Charaterization of Liquid Solid Reaction

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 8

Jakob, M., Phys. Z. 22, 65 (1921). Price, D., Ind. Eng. Chem., Chem. Eng. Data Ser. 1, 83 (1956).

Joule, J. P., Thomson, W., Phil. Mag. 4, 481 (1852). Redlich, O., Kwong, J. N. S., Chem. Revs. 44, 233 (1949).
Koeppe, W., Kaeltetechnik 14, 399 (1962). Reid, R. C., Sherwood, T. K., “Properties of Gases and Liquids,”
Koeppe, W., Proceedings of 10th International Congress on 2nd ed., pp. 50-2, McGraw-Hill, New York, 1966a.
Refrigeration (Copenhagen, 1959), Vol. 1, pp. 156-63, Per- Reid, R. C., Sherwood, T. K., “Properties of Gases and Liquids,”
gamon Press, New York, 1960. 2nd ed., Appendix A, pp. 571-84, McGraw-Hill, New York,
Kordbachen, R., Tien, C., Can. J. Chem. Eng. 37, 162 (1959). 1966b.
Lydersen, A. L., Greenkorn, R. A., Hougen, O. A., “Generalized Roebuck, J. R., Murrell, T. A., Miller, E. E., J. Amer. Chem. Soc.
Thermodynamic Properties of Pure Fluids,” University of 64,400 (1942).
Wisconsin, College of Engineering, Rept. 4 (October 1955). Roebuck, J. R., Osterberg, H., J. Amer. Chem. Soc. 60, 341 (1938).
Martin, J. J., Chem. Eng. Progr. Symp. Ser. 59, No. 44, 120 (1963). Roebuck, J. R., Osterberg, H., J. Chem. Phys. 8, 627 (1940).
Martin, J. J., Ind. Eng. Chem.. 59 (12), 34 (1967). Roebuck, J. R., Osterberg, H., Phys. Rev. 46, 785 (1934).
Onnes, . K., Commun. Phys. Lab. Univ. Leiden, No. 23 (1896). Roebuck, J. R., Osterberg, H., Phys. Rev. 48, 450 (1935).
Onnes, . K., Keesom, W. H., “Die Zustandgleichungen,” in Shah, K. K., Thodos, G., Ind. Eng. Chem. 57 (3), 30 (1965).
Encyclopedia Math. Wiss. V-10, p. 842, B. G. Teubner, Leip- van der Waals, J. D., Proc. Sect. Sci. Kon. Med. Akad. Wctcnschap.
zig, 1912. (Amsterdam), 2, 379 (1900).
Partington, J. R., “Advanced Treatise on Physical Chemistry,” Witkowski, A. W., Bull. Acad. Sci. Cracovie (July 1898).
Vol. 1, p. 622, Longmans, Green, London, 1949. Yen, L. C., Alexander, R. E., A.I.Ch.E. J. 11, 334 (1965).
Pitzer, K., Brewer, L., revisers, “Lewis and Randall's Thermo- Received for review September 29, 1969
dynamics,” Appendix 1, p. 611, McGraw-Hill, New York, Accepted July 13, 1970
1961.
Porter, A. W., Phil. Mag. (6) 11, 554 (1906). W’ork carried out under the auspices of the L'. S. Atomic Energy
Porter, A. W., Phil. Mag. (6) 19, 888 (1910). Commission.

Characterization of Liquid-Solid Reactions

Hydrochloric Acid-Calcium Carbonate Reaction

Bert B. Williams,1 John L. Gidley,2 James A. Guin,3 and Robert S. Schechter4


Esso Production Research Co., Houston, Tex. 77001

Techniques commonly used to determine reaction rates for heterogeneous liquid-solid


reactions such as the hydrochloric acid-calcium carbonate reaction are mass transport-
limited and do not reflect surface kinetics. A procedure where liquid flows through a
channel composed of solid reactant is proposed for obtaining kinetic data. An exact solu-
tion for reaction rate with an arbitrary kinetic model is obtained by numerical methods
based upon Duhamel’s theory, for first- and second-order reversible and irreversible reac-
tions. The method is easily extended to other models. An exact boundary layer solution is
given for the first-order irreversible reaction and an approximate boundary layer tech-
nique for solution of the arbitrary reaction rate case is developed. The approximate tech-
nique yields simple solutions which are accurate under certain prescribed conditions.
Existing experimental data for the hydrochloric acid-calcium carbonate reaction are
analyzed in light of the proposed kinetic models.

^Xcidizing of subsurface reservoirs to increase oil or gas pendent upon knowledge of the surface reaction rate of the
production has been practiced since the late 1800’s. In these acid. It is, therefore, of practical importance to be able to
techniques an acid, such as hydrochloric acid, is forced to flow obtain correct expressions for the surface reaction rate under
within the pore structure of the rock matrix, or along a hy- a variety of conditions of temperature, pressure, and acid
draulically induced fracture, reacting with the rock and alter- composition. The reaction rate is relatively fast and only care-
ing reservoir characteristics. Schechter and Gidley (1969) have fully performed experiments, analyzed to account for diffu-
developed a method of computing the rate of acid spending sion correctly, are of general applicability.
and the effect of acid reaction on formation flow capacity for Reaction of an acid with calcium carbonate can be gener-
the case of matrix acidizing. These computations, as well as alized as
calculations of reaction rate in flow along a fracture, are de- hi
2H+ + CaC03 <=í C02 + Ca2+ + H20
1
To whom correspondence should be sent.
2
Present address, Humble Oil & Refining Co., Houston, Tex.
77001
Present address, Auburn University, Auburn, Ala. 36830
3

Present address, Department of Chemical Engineering,


4
2A + B<=*C + D + E
University of Texas, Austin, Tex. 78712 k2

Ind, Eng. Chem. Fundam., Vol. 9, No. 4, 1 970 589


Therefore, the general rate expression for reaction of slightly
Table I. Effective Mass Transfer Ratio for 1 to 1 dissociated acids will have the form
Ratio of Acid Volume to Rock Area
Time, Min f (D//2)°.5
-r =
ki{Cs?Kd)a -

hCcoaCc,.^a (8)
10 0.25 0.009 Historically, acid reaction rate data have been determined
20 0.40 0.010 in a static reaction test, in which a cube of rock of known sur-
30 0.50 0.010 face area is placed in a basket at the top of a reaction vessel
40 0.58 0.010 and a known volume of acid is placed in the bottom of the
50 0.63 0.010
container. Test pressure is usually applied to the chamber by
60 0.71 0.010
70 0.76 0.010 injecting nitrogen so as to stimulate reservoir conditions, and
a test is started by inverting the equipment, thereby bringing
. 0.010
rock and acid in contact. After a set time, the cell is righted
and an analysis made to determine the quantity of rock dis-
solved. Data from this test are normally reported as the frac-
For this reaction, tion of acid spent, /, as a function of contact time for various
a general rate expression is
ratios of acid volume to rock surface area.
—r =
k1aAaacyaDs -

kzaAa'acy'aDs' (1) Some insight into what is occurring during static tests can
be obtained as follows. First, assume a model with two parallel
where o< is the activity of the fth species. To satisfy equi-
walls separated by a distance l with no fluid circulation, a
librium requirements, the ratio of the rate coefficients must be
reaction rate at the solid surface very fast relative to the rate
equal to the equilibrium constant to a power n. This means
that the exponents of the activities in Equation 1 are related
of acid transfer to the surface (C 0 at y l), no alteration
= =

in diffusivity of hydrogen ion with increase in product concen-


by Equation 2.
tration, and at y 0, an inert surface where the concentration
=

(^~ gradient is zero [dC(y 0)/dy =


0]. For this system solution
=

=
(7' -

7) =
(S' ~

5) =
n (2) of the mass-balance equation can be simplified for typical
values of D, l, and t to the form
Highly dissociated acids such as hydrochloric acid are
most commonly used in stimulation techniques. If we assume f = 1 -
C =
(IDí/tt/2)1''2 (9)
that the activity coefficient for all reactants at the surface is where C is the mean concentration in the cell at time t divided
1, and the reaction is first-order in hydrogen ion activity
by the initial concentration. / is the fraction of the acid that
(a =
1), the rate expression is has reacted. Many results obtained in the static cell can be
—r =
kxCA -

hCc^Cy-'2 (3) explained with this equation in terms of mass transfer proper-
ties independent of rate of surface reaction.
Experimental data show that even at high C02 partial pres- Values for the group (Z)/Z2)0,5 computed for the data of
sure, reaction with strong acids goes essentially to completion. Hendrickson et al. (1960) for a 1 to 1 acid volume to rock
Therefore, the backward rate will be negligible in most appli- surface area ratio are given in Table I.
cations and the rate expression can be written The average spacing from the rock sample to a wall in the
test cell is not known, but would appear to be about 2 cm.
—r =
kiCA (4)
Assuming this is representative, the effective mass transfer
By a similar argument, the rate expression assuming the coefficient for the test is 4 X 10-4 cm2 per second, or about 20
reaction is of order N can be written times the molecular diffusion coefficient-. Therefore, this reac-
tion is entirely controlled by mass transfer to the rock surface.
—r =
ZciCa-v (5) In these tests, the apparent mass transfer rate is much more
rapid than is possible by diffusion alone, indicating that con-
Acids that only slightly dissociated in aqueous solution
are
vection occurring in the aqueous phase controls reaction rate.
are also used to acidize formations: formic, acetic, propionic, It appears that this test is sensitive to any change that
and other organic acids. Study of these acids is complicated alters fluid circulation properties and is completely insensitive
by the dissociation process. A general acid (HP) dissociates to reaction rate at the rock surface. Therefore, changes in
according to the relation over-all reaction rate attributed to changes in kinetic rate by
HP^H++ P- Chamberlain and Boyer (1939), Hendrickson et al. (1960), and
van Poollen (1968) are apparently caused by changes in rate
wit-h equilibrium described by an equilibrium constant, Kd, of fluid convection.
defined as Since these tests are mass transfer-limited, data will indi-
cate a first-order reaction regardless of the true order of the
v + ~ surface reaction. For these reasons, kinetic data from static
Kd = -

(6)
Ohp tests cannot be used with confidence to predict rate of acid
spending in field operations. In this paper a procedure for
Again assuming first-order kinetics in hydrogen ion activity
and unit activity coefficients, and introducing acid dissocia- determining accurate kinetic data is proposed and a theory
developed for expected kinetic rate expressions.
tion, gives the rate expression
—r =
kiCs* —

A)2Cco1/2Cca2+1/2 Dynamic Acid Reaction Tests

Or, in terms of the undissociated acid concentration (approxi- General Consideration. Experiments cited to demon-
mately equal to total acid concentration), strate the first-order dependence of acid reaction rate on
acid concentration are diffusion-controlled and therefore,
—r =
fci(CHpXt¡)1/2 —

Zc2Cco1,2Cca2+1/2 (7) are not reliable indicators of the true nature of the surface

590 Ind. Eng. Chem. Fundam., Vol. 9, No. 4, 1970


reaction. Flow tests in which acid in fully developed
laminar flow comes in contact with a reactive surface are X CENTERLINE
Y V = AVERAGE
well suited for experimental determination of reaction rate t VELOCITY X/VELOCITY =

coefficients—for example, laminar flows through a tube or *x 1


h
between parallel walls of calcium carbonate can be used to 1

determine reaction rate, if the experiment is devised so


that the steady-state acid effluent concentration can be Figure 1. Geometry for parallel plate reaction system
obtained before there is a significant change in the dimen-
sions of the system owing to solution of the carbonate
walls. An additional assumption is that diffusivity for each To solve the unit flux problem, the following change of
ion is constant for all ionic concentrations and independent variables is convenient:
of concentration of other ions. Experiments satisfying the
above conditions have been performed by Hoelscher and
Cowhead (1965) in studying kinetics of interfacial reac-
gt(e{,v) =
Ui(u,v) —

Hr Jo Jo
f () \ (14)

tions. where
Consider first reaction in flow between parallel walls of a
2
reactive solid, as shown in Figure 1. Since the correct reac-
Qr 3
tion rate expression at the rock surface is not known a
priori, a general solution for this problem follows for the for laminar flow between parallel plates. The gt{( (, ) must then
surface chemical reaction symbolized as < 0, where =

satisfy the differential equation


di are stoichiometric coefficients (positive for reactants
and negative for products). A mass balance for the fth Ogi Wgt v(y)
v(v) (15)
component gives 0(t 2 Qr
* and the boundary conditions
VW =
(10)
de, 2
=
p p (& ;
1 OgÁufi) OgÁtu i)
Qi{0,v)
This equation must be solved subject to the boundary condi- ~q/J0 J0
tions
The solution of this problem is
\ n OCiiti, 1) pq(tt) J act(e¡,0)
and---=0 n
——; (11)
^Hr
= =
Ci(0,77) Cid,·
---

0v Di gÁu,v) =
Bsfs(v)ex (- ) + (16)
5=0
Solutions for mass transport problems in flowing fluids are where the /„(»?) are eigenfunctions defined by the differential
often expressed in terms of eigenvalue expansions, but near
equation
the inlet (e< 0) such expansions converge slowly and bound-
=

ary-layer solutions are more convenient. This development d2fn / n


+ xMv).fn
.
2 =
0 (17)
makes use of both solutions. To write the solution to the j^F
general problem, the solution to the unit flux problem is
denoted as ^ , ), where ut satisfies Equation 10 together subject to the boundary conditions
with the boundary conditions dfn{ 1) dfn{0)
_ _
Q (18)

Mi( 0, )
,) 1
The Bs are determined by
and

dMXoO)
;r
Hr Jo Jo Jo_
(f \/ ( ) ( )
= B, (19)

Then by Duhamel’s theorem,


The solution to the unit flux problem is then
" pg(x)
=
I
Jo ¿-'ll 0(1
[Ui(«i —

\, )] \ + Cid (12)
Ui((i,v) =
Bsfs(v) exp (-Xs26i) + +
q
Integrating this expression by parts, a form more convenient
for numerical computation is obtained: 7T
Hr Jo Jo
() (20)

For fully developed laminar flow where ( ) 1 v2 and = —

Qf —

2/3, the eigenvalues and eigenfunctions evaluated at the


Ci((,v)
wall are tabulated in Table II.
A similar analysis for a circular tube with a reactive wall
(13) gives for the solution of the unit flux problem

where all axial positions are referred to e = e4. ut{Zi,y) =


¿
$

0
DsFs{y) exp (-XS2Z4) + W +
Vc
Equation 13 provides a means of obtaining the solution for
acid concentration for an arbitrary surface reaction, once the 1 1 f
unit flux solution has been obtained.
7
-

v&md\ (21)
yc Jo a jo

Ind. Eng. Chem. Fundam., Vol. 9, No. 4, 1970 591


Table II. Mathematical Properties of Unit Flux Solution
Parallel Plates Circular Tubes
s » Ml) 8, , F.(l> D,
0 0 1 -39/280 0 1 -7/24
1 4.2872 -1.2697 0.17503 5.0675 -0.49252 0.40348
2 8.3037 1.4022 -0.05173 9.1576 0.39551 -0.17511
3 12.3106 -1.4916 0.02505 13.1972 -0.34587 0.10559
4 16.3145 1.5601 -0.01492 17.2202 0.31405 -0.07328
5 20.3171 -1.6161 0.009969 21.2355 -0.29125 0.05504
6 24.3189 1.6638 -0.007164 25.2465 0.27381 -0.04348
7 28.3203 -1.7054 0.005415 29.2549 -0.25985 0.03560
8 32.3214 1.7425 -0.004248 33.2615 0.24833 -0.02991

where artificial introduction of additional terms as was done by

-f
Solbrig (1967).
yv(y)dy By assuming a value of gw+1), the concentrations at point
ew+1> along the surface can be computed using Equation 23,
and the Fs satisfy the differential equation and in turn, be used to compute a corrected value of g(2f+1).
By continuing this iterative process, new reaction rates and
d_ d,Fs\ concentrations satisfying Equation 23 can be found. This
7 ’-' $$ =
0
dy ~dy J process of calculation can be continued to yield both the
and the boundar)· conditions concentrations and the reaction rates evaluated along the
surface. The amount of acid reacted is given by
dFs( 0) dF,(l) =
0 p rL*
dy dy R*
-gJQ g(')d*
=
(24)
The coefficients Ds are determined from
This quantity is easily shown to be related to the mean concen-
f1 r i
I Fs711(7) I -

I v(£)£didXdy tration divided by the inlet concentration according to the


Jo Jo Jo relationship
D, = -

6,
/‘ FJyv(y)dy R* = 1 -

Ci

where Ci is the average dimensionless concentration at L*.


(25)

Using the method given below, it can be shown that the


boundary-layer solution to the unit flux problem evaluated
along the surface of the wall is First-Order Irreversible Reactions

While it appears that the experiments leading to the ac-


M<(e<,l) =
G-r
'2 (22) cepted conclusion that the rate of acid reaction with carbonate
is first-order are not valid, it is possible that this is, in fact, the
*3 nature of the reaction. First-order irreversible reactions are
also of interest, since the resulting equations are linear and
where («) is the gamma function of the argument e.
can be solved without using Duhamel’s theorem or the unit
In the general case, the dimensionless reaction rate g(e)
flux solutions developed above. The solutions developed by
depends on the local value of the concentrations at the reac- numerical methods based on Duhamel’s theorem were tested
tive surface. Equation 13 can be used in an iterative fashion
against these exact solutions and found to be in agreement
to evaluate both the surface concentration and the reaction
within 1% over the entire range of e. Also, approximate solu-
rate along the surface. Suppose that the rate of reaction has
tions introduced below can be compared to these exact solu-
been established at positions e(0), e(1), ew), giving the . ..

tions.
particular values g(1), g(2), gw). It is desired to establish
The first-order irreversible case is studied by putting q(ei)
. . .
=

g(.v+iy From Equation 13, we can write —

Ci. For larger values of the dimensionless distance ei,


n ££(0) U\
p N
q(‘+D

q(’) (d = e), solution in terms of an infinite series using the


Cío -j-
1

n + methods of separation of variables can be found. However, in


On U H s = 0 e('+i) —
e(*)
the entry region, the rate of diffusion to the reactive surface is
controlled by the development of a diffusion boundary layer,
U -

), dX (23)
) the rate-limiting processes occur very near the wall, and the
assuming that bq/be is a constant between the nodal points. velocity distribution can be replaced by the approximation
In performing the calculations, it must be remembered that
() =
1(1
-

v)
the boundary layer solution, Equation 22, should be used for
small values of e and then switched to Equation 20 for larger For this special case, the expression developed by Acrivos
values of the argument. The solutions as computed from 20 and Chambre (1957) reduces to
and 22 agree to within 2% over the region 10~3 to =

£(=
7 X 10'3 and thus the switching point should be in this dCi(|,l)
interval. This method of calculation does not require the

Ci(x*,l)

592 Ind. Eng. Chem. Fundam,, Vol. 9, No. 4, 1970


where
Table III. Values of Coefficienls in Power Law Solution

[-TM0 €
n

1
2
-0.82699
ß

0.60412
Making the substitution
3 -0.40275
Z3 =
4 0.24980
5 -0.14598
Zi3 =
£
6 0.08110
gives 7 -0.04306
8 0.02143
fz dCtZi.l)
' ’ 9 -0.00724
-C(Z, 1) =
[Z3
-

Z,3]-!/3 dZi (26) 10 0.00047


Jo “Zi
11 -0.00000
The concentration at the wall is obtained in terms of the
power series

Cl = 1 +
=
ß (27)
1

Substituting Equation 27 into Equation 26, it is found that

-1 -

ß =
(n + l)/w„ (28)
n—l =
0

where

-f [Z3
Zl”
-Zv
dZ i (29)

Equation 29 can be integrated to give


Figure 2. Surface reaction rate in circular tube
+ Irreversible reaction
3 /
;z" (30)

<)
Substituting Equation 30 into Equation 28 and equating
coefficients of like powers, it is found that

ft =

and

r
() Figure 3. Surface reaction rate between parallel plates
Irreversible reaction

The first eleven values of ß„ are given in Table III.


Reversible Reactions
With these values known, concentration at the wall is given by
The reaction of weak acids should be considered as revers-
ible. Figures 4 and 5 show dimensionless conversion R* as a
function of length in a parallel plate system for reactions hav-
Ci(e,l) —
1 + ß (31)
= 1 ing the form

Using Equation 24 gives the final result -g(e) =


Ci -

KC21,2C31/2, K = 0.1

-g(e) =
CJ -

KC2C3, K =
1.0

PL* 3 In these figures the dimensionless quantity, P, is defined as


R* 1 + ß (32)
~Q~f =
1 71 3
k\ fclCl
A similar analysis for a circular tube gives Equation 32, again P =

and P =
respectively
üi Di
with R*, Qf, and L* replaced by their analogs for the circular
case. Results predicted by combination of infinite series These calculations have been performed for Dn —0.125 =

solution and the boundary layer solution are shown in Figures and Dzi = —0.375, which are approximately valid for the
2 and 3. system HCl(Ci), CaCl2(C2), and CO2(C3).

Ind. Eng. Chem. Fundam., Vol. 9, No. 4, 1970 593


with the boundary conditions C^ei, oo) =
Ci0 and

&C,(x„0) =
/9ei\1/3 mW
d7< V 2 / Dtl
Equation 34 can be integrated directly to yield, after applying
the boundary conditions,

Ci(et, 0) (35)

This expression represents an approximate solution for the


wall concentration for an arbitrary surface reaction rate.
Figure 4. Surface reaction rate between parallel plates Applying this expression to the unit flux problem in which
Reversible reaction dw,(e,,0)
dy<
and
Mi(e<,°°) =
0

yields

(36)

When compared with Equation 22, Equation 36 is in error


by about 20%. The approximation improves considerably as
the order of the reaction increases with the unit flux problem
being of order zero.
For an nth-order irreversible reaction,

9(e) =
-[Ci(e,0)]B
Reversible reaction
and Ci is given implicitly by Equation 35. In particular, if
n =
1,
The simple iteration scheme previously suggested did not
converge in all cases for reversible reaction and it was some-
times necessary to use a direct search technique to find values
for Ci, C2, and C3 that satisfied Equation 23. The technique
of Powell (1964) worked well in these cases. In this tech-
3

ñique, the quantity 23 [C,p+1 —

C,27]2 is minimized, where


i = l
C^ is the concentration of component i at the pth iteration
as found from Equation 23.

Approximate Solutions
The results of an approximate boundary layer technique
introduced by Solbrig (Solbrig and Gedaspow, 1967) are
compared with the exact results obtained earlier. This ap-
proximate method is useful because the calculations involved
are simpler than those required to solve the integral equations This approximate result is compared with the exact solution
resulting from the exact boundary layer method. in Table IV.
In this procedure t is defined as the dimensionless distance Similarly for n 2, =

measured from the reactive surface—i.e., t .1 (for = —


1/2
small , 1 —

t;2 ·—' 2£). Changing variables in Equation 10 from


(ei,v) to (<·,,.£,) where Zi <e,-1/3 gives
= -3A + 3A 1 + 1
—--—-
4 (t)
L_
3 3AA
3 J
dC,
7 1 2 ^C &!£<
*
,
2 (33) 2pr
(I)
Z te i '
de¿ 3 c)Zt dZ,2
Neglecting the term on the left-hand side of Equation 33 and
For the second-order reversible reaction where q(x) =

defining —
Ci2 + KC2C¡ and assuming that the inlet concentrations of
1/3
2 both substances 2 and 3 vanish from Equation 35, it is found
y t
=
Zi that
Equation 33 can be written Ci(e,l) =

d2C, c)Ct (1 -

2BE) + {(2BE -

l)2 4(1 - -

BE)(B -

BE)}112
C—, + 37 i2 c- =
0 (34)
97< 97, 2(B BE) -

594 Ind. Eng. Chem. Fundam., Vol. 9, No. 4, 1970


Table IV. Comparison of Approximate with Exact Boundary Layer Solutions for First-Order Irreversible Reactions
R*(P =
0.1) R*(P =
1.0)
L* Exact Approx. L* Exact Approx.
10 ~4 1.493 X 10-5 1.492 X 10-5 10-3 1.374 X 10~3 1.352 X 10-3
10-3 1.486 X 10 ~4 1.484 X 10-4 10-2 1.251 X -2 1.215 X "2
10-2 1.471 X 10-3 1.465 X 10-3 10-1 1.047 X 10-1 9.997 X -2
10-1 1.439 X 10-2 1.427 X 10-2

Table V. Apparent Reaction Rate Coefficients from


Data of Barron et al. (1962)
Plate Spacing, Inch P

0.025 1.25
0.05 2.5
0.10 5.0
0.20 10.0

where

B =
K(D2D3yi*
Pl2lSP2lPsi
Figure 6. Experimental and theoretical data for HCI reac-
Similarly, tion between parallel walls

=
Cl -1
P3l(Pl/P3 )1/3
cm2 per second, the rate coefficient for reaction of HCI and
and calcium carbonate is approximately 1.0 X 10“3 cm per
second. This value for the rate coefficient should be con-
c2 =
(D2/p3yiyp31/p2i)C3
sidered as qualitative until additional, carefully controlled

These expressions can be substituted into Equation 24 and experiments have been obtained. To verify the order of the
the dimensionless reaction rate computed. These approximate reaction, these experiments must include variation in acid
expressions have been found to represent a very accurate (to
concentration as well as flow rate and system geometry.
within 2 %) approximation of the solution for values of L* less
than 0.1 for P 0.1. Since they are much more compact and
=
Nomenclature
provide straightforward solutions compared to the use of
Duhamel’s theorem, their use in the inlet region is recom-
a¿ activity coefficient for i'th component
5, coefficients in unit flux solution
mended. C dimensionless average concentration
actual inlet concentration of component i, moles/liter
o concentration of component i at pth iteration as
Dynamic Reaction Rate Tests calculated from Equation 23
Ci dimensionless reactant concentration—i.e., Ca/Ci°
The only available data from which a reaction rate coeffi-
dimensionless initial concentration of reactant—i.e.,
cient for hydrochloric acid can be obtained are those of CS/CS
Barron et al. (1962), who studied the reaction of 15% HCI in Z), molecular diffusivity for ith component, cm2/second
flow between parallel walls of calcium carbonate. The results Pa dimensionless diffusivity ratio, PSi/Piii
of their experiments are compared to predicted reaction rates f fraction of spent acid
h distance from center line of parallel plate system to
for a first-order irreversible model in Figure 6. At low- velocity reactive wall, cm
these data appear to fit the first-order irreversible reaction ki reaction rate coefficient, cm/second
model with the approximate values of dimensionless rate Keq equilibrium coefficient, k2/ki
given in Table V. L total system length in flow experiments, cm
At the higher velocities, observed reaction rates are larger
2 LPi i i LPi
, \ . ....

L* i i i
dimensionless length, (plates), (cylinder)
g_^2
than expected if mass transport to the rock surface is diffusion- surface in
l spacing between reactive and nonreacting
limited. This would indicate that instability in the flcnv may static test, cm
occur even though the Reynolds number is well in the laminar n order of reaction with respect to component 1
range. These instabilities may in part be caused by heat hki(C i°)n_1
P reaction rate coefficient,-^- (parallel plates),
evolution at the surface, surface roughness, or entry effects in Ui
the model. Rh(C i0)"-1
Pi (cylinder)
Assuming that experiments at knv velocity represent the dimensionless reaction rate, r(e)//cl(C'i°)n
g(e)
diffusion-controlled reaction, that first-order kinetics are
Q constant, Qf for plates, Qc for cylinder
applicable, and the diffusion coefficient for HCI is 2.6 X 10~5 r(e) reaction rate at position e, moles/cm2/second

Ind. Eng. Chem. Fundam., Vol. 9, No. 4, 1970 595


B =
radius of tube in calculations for reaction along cylin- 5, =
stoichiometric coefficient
drical channel, cm 2xDi
R* =
dimensionless position along x coordinate, (flat
=
dimensionless reaction rate, Ci
g^r
(cylinder) plate)
{w(plate); y =
dimensionless position along y coordinate, y/h
fíave =
average reaction rate over total reactive surface p =
density, grams/cm3
Re =
Reynolds number, dvp/µ
t =
seconds Literature Cited
=
average velocity, centimeter/second
() =
dimensionless velocity profile, for laminar flow, Acrivos, A., Chambre, P. L., Ind. Eng. Chem. 49, 1025 (1957).
v(y)=
(1
-

42) Barron, A. N., Hendrickson, A. R., Wieland, D. R., Trans. .


x =
distance in axial direction, cm 225, 409 (1962).
y =
distance in transverse direction, cm Chamberlain, L. C., Boyer, R. F., Ind. Eng. Chem. 31, 400 (1939).
Zi —
dimensionless distance along x coordinate xDi/2vR2 Hoelscher, . E., Cowhead, C., Ind. Eng. Chem. Fundam. 4,
150-4 (1965).
(cylinder) Hendrickson, A. R., Rosene, R. B., Wieland, D. R., Division of
Petroleum Chemistry, 137th Meeting ACS, Cleveland, Ohio,
Greek Letters April 1960.
Powell, J. J. D., Computer J. 7, 155 (1964).
a =
apparent reaction order for weak acids Schechter, R. S., Gidley, J. L., A.I.Ch.E. J. 15, 339 (1969).
ß —
coefficients in series approximation to boundary layer Solbrig, C. W., Gidaspow, D., Can. J. Chem. Eng. 45, 35 (1967).
van Poollen, . K., Jargon, J. R., OilGas J. 66, 84 (1968).
problem
7 =
dimensionless radial position, r/R Received for review November 12, 1969
r(n) =
gamma function of variable n, tabulated function Accepted July 15, 1970

Flow of Single-Phase Fluids through Fibrous Beds


Chwan P. Kyan, Darshanlal T. Wasan,1 and Robert C. Kintner
Department of Chemical Engineering, Illinois Institute of Technology, Chicago, III. 60616

A pore model for the flow of a single-phase fluid through a bed of random fibers is proposed. An effective
pore number, Ne, accounts for the influence of dead space on flow; deflection number, N<j, characterizes the
effect of fiber deflection on pressure drop. Experimental data were obtained with glass, nylon, and Dacron
fibers of 8- to 28-micron diameter and with fluids of viscosity ranging from 1 to 22 cp. A generalized fric-
tion factor-Reynolds number equation is presented. The effects of dead space in a fibrous bed on flow
and of fiber deflection on pressure drop have no parallels in a granular bed.

The flow of fluids through porous media has been a subject In this equation, the fact that k depends on fiber orientation
of investigation for many years. A considerable amount of and porosity had been observed and discussed by Sullivan
research has been done on the flow through granular beds and Hertel (Sullivan, 1941; Sullivan and Hertel, 1940)
and many useful results have been obtained (Brownell and based on their experimental work. Thus Equation 1 was in-
Katz, 1947, 1956; Ergun, 1952; Ergun and Orning, 1949). adequate for pressure drop correlations. Various workers
A lesser number of investigations have been done on the using the channel model have elaborated upon Equation 1
phenomena of flow of fluids through fibrous media, mostly with modification for shape and orientation (Davies, 1952;
in connection with aerosol filtration. Fowler and Hertel, 1940; Langmuir, 1942; Sullivan and
General approaches pursued by most workers on the flow Hertel, 1941).
of fluids through fibrous beds involved the development of Most workers (Chen, 1955; Iberall, 1950; Wong et al.,
theoretical pressure drop equations from either a “channel 1956) using the drag model rejected the applicability of the
model” or a “drag model.” The former was the more exten- channel model because of the high porosity of a fibrous bed
sively used. and derived a pressure drop equation by considering the drag
Most workers using the channel model started with the forces due to fluid flow on the fibers.
Kozeny-Carman equation, Wong et al. (1956) employed an effective drag coefficient,
Cdc, to account for the fiber orientation, interference of
=
frUS* ~

(1) neighboring fibers, fiber ends, and nonuniformity of fiber


L e3
distribution in the bed. They concluded that the fiber volume
which in the friction factor form becomes fraction, y, has a marked effect on CDe. The higher the value
of 7, the higher is the neighboring fiber interference which
k
/ =
(2) leads to a higher CDe. They also noticed the leveling off of
kATRe' the effective drag coefficient-Reynolds number plot at
Reynolds numbers greater than 6.
1
To whom correspondence should be sent. Gunn and Aitken (1961) in their study of the mechanism

596 Ind. Eng. Chem. Fundam., Vol. 9, No. 4, 1 970

You might also like