Sie sind auf Seite 1von 13

KEYWORD INDEX AUTHOR INDEX TITLE INDEX

8th International Conference on Role of Engineering Towards a Better Environment


(Global Environmental Changes & Sustainable Development), Alexandria, Egypt, 20 of December, 2010
th

Mass Transfer Controlled Corrosion of the Wall of Air Sparged Agitated Vessel with Ring Sparger
Yehia M.S. El-Shazly Chemical Engineering Department, Faculty of Engineering, Alexandria University, Alexandria 21544; Egypt. Abstract
Air sparged mechanical agitated vessel is an important gas liquid contactor that is used in many processes: in gas stripping, in bioreactors, in fermentation, in pulp bleaching, and in wastewater treatment to reduce the oxygen demand in waste streams. In this study, the mass transfer coefficient from the wall of an air sparged mechanically agitated vessel with Rushton turbine is studied as a function of the air superficial velocity (3, 9, 15 and 21 L/min) and the speed of rotation of the impeller (100, 200, 400 and 600 rpm). The dissolution of copper in acidified dichromate solution technique was adopted in this study. It was found that the rate of mass transfer increases with the increase of the speed of agitation and to a lower extent with the increase in the air flow rate. Moreover, at low speed of agitation (100 rpm), the mass transfer is controlled by the air flow rate, while the effect of mechanical agitation is dominating at the higher speeds. The data were correlated with the mass transfer equation:

It was found that the value of the mass transfer coefficient from this study is much higher than expected from the correlation of Calderbank and Moo-Young. Keyword: Mass transfer; Sparged vessel; diffusion; corrosion; Power consumption

1. Introduction
Air sparged agitated vessel is an important equipment that is widely used in many industrial applications where a high interfacial area between liquid and gas is required for mass and heat transfer [1]. They are used in fermenters, wastewater aeration, gas absorbers and strippers, catalytic and hydrogenation reactors The basic principle is the introduction of a gas into the liquid with the agitation by an impeller to increase the dispersion of the gas and uniformity of the mix. They are very effective in transferring mass between the two phases, subject to the constraints of physical and chemical equilibrium. Usually the gas is injected from the bottom of the liquid tank or near the impeller, as the effectiveness of the process depends primarily on the impeller speed and the gas injection rate as well as the physical properties of the system [1-2]. Different flow regimes that may occur in this process; with axial flow impellers, the upward flowing gas bubbles are allowed to rise near the shaft where the shear forces are small, so that the flow pattern and pumping effect is disrupted. Therefore, axial flow impellers are not very useful for gas dispersion. On the other hand, radial impellers are very useful for gas dispersion where the gas is introduced just below the impeller at its axis and drawn up to the blades and chopped into fine bubbles increasing the contact area [3].

Table 1 List of symbols used in the current research


A D C C0 d di g H K n N Np Pm Pm0 Psp PT Q t Re Reg Sc Sh v Vs Wall surface area; cm2 Diffusivity; cm2/s Potassium Dichromate concentration at time t; mol/L Initial Potassium Dichromate concentration; mol/L Oxygen concentration; mol/L Tank diameter; cm Impeller diameter; cm Gravity acceleration; Height of liquid; cm Mass transfer coefficient; for the depolarizer to the metal wall surface; cm/s Impeller rotation speed; rpm Dissolved oxygen flux; mol/m2s Power Number Agitation power consumption under aeration; erg/s Agitation power consumption under no aeration; erg/s Sparging power consumption; erg/s Total power consumption; erg/s Air flow rate; cm3/s Time; s Reynolds Number for impeller mixing Reynolds Number for air sparging Schmidt Number Sheerwod Number Superficial air velocity; cm/s Liquid volume; cm3 Liquid viscosity; g/cm s Liquid density; g/cm3

Previous studies on the performance of the different combinations of ring sparger and Rushton radial impeller reported that the best results in terms of gas handling and mass transfer are obtained with the sparger located below the impeller and with a diameter at least equal to the impeller diameter [4-5]. Mass transfer from the wall of the vessel can be an important consideration in the design of the reactor; most corrosion processes in industrial applications are diffusion controlled [6-8]. Corrosion of metals as a material of construction involves the formation of macroscopic or microscopic galvanic cells where two simultaneous reactions are taking place; the anodic dissolution of the metal: MM2++2e; and cathodic reduction of the oxidizing agent such as dissolved oxygen: O2+2H2O+4e4OH- . Side reactions between the cathodic product and the anodic product may lead to the formation of porous solid film of corrosion product on the corroding metal surface; Fe(OH)2 or the oxidized form Fe(OH)3, through which oxygen must diffuse for the corrosion process to continue. Under these conditions, the rate of corrosion is [7, 9]: 2

(1) Moreover, the analogy between mass and heat transfer is a well known phenomenon since the work of Colburn [10], whereas the similarity of the mathematical formulations of both the heat flux and mass flux enable the estimation of one by knowing the second, which has been used to study different systems of transfer [11-16]. Many techniques have been used for the study of the mass transfer, among them the dissolution of solid into liquid [17-23], the electrochemical technique [24-27], adsorption [28-29] and CFD [30-33].

2. Effect of power on mass transfer


The rate of mass transfer is related to the power consumed in causing the turbulence responsible for this process: according to Calderbank and Moo-Young [34], the power consumed in driving the process affects the rate of mass transfer. They presented their equation to estimate the mass transfer coefficient between a fixed phase and its surrounding fluid which is in turbulent conditions as:

(2)

3. Power consumption in sparged agitated vessels The power requirement for the mixing and agitation is an important consideration in the design and scaleup of reactor and mixing vessels [2]. For the current case, two equipment are consuming power: the motor responsible for driving the shaft of the turbine, and the compressor used for injecting the air in the liquid. The power used for mechanical agitation can estimated from the following equation [35-36]: (3) and the value of Np reported for the six blade Rushton turbine is 5.75. For the case of presence of air sparging along the mechanical agitation, the power used for mixing is decreased and can be estimated according to the equation [37]:

(4) For the case of air sparging, the power required to drive the air against the liquid pressure can be calculated as [38]:

(5) 3

The total power is the sum of its two components: mechanical mixing and air sparging: (6) 4. Materials and Methodology

4.1. Experimental technique To study the rate of mass transfer from the wall of the vessel under mechanical agitation and air sparging, the dissolution of copper in acidified dichromate method is adopted[39]. The arrival of the oxidizing agent, the Cr6+, to the surface of the copper wall, determines the rate of the reaction: (7) This technique has been used widely in the study the role of the surface geometry and hydrodynamics in the rate of diffusion controlled corrosion in view of simplicity and accuracy [7, 40-41]. To report the rate of mass transfer from the bulk of the liquid to the surface of the wall, the method of dimensional analysis is adopted, and a correlations in terms of Sh, J, Re, Fr, and Sc numbers are reported depending on the conditions. For example, for the case of agitated vessels, a correlation in the form Sh number as a function of Re and Sc is usually reported [7-8, 42] (8) While for the case of gas sparging, a correlation in the form of (9) is usually reported [40, 43-44]. While in the case of the presence of mechanical agitation and air sparging, the correlation is reported is the form [45]: (10) where, x1, x2, x3, , , , and are constant to be determined experimentally depending on the geometry and conditions. The value of the power of Schmidt number is usually fixed at 0.33[9].

4.2. Experimental set-up Figure 1 shows the set up of the experiment: A tank made of plexiglass with it internal wall covered with a pure copper sheet to form the inner wall of the tank. The outer side of the sheet is completely isolated to insure that the exposed area is only the inner one. Four vertical plexiglass baffles were attached to the copper wall to prevent swirl flow of the solution. The Rushton turbine and its shaft are completely isolated with an epoxy coating. The shaft is connected to a variable speed motor with a digital controller. 4

The air is introduced through a ring sparger of 7.5 cm in diameter and is fixed at 2 cm below the turbine. It is connected through a plastic tube that is attached to the wall of the tank to a flowmeter and then to the air compressor. Table 2 summarize the dimensions used for the agitated vessel. The speed of rotation tested was 100, 200, 400 and 600 rpm and the air flow rate was 3, 9, 15, and 21 liter/min. A solution of 1 Molar sulphuric acid and 0.003 Molar potassium dichromate was used. The volume of the solution was 2.7 liter in each run, and a 5 mL sample was drawn every 5 minutes for analysis. The dichromate concentration was determined by titration against standard ferrous ammonium sulphate using diphenylamine indicator [46]. The density and viscosity of the solution were determined using a density bottle and an Ostwald viscometer respectively to be 1.061 g/cm3 and 0.0107 g/cm s with 6% reproducibility [47]. The diffusivity of the dichromate was taken from literature [48]. ANALAR grade chemicals and distilled water were used in the experiments.

1. 2. 3. 4. 5. 6.

DC Motor Speed controller Flowmeter Baffles Rushton turbine Ring sparger

Figure 1 Schematic diagram of the experimental setup.

Table 2 Dimensions used in the experimental setup

5. Results and Discussion


The rate of the mass transfer diffusion of the wall of the vessel is expressed in terms of the mass transfer coefficient which can be obtained from the dichromate concentration data. For these experiments, the rate of reaction is given by:

(11) Integrating this equation gives:

(12) 15 cm Figure 2 shows a Diameter of the vessel (d) typical ln(C0/C) vs. time plot for the effect of the Diameter of the Impeller (di) degree of agitation on 5 cm the rate of dichromate Diameter of the sparger (dS ) deletion, while Figure 3 7.5 cm shows the effect of air Height of solution in the vessel (H) sparging on dichromate 15 cm depletion. It is seen that the increase in either the Impeller clearance(H2) 5 cm speed of rotation or the air flow rate result in Sparger clearance(C) 2 cm increasing the rate of dichromate arrival to the 1 cm wall. Figure 4 shows Width of Baffles (Bw ) the combined effect of the mechanical agitation and air sparging on the calculated value of the mass transfer coefficient, K. As expected, the mass transfer coefficient increases with the increase of the rotation speed and the air flow rate. This is attributed to the increase of turbulence resulting from the eddies generated by the movement of the impeller or from the bubbles coalescence and breakdown. This turbulence results in the decrease of the thickness of the hydrodynamic boundary layer and the faster transport of the oxidizer to the wall of the vessel. It is also noted a higher increase in mass transfer coefficient upon increasing the rate of mechanical agitation than on the increase of the air flow rate for the tested range.

Figure 2 ln(C0/C) versus time curves for the different mechanical agitation speed.

Figure 3 ln(C0/C) versus time curves for the different air flow rate.

Figure 4 Combined effect of mechanical agitation and air sparging on the rate of mass transfer from the wall of the vessel.

For the effect of air sparging, Figure 5 shows that the mass transfer coefficient is related to the superficial air velocity by the equation: (13) It is seen that the exponent of the velocity is lower than reported in literature [40, 49]. This might be attributed to the fact that the diameter of the sparger is smaller than the diameter of the tank, with the result that the air bubbles flowing out of the sparger are more concentrated at the centre of the tank than at the wall, thus resulting in higher turbulence in the bulk zone away from the wall. Moreover, due to the fact that the air bubbles move to the upward direction, a dead zone where no turbulence occurs is formed below the sparger, makes the actual affected area by the turbulence lower than the wall area and the lower area of the tank not affected by gas bubbles. Figure 6 shows that for the case of sparging alone, the mass transfer correlation is: (14) Figure 7 shows that for the case of mechanical agitation in the absence of air sparging, the non dimensional mass transfer correlation can be represented as: (15)

Figure 5 Effect of superficial air velocity on the rate of mass transfer for the case of air sparging alone.

Figure 6 Mass transfer correlation for the case of air sparging alone.

Figure 7 Mass transfer correlation for the case of mechanical agitation. 9

The mass transfer correlation for the combined mechanical agitation and gas sparging is ():

(16) with a correlation coefficient equal to 0.92. It is seen from this correlation that the mass transfer is affected by the combined effect of the mechanical agitation and air sparging. The higher exponent of the Re compared to that of Reg shows that under the present range of conditions mechanical stirring is more influential than gas sparging in enhancing the rate of mass transfer.

Figure 8 Overall mass transfer correlation for the case of the combined effect of mechanical agitation and air sparging.

Figure 9 shows a comparison between the values of K obtained from the current study with the values obtained from Calderbank and Moo-Young correlation (equation 2) [34]. For the power calculations, equations 3 through 6 were used. At low mechanical agitation speed, a small difference between the experimental and the calculated values of K is found, though upon increasing the air flow rate, the difference increases in favor for the experimental results. On increasing the rotation speed, the difference increases largely. This discrepancy may be attributed to that the value of the rate of mass transfer in this case is not only a result for the decrease in the thickness of the hydrodynamic layer at the wall from the agitation turbulence, but also the radial flow resulting from the impeller, directs the air bubbles outwards where they impinge on the wall, exposing fresh surface for mass transfer and enhancing the turbulence near the wall. This is different from the isotropic turbulence for which Calderbank equation applies.

6. Conclusions
In this study, the rate of mass transfer of the wall of a mechanical agitated sparged vessel is examined. A mass transfer correlation relating the mass transfer to the turbulence and physical properties of the solution is reported. It is found that the rate of mass transfer increases with the increase in speed of rotation, and to a lower extent, with the increase in the air flow rate. The experimental results showed a higher mass transfer coefficient than expected from Calderbank and Moo-Young correlation. 10

; cm/s
Figure 9 Comparison of the mass transfer coefficient estimated from the Calderbank and Moo-Young correlation and the mass transfer coefficient obtained from the current study.

References
1. 2. 3. 4. 5. 6. 7. 8. Richardson, J.F. and D.G. Peacock, Coulson and Richardson's Chemical Engineering Volume 3 Chemical and Biochemical Reactors and Process Control (3rd Edition), Elsevier. McCabe, W.L., J.C. Smith, and P. Harriott, Unit Operations of Chemical Engineering. Seventh ed. McGraw-Hill Chemical Engineering Series2005: McGraw-Hill Jakobsen, H.A., Chemical Reactor modeling : Multiphase Reactive Flows2008: Springer. Birch, D. and N. Ahmed, The Influence of Sparger Design and Location on Gas Dispersion in Stirred Vessels. Chemical Engineering Research and Design, 1997. 75(5): p. 487-496. Sardeing, R., et al., Gas-Liquid Mass Transfer: Influence of Sparger Location. Chemical Engineering Research and Design, 2004. 82(9): p. 1161-1168. Sedahmed, G.H., et al., Mass transfer at a pipe inlet zone in relation to impingement corrosion. International Communications in Heat and Mass Transfer, 1998. 25(3): p. 443-451. Sedahmed, G.H., et al., Mass transfer at the impellers of agitated vessels in relation to their flowinduced corrosion. Chem. Eng. J., 1998. 71(1): p. 57-65. El-Shazly, Y.M., et al., Mass transfer in relation to flow induced corrosion of the bottom of cylindrical agitated vessels. Chemical Engineering and Processing, 2004. 43(6): p. 745-751. 11

9. 10. 11. 12. 13.

14. 15.

16.

17. 18.

19. 20.

21. 22. 23. 24. 25.

26.

27. 28.

Kelly, R.G., et al., Electrochemical Techniques in Corrosion Science and Engineering. Corrosion Technology, ed. P.A. Schweitzer2002: Marcel Dekker, Inc. Colburn, A.P., A method of correlating forced convection heat-transfer data and a comparison with fluid friction. International Journal of Heat and Mass Transfer, 1964. 7(12): p. 1359-1384. Bedingfield, C.H. and T.B. Drew, Analogy between Heat Transfer and Mass Transfer. Industrial & Engineering Chemistry, 1950. 42(6): p. 1164-1173. Han, S. and R.J. Goldstein, The heat/mass transfer analogy for a simulated turbine blade. International Journal of Heat and Mass Transfer, 2008. 51(21-22): p. 5209-5225. Boukadida, N. and S. Ben Nasrallah, Mass and heat transfer during water evaporation in laminar flow inside a rectangular channel -- validity of heat and mass transfer analogy. International Journal of Thermal Sciences, 2001. 40(1): p. 67-81. El-Riedy, M.K., Analogy between heat and mass transfer by natural convection from air to horizontal tubes. International Journal of Heat and Mass Transfer, 1981. 24(3): p. 365-369. Ambrosini, W., et al., On various forms of the heat and mass transfer analogy: Discussion and application to condensation experiments. Nuclear Engineering and Design, 2006. 236(9): p. 1013-1027. Bieniasz, B., Convectional mass/heat transfer in a rotary regenerator rotor consisted of the corrugated sheets. International Journal of Heat and Mass Transfer, 2010. 53(15-16): p. 31663174. Duda, J.L. and J.S. Vrentas, Heat or mass transfer-controlled dissolution of an isolated sphere. International Journal of Heat and Mass Transfer, 1971. 14(3): p. 395-407. Shehata, A.S., S.A. Nosier, and G.H. Sedahmed, The role of mass transfer in the flow-induced corrosion of equipments employing decaying swirl flow. Chemical Engineering and Processing, 2002. 41(8): p. 659-666. Rakoczy, R., Enhancement of solid dissolution process under the influence of rotating magnetic field. Chemical Engineering and Processing: Process Intensification, 2010. 49(1): p. 42-50. Shen, G.C., C.J. Geankoplis, and R.s. Brodkey, A note on particle--liquid mass transfer in a fluidized bed of small irregular-shaped benzoic acid particles. Chemical Engineering Science, 1985. 40(9): p. 1797-1802. Guo, G. and K.E. Thompson, Experimental analysis of local mass transfer in packed beds. Chemical Engineering Science, 2001. 56(1): p. 121-132. Turitto, V.T., Mass transfer in annuli under conditions of laminar flow. Chemical Engineering Science. 30(5-6): p. 503-509. Goldstein, R.J. and H.H. Cho, A review of mass transfer measurements using naphthalene sublimation. Experimental Thermal and Fluid Science, 1995. 10(4): p. 416-434. Sedahmed, G.H. and I. Nirdosh, Free convection mass transfer at horizontal cylinders with active ends. International Communications in Heat and Mass Transfer, 1990. 17(3): p. 355-366. Zaki, M.M., I. Nirdosh, and G.H. Sedahmed, Mass transfer characteristics of reciprocating screen stack electrochemical reactor in relation to heavy metal removal from dilute solutions. Chemical Engineering Journal, 2007. 126(2-3): p. 67-77. Sara, O.N., et al., Electrochemical mass transfer between an impinging jet and a rotating disk in a confined system. International Communications in Heat and Mass Transfer, 2008. 35(3): p. 289-298. Tzanetakis, N., et al., Mass transfer characteristics of corrugated surfaces. Applied Thermal Engineering, 2004. 24(13): p. 1865-1875. Sonetaka, N., et al., Characterization of adsorption uptake curves for both intraparticle diffusion and liquid film mass transfer controlling systems. Journal of Hazardous Materials, 2009. 165(13): p. 232-239. 12

29.

30.

31.

32. 33. 34. 35. 36.

37. 38. 39. 40. 41. 42.

43. 44. 45.

46. 47. 48. 49.

Fujiki, J., et al., Experimental determination of intraparticle diffusivity and fluid film mass transfer coefficient using batch contactors. Chemical Engineering Journal, 2010. 160(2): p. 683690. Santos, J.L.C., et al., Characterization of fluid dynamics and mass-transfer in an electrochemical oxidation cell by experimental and CFD studies. Chemical Engineering Journal, 2010. 157(2-3): p. 379-392. Dixon, A.G., M. Nijemeisland, and E.H. Stitt, Packed Tubular Reactor Modeling and Catalyst Design using Computational Fluid Dynamics, in Advances in Chemical Engineering, B.M. Guy, Editor 2006, Academic Press. p. 307-389. Li, F., et al., Experimental validation of CFD mass transfer simulations in flat channels with nonwoven net spacers. Journal of Membrane Science, 2004. 232(1-2): p. 19-30. Esteban Duran, J., F. Taghipour, and M. Mohseni, CFD modeling of mass transfer in annular reactors. International Journal of Heat and Mass Transfer, 2009. 52(23-24): p. 5390-5401. Calderbank, P.H. and M.B. Moo-Young, The continuous phase heat and mass-transfer properties of dispersions. Chem. Eng. Sci., 1961. 16: p. 39-54. Coulson, J.M. and J.F. Richardson, Chemical Engineering. 6th ed. Vol. 1. 1999: Butterworth Heinemann Garcia-Ochoa, F.F. and E. Gomez, Theoretical prediction of gas-liquid mass transfer coefficient, specific area and hold-up in sparged stirred tanks. Chemical Engineering Science, 2004. 59(12): p. 2489-2501. Michel, B.J. and S.A. Miller, Power requirements of gas-liquid agitated systems. AIChE Journal, 1962. 8(2): p. 262-266. Joshi, J. and L. Doraiswamy, Chemical Reaction Engineering, in Albright's Chemical Engineering Handbook2008, CRC Press. Gregory, D.P. and A.C. Riddiford, Dissolution of Copper in Sulfuric Acid Solutions. J. Electrochem. Soc., 1960. 107(12): p. 950-956. Nosier, S.A., Diffusion-controlled corrosion of gas agitated vessel under forced convection conditions. Materials Letters, 1997. 31(3-6): p. 291-296. El-Shazly, Y.M., et al., Mass transfer in relation to flow induced corrosion of the bottom of cylindrical agitated vessels. Chem. Eng. Process., 2004. 43(6): p. 745-751. Sedahmed, G.H., et al., A mass transfer study of the electropolishing of metals in mechanically agitated vessels. International Communications in Heat and Mass Transfer, 2001. 28(2): p. 257265. Zarraa, M.A., et al., Effect of gas sparging on the rate of mass transfer at a single sphere. The Chemical Engineering Journal, 1991. 47(3): p. 187-191. Noseir, S.A., et al., Solid - liquid mass transfer at gas sparged fixed bed of rasching rings. International Communications in Heat and Mass Transfer, 1997. 24(5): p. 733-740. Konsowa, A.H., et al., Mass transfer at gas sparged interface between two immiscible nondispersed liquids in batch and continuous reactors. Chemical Engineering Journal, 2004. 102(2): p. 131-140. Mendham, J., et al., Vogel's Quantitative Chemical Analysis. 6 ed2000: Prentice Hall. Findlay, A. and B.P. Levitt, Practical physical chemistry. 9 ed1972: Wiley. Gregory, D.P. and A.C. Riddiford, Transport to the surface of a rotating disc. J. Chem. Soc., 1956: p. 3756-3764. Sedahmed, G.H., et al., Effect of gas sparging on the rate of mass transfer during electropolishing of vertical plates. Chemical Engineering and Processing, 2001. 40(3): p. 195-200.

13

Das könnte Ihnen auch gefallen