Sunteți pe pagina 1din 7

TWO-PHASE FLOW IN PACKED BEDS

Evaluation of Axial Dispersion and Holdup by Moment A n a l s i s


V. E. SATER1 AND OCTAVE LEVENSPIEL

Illinois Znstiute of Technology, Chicago 76, Ill.

An experimental investigation of the extent of axial dispersion in both gas and liquid phases in a bed packed with /*-inch Raschig rings or Berl saddles is presented. Radioisotopes, argon-41 in the gas phase and iodine-1 31 in the liquid phase, were used as tracers. The concentration at a position along the bed could then be measured from outside the bed without disturbing the flow patterns by using scintillation detectors. By comparing the moments of the responses measured a t two points along the column, the velocity and dispersion coefficient were evaluated for that section of the bed between the two points. This use of the moments of the curves circumvented the need for an exact tracer input (such as a step function) and greatly simplified the experimental problems by allowing an arbitrary sloppy pulse input. Correlations show the effect of flow rates and packing geometry on the axial dispersion numbers of both phases.

M are solved by making simplifying assumptions concerning


the flow behavior of the system. Probably the most familiar example is in heat transfer. I n calculating film coefficients for laminar flow through pipe, the conduction term in the axial direction is neglected on the assumption that it is small compared to the convective transfer term. This greatly simplifies the problem and leads to no great error in predicting film coefficients. I n the same manner, axial dispersion or mixing in the axial direction has been neglected in the calculation of mass transfer coefficients in packed towers. True values of mass transfer coefficients, those based on the actual existing concentrations of materials, can be evaluated only when the extent of this axial dispersion is known. There are two useful models for describing the mixing in a packed bed. The first (the mixing cell model) assumes that the packing can be characterized by a series of completely mixed cells. The mixing in a bed is therefore a function of only one parameter, the number of mixing cells in the bed. The other approach (the dispersion model) assumes that the various factors causing axial mixing can be described by a diffusional-type process superimposed on plug flow. This is reasonable if the length over which a single mixing effect acts is small and if a large number of such events take place in the vessel. In laminar floiv, the factors causing axial mixing are molecular diffusion and the overtaking of fluid elements due to the velocity profile. For turbulent flow, an additional factor, turbulent eddy diffusion, plays a role in causing axial mixing. With the dispersion model then, deviation from plug flow is accounted for by a flux term

OST

chemical engineering problems involving fluid flow

transfer of material in and out of stagnant pockets and local channeling, in addition to the effects of turbulent eddy diffusion. Admittedly, the use of a single term to account for these many effects may be an oversimplification, but it should describe the flow characteristics in a packed bed better than the plug flow model.
Technique of Evaluating Effective Diffusivity

T o describe the flow through a packed bed mathematically, some simplifying assumptions must be made. The first is that the axial velocity is constant over the cross section of the bed. If the ratio of tower diameter to packing size is greater than 8 to 1, this assumption is reasonable. The second assumption is that no concentration gradients exist in the radial direction. If the axial velocity is uniform and any material is introduced over the entire cross-sectional area, thi3 also is a realistic assumption. To evaluate the mixing term, a tracer is injected into the main flow stream and its behavior as it moves through the bed is analyzed. By applying a material balance to a cross section of the bed, the following partial differential equation is obtained :

-D - $I u bX2

bC 2

bc

ax

bC

dt

Putting the equation into a dimensionless form gives :

-($)(>+-=-bC

dz2

bz

bc be

-D-

dc
dx

where D is termed the axial dispersion coefficient and accounts for all the factors causing mixing in the axial direction, such as
1

Present address, Arizona State University, Tempe, k i z .


I&EC FUNDAMENTALS

T o solve this equation, appropriate boundary conditions must be used. I n most cases, the concentration is taken to be zero at time equal to zero. The form of tracer input serves as the second boundary condition-e.g., a change in concentration of tracer entering the bed in the form of a pulse, step, or sine function of time can be used. The third boundary condition specifies conditions at the end of the bed-e.g., for a n infinitely long bed, the concentration must be finite throughout the bed. The solution of Equation 2 will give the concentra-

86

tion at any point in the bed as a function of time with the axial dispersion group, D / u d , as a parameter. Thus by comparing this function with the experimental concentration-time data, the dispersion group can be determined. The technique has been used xvith the three forcing functions mentioned above (3-8, 70, 72-74, 7'9). The method described above is easily handled mathematically but difficult to realize experimentally. To inject a tracer into a flowing stream so that the concentration-time relationship corresponds to a siinple mathematical function such as a sine wave without disturbing the flow conditions is not an easy task. This problem can be circumvented by injecting an arbitrary amount of tracer into the bed and measuring the concentration-time response a t two positions far enough downstream so that the flow patterns are re-established (Figure 1). One of the responses would be equivalent to a boundary condition, even though it could not be represented by an analytical function of time. I t then becomes a problem of relating the shapes of the two response curves to the dispersion occurring in the bed. Aris ( I ) , Bischoff and Levenspiel (Z), and Sater (77) have developed the solution for the physical model illustrated in Figure 1 based on a comparison of the moments of the two response curves. The restrictions on the tracer input are that the concentration must be equal to zero for time less than zero and return to zero after some finite time interval. Thus any nondescript pulse can be used as a n input. As the tracer moves through the packing, the pulse will flatten out as shown in Figure 1 as a result of axial mixing. The variance of each curve is an indication of the spread of the curve and, therefore, the difference in the variances of the two curves is a measure of the mixing occurring between the two measurement positions. Likeivise, the dij'ference in the first moments (or the centers of gravity) of the two curves is directly related to the flo\v velocity in the test section. The mathematical relationships between the moments and the physical parameters of the model are given by the following [see ( 2 ) and (77) for details on the derivations] : (3)

where

pi=--

-_

cit at

cidO
0

''

J,

Pp

(5)

c,dt .

and ct is the ratio of length of packing downstream of the second measurement point to the length of the test section (Figure 1). The functions, f and g, are complex and account for the fact that the packed bed has a finite length and thus end effects must be considered. If CY can be made large enough, the values o f f and g in the above equations can be set equal to zero and Equations 3 and 4 reduced to:
Ap =

() 7

Aa2 = ):2 (

(f)

Using Equations 5 and 7 lvith the experimental data will directly yield a value for the axial velocity of the measured phase, u , and indirectly, the fraction of the bed occupied by that phase, E , since:
u = uo

(9)

Equations 4 and 6 allow the axial dispersion group, D l u d , and the axial dispersion coefficient, D,to be calculated for the test section.
Tracers

INJECT

MEASUREMENT

ENTRANCE

PAC KING

EXIT

T o avoid disturbing flow patterns by inserting probes or drawing off samples, radioactive tracers measurable from outside the column could be used. But because the radiation had to pass through the packing and the tolver wall, it was necessary to use a strong gamma emitter. Also, a water-air system was studied, so the gas-phase tracer had to have a low vapor pressure to avoid any transfer between phases. (This experiment was performed not to measure adsorption rates but to describe the flow characteristics of each phase.) Unfortunately, the only insoluble gaseous gamma emitter is argon-41, which has a half life of only 2 hours and therefore is not commercially available. So quartz ampoules were filled with high purity argon gas, irradiated by the Argonne Sational Laboratory for a 6-hour period using a flux of 6 X 1 O I 2 neutrons/sq. cm.-sec. (the gas had a specific activity of 1.9 millicuries per milliliter a t the time of removal from the reactor), rushed to the Illinois Tech campus, and used immediately. They were sized to contain about 0.5 millicurie of argon-41 each at the time of use. Iodine-131 in the form of a sodium iodide solution was available from Tracerlab, Inc. It has a low vapor pressure and a half life of 8.1 days, making it a good liquid-phase tracer. The iodide solution \vas purchased in glass vials containing about 50 microcuries each at time of use.
VOL. 5

INJECTION

CURVE I
't-

CURVE 2

Figure 1. Schematic of experimental setup with general response curves

NO. 1

FEBRUARY 1966

87

AIR T O V E N T

Apparatus

POULE

INSERTION

DEVICE, S E E F I G , 3

SUPPORT

PLATE

TO

SEWER

Figure 2.

Packed tower

A packed tower was constructed using 4-inch i.d. glass pipe with tees fitted on the end as shown in Figure 2. The feed streams and tracers were introduced through the branch sides of the tees. The pipe was packed to a depth of about 1 2 feet with '/?-inch ceramic Berl saddles or '/?-inch Raschig rings, both obtained from the hl. A . Knight Co. This height of packing probably caused channeling but was necessary in order to ignore end effects and use Equations 7 and 8. The use of a redistributor would have introduced disturbances in the floiv patterns in the test sections. A support plate for the packing was made by perforating a '/*-inch brass plate so that the perforations accounted for 507, of the total plate area. The air leaving the top of the tower \vas passed through a rubber hose to the laboratory's ventilation system and there diluted with more air before being vented from the roof. Because of the radioactivity, the system was made leakproof. A U-bend at the bottom of the tower served as a liquid seal. Air was taken from the building supply lines and the flow rates were measured Tvith rotameters. At low liquid flow rates, the \vater \vas taken from an overhead tank in order to maintain a steady flow. .4t the higher rates, the water was taken directly from the building main. Because of the relatively small cross-sectional area of the toxver, a pipe with a hole drilled in the wall served as a liquid distributor. The device used to insert the gas-phase tracer into the tower is shoivn in Figure 3. The ampoule was placed in the smasher by opening the gate valve and lo\vering the ampoule into position with a pair of tongs. The valve was then closed and the smasher \cas slowly pushed into the tower against the butt plate. After the flow was established, the rod !vas given a hard push and the piston in the smasher shattered the glass or quartz ampoule. The smasher \vas perforated ivith small holes so that the tracer would quickly be flushed from the smasher.
Instrumentation

AMPOULE IN

ER

GLASS

FRAGMENTS

The radioactivity 1% as detected by two Baird-.i\tomic Model 812B scintillation probes positioned and shielded with lead as shoivn in Figure 4. The shielding allowed only the radiation from a thin cross section of the packing to be detected. The probe outputs Xvere fed into research ratemeters which converted the pulses into a voltage proportional to the rate of pulses received. This voltage was recorded with a two-channel Offner Dynagraph recorder and the strip charts \vere used as the response curves.
Results

Figure 3.

Device for loading ampoules into tower

42"

rh

60"

7 1e
SCINTILLATION

LEAD

SHIELDING

Figure 5 represents a typical pair of response curves. The lines drawn by the recorder are not smooth but fluctuate about a general trend because of the random nature of the radiation. The tails of the curves approach the zero level very slo\vly and at this low radiation level the data are not accurate. To facilitate the processing of these curves, a smooth line was drawn by eye to average out the fluctuations. The tails have a heavy weight when the second moment is calculated. Because the values of the concentration are not reliable at large values of time, a cutoff point was established (Figure 5). The contributions to the moments of the curve to

39"

!UTECTOR
Cut- o f f

Po i n t

Figure 4. detectors

Position of scintillation

+t---Figure

5. Reproduction of actual recorder response for

a typical run

aa

I&EC FUNDAMENTALS

the left of the cutoff point were calculated by a numerical technique. Levenspiel and Smith (73) have shown that the tail of the response to a true impulse would decay exponentially. It was assumed that the response to an arbitrary pulse would also decay exponentially and this was confirmed by plotting the data to the right of the maximum of the response curves on semilogarithmic coordinate paper and obtaining a straight line. This line was then extrapolated beyond the cutoff point and used to caiculate the contributions of that portion of the curve to the total moments. The area under each curve and the first and second moments were calculated using this method. Details are given elsewhere (77). Single Phase. Eighteen runs were made with the tower operating under single-phase conditions-Le., with water only flowing through the bed-as a check on the experimental method, since an abundance of data is available in the literature for single-phase sysi.ems. Because the results substantiate previous work, the numerical values are not presented in this paper (77). Two-Phase Flow, Liquid Phase. The fraction of the bed occupied by a phase can be determined from the difference in the first moments of the response curves for that phase. In the

case of a liquid phase, this fraction is known as the total holdup, H2.Holdup data found by this method are compared with literature values in Figures G and 7 . Although holdup appears to vary with the gas rate, a statistical analysis of the data indicates that any such dependence is masked by the experimental error (77). This substantiates the findings of previous workers that the liquid phase is not affected by the gas flow rate at conditions below the loading point. The total holdup is the sum of the operating holdup, Hop, and the static holdup, Hst. The static holdup is the amount of liquid remaining in the bed after the liquid inlet is shut off and the column allowed to drain. It is therefore independent of the liquid flow rate and is a constant depending only on the packing used. Shulman, Ullrich, and Wells (78) reported values of the static holdup as 0.0325 and 0.0317 for l/Z-inch Raschig rings and l/2-inch Berl saddles, respectively. Thus the operating holdup, the holdup dependent on flow rate, can be obtained by difference. Otake and Okada (76) obtained a good correlation of their holdup data along with points obtained by other investigators by using the following equation :

Hop = 1.295 ( R e ) ~ ~ f ~ ~ ~ ( G a ) ~ - ~ * ~ ~(1 0)d ) (a

0 v 0

Present Work Jesser a n d E l g i n ( l l ) S h u l m a n , U l l r i c h , a n d W e l l r (18) E l g i n and W e i s s ( 9 )

Ht

0.04

400

1000
L - (Ib./hr,-sq.ft.)

10,000

60,000

Figure 6. Comparison of liquid holdup data with literature values, '/Z-inch Berl saddles

A
0

Present Work Jesser and Elgin(1l) Shulrnan,UIIrich, and Wello(l8)

A A

0
I
I I

I l l

400

1000

10,000

60,000

L- ( I b./ hr.7

sq. f 1 . )

Figure 7. Comparison of liquid holdup data with literature values, 1/2-inch Raschig rings
VOL. 5

NO, 1 F E B R U A R Y 1 9 6 6

89

1 0 0

ro

IO

4 IO

100

1000

(+IFigure 8. Comparison o operating holdup values with the correlation of f Otake and Okada (16)

-i15% ---where

Correlation of Otake and O k a d a (76) (see Equation 10) range

(Re), = dL/,u, Reynolds number of the liquid phase (Ga), = d3gp2/pc",Gallileo number of the liquid phase a = surface area per unit volume of packed bed d = nominal particle diameter Figure 8 shows the data of the present work according to this correlation. Because the dispersion coefficient, D, depends on the flow patterns in the bed, the dispersion group should be a function of the same factors that determine the holdup, and because of the success of Equation 10 in correlating holdup data, the same form was used to correlate the dispersion group.

T o evaluate exponents B, C, and E, a t least three of the variables, one in each dimensionless number, must be varied. Only two variables, L and a, were deliberately varied in the present work. For the Berl saddles, a was equal to 155 sq. feet per cu. foot and for Raschig rings, a was equal to 114. The viscosity varied between 1.1 and 1.35 centipoises, depending on the temperature of the tap water when the run was made. A least squares analysis of the data gave the following values for the constants along with the 95% confidence limits:

The 95% confidence limit on the exponent is zt0.238 and includes the error incorporated by assuming the viscosity and surface area to be constant. Otake and Kunugita (75) obtained liquid phase axial dispersion coefficients by measuring the response to a step change introduced a t the top of the tower. I n the correlation of their results, they use a Reynolds number containing the interstitial velocity. Thus the total holdup must be known before a Reynolds number can be calculated. I n order to determine the axial dispersion group, two correlations, one for the holdup and one for the dispersion group, must be used. T o compare their results with the present work, their raw data were obtained and plotted on Figure 9. Also included are three points reported by McHenry and Wilhelm (74). Two-Phase Flow, Gas Phase. The amount of liquid in a packed bed will determine the shape and size of the channels through which the gas can pass. The greater the holdup, the smaller these channels will be. Thus the flow pattern of gas is a function of the holdup, which in turn is a function of the Reynolds and Gallileo numbers of the liquid phase and also the packing characteristic, ad. And the patterns will also depend on the Reynolds number of the gas phase. Therefore,

A B

= = C = E =

19.4 0.747 i 0.147 -0.693 zk 1.095 1.968 =k 0.997

(1lb)

Equation 11 is shown in Figure 9. Because of the slight variations in viscosity and surface area, the values of C and E are unreliable, as shown by the wide confidence limits. Therefore, the viscosity and surface area were considered constant and the dispersion group was correlated with the Reynolds number only, with the following result:

Not enough variation was obtained in the Gallileo number to determine its functionality. By using data at constant gas flow rates for each packing material, the dependence of the gas-phase dispersion group on the liquid phase Reynolds number was found to be of the form :

(g)
90

= 7.58

dL

0.703

10-3

(-L -)

It was found that the value of n does not vary significantly with either rate of gas flow or type of packing. A similar treatment of the data for constant liquid rates showed that the effect of gas flow rate could be described as:

I&EC FUNDAMENTALS

0.7

0-

1/2 INCH BERL SADDLES

4Figure 9.

EQUATION I I

Liquid-phase axial dispersion

0.0 I 0.4

io

0 '/n-inch Berl saddles, present work


'/n-inch Rcischig rings, present + 10-mm. Raschig rings, McHenrywork Wilhelm ( 1 4) and

1.0
IO
-Oa6

x1 0
Figure 10.

- 0.00 259 (F) L

- - 0.785-cm.
V 1.55-cm.

I?aschig rings, Otoke and Kunugita ( 1 5 ) Raschig rings, Otake and Kunugita ( 1 5 )

Gas-phase axial dispersion

These two results can be combined as:

(:)>.

K(ReG)-m

form of this dependency cannot be determined but in this case is assumed to be of the form (ad)'. This is strictly an arbitrary choice with no foundation, but is used for convenience. Grouping the Raschig ring and Berl saddle data and fitting them to
ReL

lo-"

(12)

A least squares analysis performed on the data for each packing separately yielded the values and 95% confidence limits shown in Table I. The confidence ranges of both m and n overlap, so there is a possibility that the packing shape has no effect on these constants and that the difference in results is due to experimental error. T o test this hypothesis, a Student's t test as described by Volk (20) was performed on the data, with the result that if there were no difference in the real values of the constants, there would be one chance in six that experimental error could cause differences as great as shown in Ta.ble I. This is not a small enough chance to justify treating the Berl saddle and Raschig ring data separately. Although confidence limits on the coefficients, K , were not evaluated, it is obvious that the coefficient depends on the packing geometry. Since only two geometries were used, the

c:)c

= K(ad)S(ReG)-mX

(13)

gave the following results and 95y0 confidence limits. Number of points = 34 m = 0.668 i 0.184 n = 0.00259 0.00053 s = 2.58 f 0.78 K = 0.0585 Figure 10 shows the data plotted according to Equation 13. Because of the uncertainty as to the effects of packing geometry, the above result should be applied only to systems similar to those used in this work. DeMaria and White (7) arrived at the following correlation for I/d-, 3/g-, and 1/2-inch Raschig rings:
(:)c =

2.4 (Rec)-0.20 x 10

- (0.013 -

0.088 d ) R e L dc

(14)

Table 1. Constants from Analysis of Individual Sets of Data '/2-Inch Bed Saddles '/n-Inch Raschig Rings Constant

K
No. of data points
m n

3.72

6.23

o.oo228f 0.20 0.54 o,ooo52 o.oo299 f 0.29 0.79 o.ooo93 15 19

Because they used only one type of packing, a term to describe the effect of packing geometry was not included. H ~ they did vary the packing size and found that coefficient n in Equation 13 is a function of packing diameter. Since DeMaria and White used only one tower diameter, 4 inches, they are not justified in including the tower diameter, d,, in their correlation.
VOL. 5

NO. 1

FEBRUARY 1 9 6 6

91

2*o

1
a
G

Nomenclature

= surface area of packing per unit volume of packed bed,

sq. ft./cu. ft.

= concentration of tracer
= = = =

d
dl

nominal packing size, ft. axial dispersion coefficient

= column diameter, ft. = acceleration of gravity, ft./(sec.)2

D
g G Ga h

mass flow rate of gas phase, lb./(hr.-sq. ft.) Gallileo number, defined in Equation 10 = length of test section, ft. Hop = operating holdup, cu. ft./cu. ft. H,, = static holdup, cu. ft./cu. ft. H t = total holdup, cu. ft./cu. ft. L = mass flow rate of liquid phase, lb./(hr.-sq. ft.) Re = Reynolds number, d L / p or dG/p t = time, sec. u = axial velocity, ft./sec. u, = superficial velocity, ft./sec. x = position along bed, ft. z = dimensionless position, x / h GREEK LETTERS = ratio of length of end section to length of test section C Y 0 = defined in Equation 5 = fraction of bed volume occupied by a particular phase e e = dimensionless time, ut/h = first moment, defined by Equation 5 or viscosity when p used in Reynolds or Gallileo number p = density, lb./cu. ft. = second moment, defined by Equation 5 u SUBSCRIPTS = position of response curve = end section G = gas phase L = liqud phase
literature Cited

10 0

zoo
R e L = (L)
t L

300

Figure 1 1 . Comparison o results o present work f f with data o DeMaria and White (7), f /Z-inch Raschig rings

To compare Equation 14 with Equation 13, the appropriate values of d, a, and dt for l/Z-inch Raschig rings are used in the equations, with the result:
This work:

i E

DeMaria and White:


=

e)G

2.4 (ReG)-0.20 X 10-o*oo2 ReL

(144

A graphical comparison of the equations and the ranges covered are shown in Figure 11. For similar conditions, Equation 14 gives dispersion groups that are approximately one eighth the values given by Equation 13 based on the present investigation. Further work is needed to resolve this large disagreement. I n the meantime, since the results of the present study on Equation 13 indicate a greater deviation from plug flow, its use will give a conservative performance estimate in design applications.
Acknowledgment

The authors gratefully acknowledge the financial support of the Esso Research and Engineering Co. and the cooperation of the M. A. Knight Co., which supplied the packings.

(1) Aris, R., Amundson, N. R., A.Z.Ch.E. J . 3, 280 (1957). (2) Bischoff, K. B., Levenspiel, O., Chem. Eng. Sci. 17, 245 (1962). J., Bretton, R. H., A.Z.Ch.E. J . 4, 367 (1958). P. V., Chem. Eng. Sci.2, 1 (1953). (5) Danckwerts, P. V., Jenkins, J. W., Place, G., Zbid., 3, 26 (1954). .,(6 Deisler, P. F., Wilhelm, R. H., Znd. Eng. Chem. 45, 1219 (1953). (7 DeMaria, F., White, R. R., A.Z.Ch.E. J . 6 , 473 (1960). (8 Ebach, E. A., White, R. R., Zbid., 4,161 (1958). 9) Elgin, J. C., Weiss, F. B., Znd. Eng. Chem. 31,435 (1939). 10) Jacaues, G. L.. Vermeulen, T., University of California Radiation Laboratory, UCRL-8029 (November 1957). (11) Jesser, B. W., Elgin, J. C., Trans. Am. Znst. Chem. Engrs. 79.277 (1943). .-,(l$ Kramers, H., Alberda, G., Chem. Eng. Sci.2, 173 (1953). (13) Levenspiel, O., Smith, W. K., Zbid., 6,227 (1957). (14) McHenrv. K. W., Wilhelm, R. H., A.Z.Ch.E. J . 3, 83 (1957). 15) Otake, T.; Kunugita, E., Chkm. Eng. (Japan) 22,144 (1958). 16) Otake, T., Okada, K., Zbid., 17, 176 (1953). 17) Sater, V. E., Ph.D. thesis in chemical engineering, Illinois Institute of Technology, Chicago, lll., 1963. (18) Shulman, H. L., Ullrich, C. F., Wells, N., A.Z.Ch.E. J . 1, 247 (1955). 19 Strang, D. A., Geankoplis, C. J., Znd. Eng. Chem. 50, 1305 (1958). (20 Volk, W., Applied Statistics for Engineers, p. 260, Mcdraw-Hill, New York, 1958.

\ - - -

7 -

\--

RECEIVED review March 29, 1965 for ACCEPTED August 11, 1965

92

I&EC FUNDAMENTALS

S-ar putea să vă placă și