Sei sulla pagina 1di 4

APCA NOTE-BOOK

References
1. G. L. Hobart, Hines Oil Co., Topsham, ME, private communication. 2. D. B. Turner, Workbook of Atmospheric Dispersion Estimates, P.H.S. publication 999-AP-26, U. S. Dept. of Health Education and Welfare, Cincinnati, 1969.

3. V. Sharma, "An area source model for urban air pollution application," Atmos, Environ. 10:1027 (1976). 4. S. R. Hanna, "A simple method of calculating dispersion from urban area sources," J. Air Poll. Control Assoc. 21: 774 (1971). 5. Compilation of Air Pollutant Emission Factors, 2nd ed., U. S. Environmental Protection Agency, Research Triangle Park, 1975.

6. M. Feldstein, in Combustion-Generated Air Pollution, ed. E. S. Starkman, Plenum Press, New York, 1971. pp. 291-318. Dr. Butcher is a professor of Chemistry and Mr. Buckley is a teaching fellow at Bowdoin College, Brunswick, ME 04011.

Pressure Loss in Venturi Scrubbers

Shui-Chow Yung, Harry F. Barbarika and Seymour Calvert


Air Pollution Technology, Inc. San Diego, California

Venturi scrubbers are used extensively for the collection of small particles. The major limitation on their use is that they have high pressure drops which result in high operating costs. With a better understanding of the fundamentals of the operation of this type of scrubber, the pressure drop could be optimized for a given level of particle collection.

Nomenclature

cD d
gc

It

p Q u
X

= drag coefficient, dimensionless = diameter, cm = conversion factor = throat length, cm = pressure, dyne/cm2 = volumetric flow rate, cm3/sec = velocity, cm/sec = axial coordinate, cm

z Greek
a P M

= surface tension, dyne/cm = density, g/cm3 = viscosity, poise Subscripts d = liquid drop G = gas L = liquid 1,2 = throat entrance and exit, respectively Dimensionless numbers NRe = Reynolds number based on drop diameter
348

The pressure drop for gas flowing through a venturi scrubber is due to the friction loss along the wall of the scrubber and the acceleration of liquid drops. Friction loss depends largely upon the geometry of the scrubber. Acceleration losses, which are frequently predominant in the venturi scrubber pressure drop, are fairly insensitive to scrubber geometry and in most cases can be predicted theoretically. Currently, there are several correlations available, both theoretical and experimental, for the prediction of pressure drop in a venturi scrubber. Correlations by Matrozov,1 Yamauchi, et al.,2 Volgin, et al.,3 Gleason, et al.,4 and Hesketh5 are experimental correlations. Matrozov's correlations and Volgin's correlation were obtained mainly on small size venturi scrubbers. Yamauchi's correlation was based on experimental data taken from a venturi scrubber with high temperature gas flow (100-900C). Hesketh's correlation is an experimental correlation he obtained after he had evaluated data obtained from many fixed throat venturi scrubbers. Equations proposed by Yoshida, et al.,6-7 Calvert,8 Tohata, et al.Q Boll,10 Behie and Beeckmans11 are theoretical correlations. All of the equations were derived from the equations of motion and momentum balance. Geiseke's equation accounts for the mass transfer between liquid and gas. Boll's equation10 is similar to that of Geiseke's except Boll had neglected the mass transfer between phases. Equations by Tohata, et a/.,9 Yoshida, et al.,6'7 and Boll10 contain terms attributed to wall friction. While Tohata, et al.,9 and Yoshida, et al.,6'7 have used different values for the friction factors in the convergent, throat and divergent sections, Boll10 suggested a single value, 0.027, for all sections. Of all the equations, Calvert's equation8 is the easiest to apply. Calvert derived his equation by use of Newton's law and
Journal of the Air Pollution Control Association

assumed that all liquid drops were accelerated to the gas velocity in the throat and that none of the drop momentum was converted to pressure in the diffuser. Since, in industrial venturi scrubbers, liquid drops do not have sufficient residence time to accelerate to the gas velocity in the throat, Calvert's equation predicts a higher pressure drop than is experimentally measured except at low liquid rates, where the frictional losses for the gas alone are significant. This paper presents a modification to Calvert's equation which takes into account the fact that liquid drops are not fully accelerated to the gas velocity in the throat.
Mathematical Model

where
(9)

The pressure loss in the venturi scrubber is then equal to:


AP =

_ ^ f l d /SL\ (i _ X2 + V(jc4-jc2)-r>)
gc
\QG'

(10)

Comparison of Model Predictions with Experimental Data

To simplify the problem, we will neglect the pressure loss due to wall friction and pressure recovery by the gas in the divergent section. This simplification is acceptable since wall friction is compensated to some extent by the pressure recovery, For this case, the pressure loss in a venturi scrubber is equal to the momentum expended to accelerate the liquid in the venturi throat. Other assumptions are: (1) liquid is introduced at the throat entrance with no axial velocity; (2) liquid is atomized at the injection point and is uniform size with Sauter mean diameter predicted by the Nukiyama and Tanasawa equation:12
dd =

omm / q a. + Q Q597 r_2i_]/to

(1)

There are numerous venturi scrubber performance data reports in the literature. However, most of these data are of a limited value for testing the model. This is because some important information such as scrubber geometry, water injection method, etc., are generally not given in those reports. In the following paragraphs, only those data which have been obtained under controlled conditions and for a well-defined scrubber system will be utilized to test the model. Boll10 derived an equation to predict the pressure drop in a venturi scrubber. He obtained some pressure drop data on a prototype venturi scrubber to confirm his equation. The cross-sectional dimensions of the venturi scrubber throat was 35.6 X 30.5 cm. The lengths of the convergent section, throat, and divergent section were 91.4, 30.5, 337.8 cm respectively. The convergence angle was 25 and the divergence angle was 7.

(3) the flow is one-dimensional, incompressible and adiabatic; (4) liquid fraction at any cross-section is small. A momentum balance around a control volume of differential length yields:
dP=-

2.0

1.8 -

o 1.27 cm nozzles
A

W^

(2)
1.6

1.91 cm nozzles

Upon integration, we obtain


(3)

This equation gives the pressure expended to accelerate the liquid from zero velocity to Ud2, the drop velocity at the throat exit. If the drop is accelerated to the gas velocity, this equation reduces to Calvert's equation. The drop velocity at the throat exit can be predicted by taking into account the acceleration of the drops from their initial velocity. A force balance on the liquid drop provides the differential equation of motion for the drop. dud 3PG (4) dz 4 PL We will use the drag coefficient expression given by Hollands and Goel13 as:
= CD1(
\UG
UG

1.4

1.2

d 1.0

0.8

(5)

Eq. 5 approximates the "standard curve" in the Reynolds numbers range from 10 to 500. Cm is the drag coefficient at the point of liquid injection, i.e. at the throat entrance. It can be obtained from the standard curve or by the following formula given by Dickinson and Marshall.14 CD1 = 0.22 +
24

a*3.2
Figure 1.

0.6 0.8 1.0 1.2 Liquid to gas ratio, / m 3 of gas Comparison of Boll's pressure drop data10 with predictions.

(1 + 0.

(6)

By substituting Eq. (5) into Eq. (4) and after rearrangement, we obtain:
Uddud pLd (7)

Figure 1 shows the comparison between Boll's data and predictions by Boll's equation and the present model. The pressure drop is expressed in terms of the number of throat velocity heads, i.e.: Number of velocity heads lost =
-AP
PGUG'

(11)

Upon integration over the length of the venturi throat, we have


Ud2 ~ April 1977 Volume 27, No. 4 (8)

The range of gas velocity in the throat covered by Boll is 45.8-91.5 m/sec. The line representing the predictions by Boll's equation is an average curve for the predicted pressure drops. Even though Boll's equation slightly underestimates
349

APCA NOTE-BOOK

pressure drop at low values of liquid-to-gas ratio, the agreement between his data and equation is considered to be satisfactory. Predicted pressure drops by the present model for throat velocities of 45.8 and 91.5 m/sec are shown in Figure 1. Since Boll operated his venturi scrubber in this throat velocity range, the area bounded by these two theoretical lines represents the range of expected pressure drops. The fact that Boll's theoretical curve, which is the average of the predicted pressure drops, lies within this area signifies that the present model and the more complicated Boll's equation predict pressure drop with equal accuracy. Figures 2 and 3 compare Brink and Contant's data15 and data by Ekman and Johnstone16 with predictions. The venturi scrubber used by Brink and Contant15 was a Pease-Anthony venturi scrubber. The scrubber had a rectangular cross-section. Its throat measured 15.2 X 86.4 cm with a straight section 30.5 cm long. The angle of the convergent section was 25; the divergent section had an angle of 2.2 for 152 cm following the venturi throat and then an angle of 15.

The model neglects the wall friction and the pressure recovery in the diffuser. Since these two factors do not exactly cancel each other, the accuracy of the prediction by this model depends on the actual values of these factors. At low liquid flow rates, wall frictional losses are more significant, so that the model might underestimate the pressure drop. The reverse is true at high liquid flow rates. In either case, the predictions by the present model are within 10% of that predicted by Boll's equation.
Acknowledgment

This study was supported by E.P.A. under Contract No. 68-02-1328, Task No. 13. Dr. Leslie E. Sparks was the project officer.

1.4

1.2
Boll's equation

2.5

I 0.8
0.6

0.4 o 2.0 Present model 0.2 o Data for radial inward injection A Data for radial outward injection I 0.2 I 0.4 I I I 1 0.6 0.8 1.0 Liquid to gas ratio, / m 3 I I I

1.2

1.4

Figure 3. Comparison of Ekman and Johnstone's pressure drop data16 with predictions.

o 90 spray jet A 45 spray jet

References 1. V. I. Matrozov and O. Soobscheniya, Nauchno-Tekhnicheskikh Rabotakh NIVIF, Nos. 6/7,152, 1953. 2. J. Yamauchi, T. Wada and H. Kamei, "Pressure drop across the venturi scrubbers," Kagaku Kogaku, 27: 974 (1963). 3. B. P. Volgin, T. F. Efimova, and M. S. Gofman, "Adsorption of sulfur dioxide by ammonium sulfite-bisulfite solution in a venturi scrubber," Intern. Chem. Eng. 8: 113 (1968). 4. R. J. Gleason and J. D. McKenna, paper presented at the 69th National Meeting of AIChE, Cincinnati, OH, 1971. 5. H. E. Hesketh, "Atomization and Cloud Behavior in Wet Scrubbers," US-USSR Symp. on Control of Fine Particulate Emissions, Jan. 15-18, 1974. . 6. T. Yoshida, N. Morishima, and M. Hayashi, "Pressure loss in gas flow through venturi tubes," Kagaku Kogaku, 24: 20 (1960). 7. T. Yoshida, N. Morishima, M. Suzuki, and N. Hukutome, "Pressure loss for the acceleration of atomized droplets," Kagaku Kogaku, 29: 308 (1965). 8. S. Calvert, "Source Control by Liquid Scrubber," Chap. 46 in Air Pollution, Edited by Arthur Stern, Academic Press, New York, 1968. 9. H. Tohata, T. Nakoda, and I. Sekiguchi, "Pressure losses on venturi -mixers, "Kagaku Kogaku 28: 64 (1964). 10. R. H. Boll, "Particle collection and pressure drop in venturi scrubbers," bid. Eng. Chem. Fundamentals, 12: 40 (1973). 11. S. W. Behie and J. M. Beeckmans, "On the efficiency of a venturi scrubber," Canadian J. Chem. Eng. 51:430 (1973). 12. S. Nukiyama and Y. Tanasawa, Trans. Soc. Mech. Engrs. (Japan), 4 (14): 86 (1937); 4 (15): 137 (1938); 5 (18): 68 (1939); 6 (22): II-7 (1940). Journal of the Air Pollution Control Association

1.0 1.0

I I I 1.2 1.4 1.6 1.8 3 Liquid to gas ratio, /m of gas

2.0

Figure 2. Comparison of Brink and Contant experimental data15 with predictions.

The venturi scrubber used by Ekman and Johnstone10 was a laboratory scale scrubber. The throat was 3 cm in diameter and 3.8 cm long. The convergent and divergent angles were 25 and 7 respectively. As can be seen from these two figures, both Boll's equation and the present model, which is much simpler to use, are in good agreement with experimental data.
Conclusions

Pressure loss in a venturi predicted by the present mathematical model, Eq. 10, is in good agreement with available experimental data. Thus, the model can be used for venturi scrubber design and optimization.
350

13. K. G. T. Hollands and K. C. Goel, "A general method for predicting pressure loss in venturi scrubbers," Ind. Eng. Chem. Fundamentals 14: 16 (1975). 14. D. R. Dickinson and W. R. Marshall, "The rates of evaporation of sprays," AIChEJ. 14:541 (1968). 15. J. A. Brink and C. G. Contant, "Experiments on an industrial venturi scrubber," Ind. Eng. Chem. 50: 1157 (1958). 16. F. 0. Ekman and H. F. Johnstone, "Collection of aerosols in a venturi scrubber," Ind. Eng. Chem. 43: 1358 (1951).

Dr. Calvert is president and Drs. Yung and Barbarika are associated with Air Pollution Technology, Inc., 4901 Morena Boulevard, Suite 402, San Diego, CA 92117.

Air Pollution Control and Instrumentation Patents

Robert W. Mcllvaine
The Mcllvaine Company

An analysis of air pollution control and instrumentation patents issued in the last twelve months leads to some rather startling conclusions. First, there may be some truth in the speculation that foreign technology is rapidly gaining on the U. S. Over 39% of all the air pollution control patents issued by the U. S. Patent Office were assigned to non-U.S. corporations and individuals. Secondly, despite the complaints about inadequate protection under present patent law, air pollution control companies are flocking to the Patent Office in greatly increased numbers. Approximately one air pollution control patent is issued every day for a total of nearly 400/year. A patent search for the 1962-1972 period uncovered only 150 scrubber patents for an average of 15/yr. But, in the last twelve months, nearly 80 patents have been issued in this area alone. An increase of nearly 500%. The largest single subject area, incineration devices, comprises 28% of all patents issued. This is shown in Table I. Incineration devices include nearly the whole gamut of automobile pollution control devices and this accounts for the unusually large percentage. SO2 removal systems represent a high percentage of the scrubber total. Much technical information is available from patents, and patents covering the same general area tend to be issued at the same time. To illustrate this point, we have selected wet electrostatic precipitator patents to explore in greater detail. Between May 25,1976 and June 1,1976, six patents were issued in the area of wet electrostatic precipitation. Three were assigned to the U.S. Filter Company, one to the Ceilcote Company, one to Alvin M. Marks, and one to TRW. Inter-

estingly, at least three of the patents cover equipment which is now actually in operation, because it often takes two years after a patent is filed before it is actually issued. The wet electrostatic precipitator described in the three U. S. Filter patents has been installed on coke ovens and other applications. One of these patents, 3,960,689, deals with the removal and recovery of sulfur dioxide from the gas stream by use of the precipitator, using an ammonium phosphate solution. Patent 3,958,958 deals with the Ceilcote Ionizing Wet Scrubber which recently has experienced increasing use. The unit consists of a packed wet scrubber through which a scrubbing liquid such as water is flowed vertically downward. The stream of gas to be treated is ionized prior to its flow through the wet scrubber to provide particles in the gas stream with an electrical charge of a given polarity, usually negative. Upon flow of the gas stream through the wet scrubber, the charged particles in the gas stream are carried into close proximity with, and are attracted to, the scrubbing liquid and/or packing elements. This scrubber is currently being installed on a refractory kiln. Specific test information is often included as in the TRW patent, 3,958,959. This device uses charged droplets for col-

Table I. Patent subject areas. Equipment type Scrubbers Fabric filters Electrostatic precipitators Adsorbers Incineration devices Mechanical & inertial collectors Other control methods Instrumentation % of Total 23% 11% 7% 5% 28% 6% 15% 5%

Mr. Mcllvaine is president of The Mcllvaine Company, 2970 Maria Avenue, Northbrook, IL 60062.

April 1977

Volume 27, No. 4

351

Potrebbero piacerti anche