reactor length. This approach
also applies fairly well to practical cases of ninny parallel
reactions provided that the kinetics can be lumped (Hutchiii-
where y b and are both evaluated a t bulk fluid, rather than son and Luss, 1970). I n a n actual three-diniensioiial reactor,
surface conditions. The analogous criterion to Equation 13 radial variai,ions will still occur.
for isothermal operation is then: An approximate criterion for evaluating the importance
of radial heat transport limitations has been derived by the
author (Mears, 1971b). It applies a t ally cross section in reac-

544 Ind. Eng. Chem. Process Des. Develop., Vol. 10, No. 4, 1971
tors with linearly decreasing dilution, and at the "hot spot" Dilution t o maintain isothermal operation is not without
in undiluted or uniformly diluted beds. At these locations, the hazard, because bypassing effects can influence the conversion
axial temperature gradient equals zero so the sensible heat at large dilut'ioii ratios. For cases with uniform dilution,
term can be eliminated from the differential energy balance. van den Bleek e t al. (1969) developed a stochastic model
-4xial conduction can also be neglected for beds with length- quaiit'itatively describing t'his influence and obtained a
to-particle diameter ratios sufficiently great that plug flow is criterion to determine the allowable degree of dilution.
approached (generally Lld, > 30). Sumerical solutioiis (Car- Their criterion eiisiires that the bypassing effect will be an
berry aiid White, 1969; Valstar et al., 1969) of the conser- order-of-magnitude less than 6 , the relative experimental
vation equations suggest that radial mass transport effects are error (percent) iu the coriversion. Rearranged and expressed
usually negligible in comparison to heat transport; thus t,he in our nomenclature i t becomes:
concentration can be assumed radially constant. These sim-
plifications permitted a n analytic solution (Chambre and
Grossman, 1955) to the radial temperature distribution.
When the heat transfer resistance a t the wall is iiegligible where L is here the bed length including diluent. The criterion
(i.e., RoirD> loo), analysis (?dears, 1971b) for a 5% deviation can be used in calculating the minimum bed length when t,he
yields the criterion: dilution ratio is fixed by requirements of bed isothermality.
I n contrast to the radial case, mass transfer iii the asial
direction must be minimized to prevent significant deviations
from plug flow and lower conversions. Peterseii (1965),
where k , is the effect'ive thermal conductivity of the bed, as applying the asymptotic solution approach to the case of
given in correlations such as Agneiv and Potter (1970) and isothermal, first-order reactions, derived an approximate
Yagi and Kunii (1959) for vapor-phase systems and Week- criterion for freedom from asial dispersion effects:
man aiid Meyers (1965) for trickle-flow, and 6b is the average
reaction rate per unit bed volume a t a given bed depth. I n 01 = dk,D,/o < 1 (28)
integral react,ors the latter can usually be evaluated from
where ka is the rate constant per unit catalyst bulk volume,
the observed rate const,aiit and the coilcentration a t the D, is the asial eddy diffusivity, aiid ij is the superficial fluid
cross section of interest. velocity. In agreement mit'h earlier findings (Wehner aiid
S o t e t,hat the fuiictioiial form of this criterion is similar
Wilhelm, 19561, it shows that the dispersion effect is generally
to Equation 6 with the reactor radius replacing the particle
negligible except for cases involving short beds and high
radius mid the thermal conductivit'y of the bed replacing
conversions. B u t the problem can be much more severe in
t h a t of the pirticle. Since R, >> r p and k , approaches the
trickle flow reactors because of the higher eddy diffusivities.
order of magnitude of X a t low Reynolds numbers, the
The author (Meare, 1971a), utilizing perturbation solutioiis
interpart,icle t,ransport problem is usually much more severe
for poiTer-law kinetics (13urghardt and Zaleski, 19681,
than the intrai)article one. derived a design criterion which will hold t,he deviation
T h e local rate per unit bed volume is related t'o a,the
in the required reactor leiigth to less than 5%. E s p r e s w l i n
average rate per uiiit particle volume, by the expression:
t h e same manner as the last criterion, it requires for the
first-order case:
CY < 0.22 (29)
in which e is thc bed void fraction, and b is the ratio of diluent Thus the new crit'erion is about four and a half times more
to cat'alyst volume. conservative than the earlier one. This results because
Comparison of Equation 23 with the interphase criterion, asymptotic criteria actually indicate the transition region
Equation 14, shows that the interparticle resistance becomes bet'ween the kinetic-controlled xiid transport-iiiflueiiced
limiting first unless: regimes (Peterseii, 1968), rather than the point a t which
transport' effects are just starting to become significant.
To minimize axial dispersion effect,s, the practical expedient
available to the experimenter is to lengthen the bed, which
This condition is achieved only for low values of the radial lowers the axial concentration gradient and, hence, the driving
aspect ratio (ROIrP)or high dilution ratio. Thus the magni- force for dispersion. Consequently the most coiivenient form
tudes of the heat, transport resistances in experimental reactors of the criterion is one giving the minimum reactor length
are generally in the order: interparticle > interphase > intra- for freedom from significant dispersion effect. Expressed as
particle. the rat,io t,o particle diamet,er, the new criterion hecomes:
m h e n the heat transfer resistance a t the wall is significant, I, kbr 20 C,
the interparticle criterion becomes: ->20- = -111-
d, Pe, Pe, Cf
in which r is the space time aiid C, and C/ are the initial
and effluent concentrations. The axial Peclet number baqed
i n which the wall Biot number is a measure of the ratio of the on particle diameter:
thermal resistance of the bed to t h a t of the wall. T h e data
of Yagi and Kunii (1959) yield wall Biot numbers of 0.8 t o 2
for R e < 100 aiid 0.05 < r,/R, < 0.2, so the denominator
becomes of the order of 2. Thus heat transfer resistance at the is a measure of the ratio of convective to dispersive transport.
wall aggravates the interparticle transport problem in small The advantage of the second equality in Equatioii 30 is that
laboratory reactors. it puts the criterion in terms of the observed conversion and

Ind. Eng. Chem. Process Der. Develop., Vol. 10, No. 4, 1971 545
the independently obtainable Peclet number. Given in numbers as low as one. However, the interparticle heat, trans-
terms of reactor length, the new criterion is 20 times more fer coefficient (Yagi and Kunii, 1959) becomes quite weakly
conservative than the asymptotic criterion. dependent on G a t low Reynolds numbers. Thus, as Chambers
For nonfirst-order reactions obeying simple power-law and Boudart noted, this diagnostic test requires very precise
kinetics, the corresponding criterion is : conversion data or an extended range of flow rates.

L 20 n C, Design Criteria for Fixed-Bed Reactors


->-In-
d, Peu Cf I n designing a laboratory reactor, transport limitations
where n is t'he order of the reaction. This general expression can be minimized through proper choice of the parameters.
gives Equation 30 for first-order reactions and zero for zero- Because of the large Arrhenius effect of temperature on the
order reactions, which are not affected by axial dispersion. reaction rate, maintaining isothermal operation is by far the
Kote that the axial dispersion problem becomes more severe most critical considerat'ion. The criteria reviewed provide a
wit8h increasing conversion or reaction order, or with de- means for assessing the best ways to achieire this.
creasing Peclet numbers. Reducing the reactor radius, a n obvious choice, is partic-
Extensive studies have shown that the Peclet number is ularly important since the interparticle criterion depends on it
geiierally close to 2 for gases a t Reynolds numbers greater both explicitly and implicitly. The latter dependence occurs
than 2 and to about 0.5 for liquids when Reynolds numbers because the interparticle heat transfer coefficient becomes
are less t'han 10 (Levenspiel and Bischoff, 1963; Miller and effectively proportional to the mass velocity, G, a t Re > 100,
King, 1966). I n trickie-flow, still smaller values are found and G is inversely proportional to RO2 (at constant mass
in the liquid phase a t low Reynolds numbers (Hochman and flow). The int'erparticle heat transfer criterion then becomes
Effron, 1969; van Swaaij et al., 1969). For example, at a effectively proportional to the fourth power of R, and the
Reynolds number of 4, typical of bench-scale react'or condi- interphase criterion to the 1.2 power a t higher Reynolds
t'ions, these correlations yield Peclet numbers of the order of numbers. Hence decreasing the reactor radius is a vital step
0.1. Coiisequently, the criterion shows that the minimum in minimizing both transport' resistances.
reactor length for freedom from backmixing effect,s may be a n The question then arises as to how far the radial aspect
order-of-magnitude greater in trickle-flow than in vapor- ratio (RO/rP)may be reduced before wall effects become
phase operation. serious. Because the packing is more open near the wall, a
larger fluid velocity and lower catalyst bulk density occur
Empirical Tests for Transport Limitation there. As a result the conversion is lower near the wall,
causing an apparent lower activity for the bed.
While the analytical criteria are useful, estimation of
parameters involved can be difficult a t times. For such cases Kondelik and Boyarinov (1969) calculated the effect of
radial void fraction distribution on reactor efficiency as a
a well-established empirical criterion is to run the reaction 011
function of Ro/rp. Their model divides the reactor into
progressively smaller particle sizes, usually produced by
annular channels based on statist'ical distribution data and
crushing. If the reaction rate per unit particle volume changes
utilizes the Kozeny equation (Carman, 1956) to compute the
with diameter, strong transport influence is indicated.
volumetric flow rates in individual channels. Radial mass
Graphical (Weisz and Prater, 1954) and analyt'ical (Gupta
transfer is accounted for by a lumped radial mass transfer
aiid Douglas, 1967) techniques are available for estimating
coefficient D. I n the Roirp range of 5-10, the calculations
the effectiveness factors in this case. When the reaction rate
show a minimum in the effective bed activity whose depth
remains iiidependent of the particle size over an appreciable
increases with conversion. But the minimum is reduced with
range, the reaction is considered free from intraparticle
improved radial transfer, and the maximum value assumed
gradients on the macroscopic scale.
for D still appears to underestimate the contribution from
An alternate empirical approach, suggested by Koros and
eddy diffusion. Consequently, the gain in isothermality
Nowak (1907), is based on the fact that the rat>eof reaction
through reducing the reactor radius appears to more than
in the kinetic regime is proportional to the number of sites
compensate for the slight channeling loss to R J R , ratios as
per unit volume (S), but proportional to S to the 0.5 power
low as 4. However, wall effects may be more serious in trickle-
in the internal diffusion regime. The test involves pelletizing
flow, so larger ratios should be used for this operation.
mixtures of finely divided catalyst with a n inert powder of
Finally, in an existing reactor of larger diameter, dilution
comparable diffusional characteristics. If the fraction of
of the catalyst bed with inert particles can be utilized to
catalyst in the mixed pellet is f , the ratio of rates must also
reduce heat generation per unit volume. Linearly decreasing
be equal to f in the kinetic regime. This method avoids
dilution ratios are st,rongly recommended for integral reactors.
distortions in the flow field which changing pellet' sizes may
But in diluting a bed of fixed length, it should be remembered
cause.
that the mass velocity is reduced proportionately a t constant
A common empirical test for the possible iiifluence of
space velocity. The interparticle criterion shows that dilution
ext,ernal heat aiid mass transfer limitations consists of
will be advantageous in minimizing radial gradients only if
checking the effect of flow rate on conversion a t consbarit
the reactor is operating a t Reynolds numbers sufficiently low
space velocity. If no such effect is found, it is concluded t'hat
that the effective thermal conductivity is relatively insensitive
transfer of heat and mass to the particles does not influence
to the mass velocity (Le.) Re < 100).
the reaction rate. Chambers and Boudart (1968) warned
that this diagnostic test may fail because of its lack of sensi-
Conclusion
tivit,y a t Reynolds numbers of the order of 10. This observa-
tion \\'as based on a reanal of the interphase heat transfer The elimination of transport influeiices in studies of cata-
data of D e Xcetis and Thodos (1960). More recent data lytic kinetics is a complex problem, The criteria presented
(Gliddon and Cranfield, 1970; Littman et al., 1969; Petrovic provide tools for evaluating and minimizing the various
and Thodos, 1968) iiidicate that both h and k , remain propor- transport resiatances. Like any diagnostic test, each must be
tional to between the 0.95 and 0.64 powers of G to Reynolds applied with due caution and understanding. If carefully used,

546 Ind. Eng. Chem. Process Des. Develop., Vol. 10, No. 4 , 1971
the criteria can help the experimenter to achieve a more X = thermal conductivity of partick, g-cal/sec. cm . C
rigorously defined experimental system . p = viscosity, g/sec cm
Y = stoichiometric coefficient
Nomenclature p = density,g/cma
7 = space time = inverse space velocity, see
a = superficial (outside) surface area of catalyst particle x = Damkohler number for interphase heat transport, =
per unit particle volume, cm2/cm3 (- A H ) &r,/hTb
b dilution ratio, cm3inert solid/cm3 catalyst (cr = Damkohler number for interphase mass transport, =
Bi, s thermal Biot number for catalyst particle, = hd,/h @r,/k,Cb
Bi, = thermal Biot number a t the reactor wall, = hwdt/ke
C* = equilibrium concentration of hydrogen in liquid, literature Cited
g-mol/cm3
C h = concentration of reactant in bulk fluid, g-mol/cma Agnew, J. B., Potter, 0. E., Trans. Inst. Chem. Eng. (London)
48. T1.i 11970).
cf = effluent concentration of reactant, g-mol/cm3 Akehata, T., Namkoong, S., Kubot,a, H., Shindo, XI., Can. J .
CO =-c initial concentration of reactant, g-mol/cm3 Chem. Eng., 39, 127 (1961).
C , = concentration of reactant at outer surface of particle, Anderson, J. B., Chem. Eng. Sci., 18, 147 (1963).
g-mol/cm3 Austin, L. G., Walker, P. L., Jr., AZChE J., 9, 303 (1963).
Barnard, J. A., Mitchell, D., J . Catal., 12, 376 (1968).
d, = diameter of catalyst particle, cm Burghardt, A , , Zaleski, T., Chem. Eng. Sci., 23, 573 (1968).
D, = axial eddy diffusivity, cmZ/sec Caldwell, A . D., Calderbank, P. H., Brit. Chcm. Eng., 14, 1199
De = effective diffusivity in a porous catalyst, cmZ/sec ( 1969).
E = intrinsic activation energy for chemical reaction, Carberry, J . J., AZChE J . , 7,351 (1961).
g-cal/g-mol Carberry, J. J., Catal. Rev., 3 (l),61 (1969).
Carberry, J. J., Ind. Eng. Chem. 58 (101, 40 (1966).
f = fraction of active catalyst in mixed pellet Carberry, J. J., White, D., ibid., 61, (7), 27 (1969).
G = mass velocity, g/sec.cm2 of total or superficial bed Carman, P. C., Flow of Gases Through Porous Media, Aca-
cross section demic Press, New York, 1956, p 12.
h = heat transfer coefficient between gas and particle, Chambers, R. P., Boudart, hI., J . Catal., 6, 141 (1968).
Chambre, P . L., Grossman, L. M., Appl. Sei. Res., A5, 245
g-caljsrc cm*. C ( 1935),
h, = heat t i , > i i i f i f w coefficient at the reactor wall, g-cal/ Uamkiihler, G., Z. Elcktrochenz., 42, 846 (1936).
sec. cni2 (
I
De Acetis, J., Thodos, G., Ind. Eng. Chem., 52, 1003 (1060).
AH = heat of cLheinica1 reaction, g-cal/g-mol Denbigh, K. G., Chemical Reactor Theory, Cambridge Univ.
Press, Cambridge, England, 1963, p. 44.
IC = intrinsic rate constant per unit particle volume, sec-1 Gliddon, B. J., Cranfield, R. It., Brit. Chem. Eny., 190, 481 (1970).
for first-order reaction Gupta, V. P., Douglas, W., Can. J . Chem. Eng., 45, 117 (1967).
kb = apparent rate c~onstantper unit bulk catalyst volume, Hochman, J. M.,Effron, E., Znd. Eng. Cheni., Funrlam., 8, 63
= k ( l - ~ ) qsec-l , ( 1969).
Hudgins, R. It., Chem. Eng. Sci., 23,93 (1968).
k c = mass transfer coefficient between gas and particle, Hutchinson, P., Luss, D., Chern. Eng. J . , 1, 129 (1970).
cm/sec Kondelik, P., Boyarinov, A. I., Collect. Czech. Chem. Coniniun.,
= effective thermal conductivity across reactor bed, 34,3852 (1969).
g-cal/sec. cm. C Koros, R. AI., Sowak, E. J., Chrni. Eng. Sci.,22,470 (1967).
Kubota, H., Yamanaka, Y., J . Chem. Eng. (Japan),2,238 (1969).
Iz LS = overall mass transfer coefficient through liquid film, Kubota, H., Yamanaka, Y., Dalla Lana, I. G., ibid., 2 , 71 (1969).
cm/sec Levenspiel, O., Bischoff, K., Advan. Chem. Eng., 4, 93 (1963).
L = reactor bed length, cm Littman, H., Barile, R., Pulsifer, A,, Znd. Eng. Chem., Fitndani.
L o = reactor bed length rvith undiluted catalyst, cm 7, 554 (1969).
AIcGreavy, C., Cresswell, D. L., Can. J . Chem. Eng., 47, 583
n = integer exponent in power law rate expression (1969).
Pea = axial Peclet number, = d,fi/D, Mears, D . E., Chem. Eng. Sci., in press (1971a).
T P = particle radius, cm
hlears, D. E., J . Catal., 20, 127 (1971b).
R = gas constant, g-cal/g-mol. O K R., Happel, J., Catal. Rev., 3, 241 11969).
R e = Reynolds number, = cpd,/p
., King, C. J., AIChE J . , 12, 767 (1966).
, Wakao, N., Smith, J. AI., ibid., 10, 130 (1964).
R O = radius of tubular reactor, cm , E. E., Chem. Eng. Sci., 20,587 (1965a).
R = observed reaction rate per unit particle volume, , E. E., ibid., 23, 94 (1968).
g-niol/sec. em3 Petersen, E. E., Chemical Iteaction Analysis, Prentice-Hall,
Englewood Cliffs, N. J., 1965b, pp. 76, 79, 198.
= reaction rate per unit bed volume, g-mol/sec. cm3 Petrovic, L. J., Thodos, G., Ind. Eng. Chcni., Fundam., 7, 274
= derivative of reaction rate expression with respect t,o (1968).
concentration, see- Satterfield, C. N., Sherwood, T. K., The Role of Diffusion in
= derivative of reaction rate expression with respect to Catalysis, Addison-Wesley, Reading, Mass., 1963.
Satterfield, C. N,, hIa. Y. 13.. Sherwood. T. K.. Insf. Chcni.
temperature, g-mol/sec. cm3.K Eng. (London),Symp.Scr., 28; 22 (1968).
s = active sites per unit particle volume, Satterfield, C. N., Pelossof, A. A , , Sherwood, T. K., A I C h E J.,
S.V. = space velocity, = o/L,, sec-1 15.227 (1969).
t = time, sec Smith, J. S?.,C/km.Eng. Progr., 64,78 (1968).
Stewart, 1%. E., Villadsen, J. V., AZChE J . , 15, 28 (1969).
T = absolute temperature, K Talstar, J. AI., Bik, J. D., van den Berg, P. J., Trans. Inst. Cheni.
D = superficial velocity, cm/sec Eng. (London),47, CE 136 (1969).
Y , = mole fraction of reactant a t catalyst surface van den Bleek, C. AI., van der Wiele, K., van den Berg, P. J..
z = distance in axial direction, cm Chem. Eno. Sci.. 24. 681 11969).
van Swaaij,W., Charpentier, J: C., and Villermaux, J., Cheni.
E m . R c i . 24. 108.1 ilRfi9).
GRI:EK LETTERS
cy = dimensionless axial dispersion number, = d h X / o
P = dimensionless maximum temperature rise, = ( - A H )
D,C,/XT, &., 17,265 (1962).
7 = dimensionless activation energy, = E / R T , Weisz, P. B.; Prate;,, C. b.,Advan. eatal., 6, 143 (1954).
Kunii, O., AIChE J., 6,97 (1939).
Yagi, S.,
6 = relative experimental error in conversion
E = bed void fraction
7 = effectiveness factor RECEIVED
for review November 2, 1970
e = volume change modulus = (1 - v ) Y 8 ACCEPTED MARCH22, 1971

Ind. Eng. Chem. Process Des. Develop., Vol. 10, No. 4, 1971 547

Das könnte Ihnen auch gefallen