Acessibilidade / Reportar erro

Hydrodynamics of an external-loop airlift reactor with inserted membrane

Abstract

The objective of this study was to investigate the hydrodynamics of an external-loop airlift membrane reactor (ELAMR). The ELAMR was operated in two modes: without (mode A) and with bubbles in the downcomer (mode B), depending on the liquid level in the gas separator. The influence of the gas distributor’s geometry and various alcohol solutions on the hydrodynamics of the ELAMR was studied. Results for the gas holdup and the downcomer liquid velocity are commented with respect to an external-loop airlift reactor of the same geometry but without the membrane in the downcomer (ELAR). Due to the presence of the membrane in the downcomer, acting as the local hydrodynamic resistance, the gas holdup in the riser of the ELAMR increases maximally by 16%, while the liquid velocity in the downcomer decreases up to 50%. The values of the gas holdup and liquid velocity predicted by the application of empirical power law correlations and a feed forward back propagation neural network (ANN) are in very good agreement with experimental values.

Keywords
membrane air lift reactor; hydrodynamics; dilute alcohol solution; geometry of distributor; multichannel membrane

INTRODUCTION

Over the past few decades airlift reactors have been widely used in chemical and biotechnological processes, wastewater treatment and various fermentation systems. The airlift reactors have many advantages, such as the self-generated liquid circulation caused by hydrostatic pressure differences, simple construction, good heat transfer, good mixing, low costs, absence of moving parts, high capacity, relatively low energy consumption and minimal space requirements. Also, they easily enable fluidization of solid particles. There are many variations of these reactors, but the two major groups are: internal loop airlift reactors (ILAR) and external-loop airlift reactors (ELAR).

The progress in waste water treatment, chemical processes and biotechnology industries requires reactors with in situ product removal. In situ separation processes recover desired product or remove harmful components from the reactor. Therefore, airlift reactors were combined with separation processes such as: liquid−liquid extraction (Gianetto et al., 1988Gianetto, A., Ruggeri, B., Specchia, V., Sassi, G., Forna, R., Continuous extraction loop reactor (CELR): Alcoholic fermentation by fluidized entrapped biomass. Chem. Eng. Sci., 43, 1891-1896 (1988).; Yuan et al., 2001Yuan, Y.-J., Wei, Z.-J., Wu, Z.-L., Wu, J.-C., Improved Taxol production in suspension cultures of Taxus chinensis var. mairei by in situ extraction combined with precursor feeding and additional carbon source introduction in an airlift loop reactor. Biotechnol. Lett, 23, 1659-1662 (2001).) and adsorption or ion exchange (Sun et al., 1999Sun, Y., Li, Y. L., Bai, S., Hu, Z. D., Modeling and Simulation of an In Situ Product Removal Process for Lactic Acid Production in an Airlift Bioreactor. Ind. Eng. Chem. Res., 38, 3290-3295 (1999).). Nowadays, in situ membrane separation process coupled with reactor has received considerable attention (Carstensen et al., 2012Carstensen, F., Apel, A., Wessling, M., In situ product recovery: Submerged membranes vs. external loop membranes. J. Membr. Sci., 394-395, 1-36 (2012).). In this integrated process, both the biomass formation and liquid separation with soluble products take place in one membrane bioreactor (MBR). This process has many advantages, such as an excellent and stable effluent quality, compact equipment, high volumetric load and less surplus sludge production (Liu et al., 2000Liu, R., Huang, X., Wang, C., Chen, L., Qian, Y., Study on hydraulic characteristics in a submerged membrane bioreactor process. Process Biochem., 36, 249-254 (2000).).One of the newest types of MBRs is the airlift MBR (Futselaar et al., 2007Futselaar, H., Schonewille, H., de Vente, D., Broens, L., NORIT AirLift MBR: side-stream system for municipal waste water treatment. Desalination, 204, 1-7 (2007).). This reactor has lower energy consumption, compared to conventional cross-flow MBR, because it preferably uses air to control membrane fouling and to recirculate liquid (Futselaar et al., 2007Futselaar, H., Schonewille, H., de Vente, D., Broens, L., NORIT AirLift MBR: side-stream system for municipal waste water treatment. Desalination, 204, 1-7 (2007).). Namely, the liquid circulation is induced without a circulation pump. Therefore, in the past years different types of airlift MBR configurations have been designed and investigated. Liu et al. (2000Liu, R., Huang, X., Wang, C., Chen, L., Qian, Y., Study on hydraulic characteristics in a submerged membrane bioreactor process. Process Biochem., 36, 249-254 (2000).) investigated hydrodynamics and its effect on transmembrane pressure in an internal airlift membrane bioreactor. They found that the cross flow velocities increase with an increase of the superficial gas velocity. Jajuee et al. (2006)Jajuee, B., Margaritis, A., Karamanev, D., Bergougnou, M. A., Mass transfer characteristics of a novel three-phase airlift contactor with a semipermeable membrane. Chem. Eng. J., 125, 119-126 (2006). conducted experiments in a 25 dm3 three-phase concentric airlift reactor with a semipermeable membrane. They observed that the mass transfer coefficient was strongly affected by p-xylene, naphthalene and solid loadings. The effect of hydrodynamic conditions and operating modes on the permeate flux in a submerged airlift hollow fiber membrane system was studied by Bérubé and Lei (2006Bérubé, P. R., Lei, E., The effect of hydrodynamic conditions and system configurations on the permeate flux in a submerged hollow fiber membrane system. J. Membr. Sci., 271, 29-37 (2006).). Li et al. (2008Li, Y. Z., He, Y. L., Ohandja, D. G., Ji, J., Li, J. F., Zhou, T., Simultaneous nitrification-denitrification achieved by an innovative internal-loop airlift MBR: Comparative study. Bioresour. Technol., 99, 5867-5872 (2008).) developed an innovative single stage continuously aerated internal-loop membrane airlift bioreactor for simultaneous nitrification and denitrification of synthetic domestic wastewaters. Xu and Yu (2008Yuan, Y.-J., Wei, Z.-J., Wu, Z.-L., Wu, J.-C., Improved Taxol production in suspension cultures of Taxus chinensis var. mairei by in situ extraction combined with precursor feeding and additional carbon source introduction in an airlift loop reactor. Biotechnol. Lett, 23, 1659-1662 (2001).) investigated the hydrodynamics of a novel multi-airlifting membrane bioreactor, constructed with four sintered stainless steel tubular filters serving as the risers and downcomers. In several studies the hydrodynamics in a draft tube airlift reactor with and without membrane have been simulated using computational fluid dynamics software (CFD) (Moraveji et al., 2012Moraveji, M. K., Sajjadi, B., Davarnejad, R., CFD Simulation of hold-up and liquid circulation velocity in a membrane airlift reactor. Theor. Found. Chem. Eng., 46, 266-273 (2012).; Prieske et al., 2008Prieske, H., Drews, A., Kraume, M., Prediction of the circulation velocity in a membrane bioreactor. Desalination, 231, 219-226 (2008).). They concluded that the presence of membrane caused an increase in the gas holdup and decreased the driving force for liquid movement. Mihaľ et al. (2013Bendjaballah, N., Dhaouadi, H., Poncin, S., Midoux, N., Hornut, J. M., Wild, G., Hydrodynamics and flow regimes in external loop airlift reactors. Chem. Eng. Sci., 54, 5211-5221 (1999).) used a hybrid system consisting of an ILAR and membrane module immersed in the downcomer for in situ 2-phenylethanol removal from the fermentation medium.

In all mentioned airlift MBRs the membranes were submerged in the reactor. The key problem of this set-up is that the reactor requires small bubbles for high gas holdup and good gas-liquid mass transfer, while the membrane requires large bubbles to control membrane fouling (Cui et al., 2003Cui, Z. F., Chang, S., Fane, A. G., The use of gas bubbling to enhance membrane processes. J. Membr. Sci., 221, 1-35 (2003).; Futselaar et al., 2007Futselaar, H., Schonewille, H., de Vente, D., Broens, L., NORIT AirLift MBR: side-stream system for municipal waste water treatment. Desalination, 204, 1-7 (2007).). To overcome these problems, external-loop airlift membrane reactors (ELAMR) have been developed. In the ELAMR the riser acts as the reactor, while the separation takes place in the downcomer with an inserted membrane. One of the first configurations of this type of reactor was constructed by Norit for commercial application (Futselaar et al., 2007Futselaar, H., Schonewille, H., de Vente, D., Broens, L., NORIT AirLift MBR: side-stream system for municipal waste water treatment. Desalination, 204, 1-7 (2007).). Fan et al. (2006)Fan, Y., Li, G., Wu, L., Yang, W., Dong, C., Xu, H., Fan, W., Treatment and reuse of toilet wastewater by an airlift external circulation membrane bioreactor. Process Biochem., 41, 1364-1370 (2006). successfully used an H-type recycling pipe external-loop airlift membrane bioreactor to treat and reuse municipal wastewater. Shariati et al. (2010)Shariati, F. P., Mehrnia, M. R., Salmasi, B. M., Heran, M., Wisniewski, C., Sarrafzadeh, M. H., Membrane bioreactor for treatment of pharmaceutical wastewater containing acetaminophen. Desalination, 250, 798-800 (2010). removed acetaminophen as the main pollutant of pharmaceutical wastewater in a rectangular ELAMR. The removal efficiency of acetaminophen was significantly higher for the ELAMR, in comparison to a conventional activated sludge laboratory system.

The hydrodynamics in an ELAMR has significance in the control of membrane fouling and maintenance of the steady operation (Liu et al., 2000Liu, R., Huang, X., Wang, C., Chen, L., Qian, Y., Study on hydraulic characteristics in a submerged membrane bioreactor process. Process Biochem., 36, 249-254 (2000).).The gas holdup and the downcomer liquid velocity are the most important hydrodynamic parameters (Chisti, 1988Chisti, M. Y., Airlift bioreactors. Elsevier Applied Science, London (1988).). It is well known that gas holdup and downcomer liquid velocity in the ELAR depend on parameters such as superficial gas velocity, ratio of the downcomer to riser cross-sectional area, horizontal connector geometries, hydrodynamic resistance to the liquid flow, liquid height in the gas separator, type of gas distributor and physical properties of the liquid phase (Bello et al., 1984Bello, R. A., Robinson, C. W., Moo-Young, M., Liquid circulation and mixing characteristics of airlift contactors. Can. J. Chem. Eng., 62, 573-577 (1984).; Bentifraouine et al., 1997Bentifraouine, C., Xuereb, C., Riba, J.-P., An Experimental Study of the Hydrodynamic Characteristics of External Loop Airlift Contactors. J. Chem. Technol. Biotechnol., 69, 345-349 (1997).; Merchuk and Stein, 1981Merchuk, J. C., Stein, Y., Local hold-up and liquid velocity in air-lift reactors. AIChE J., 27, 377-388 (1981).; Rujiruttanakul and Pavasant, 2011Rujiruttanakul, Y., Pavasant, P., Influence of configuration on the performance of external loop airlift contactors. Chem. Eng. Res. Des., 89, 2254-2261 (2011).). A membrane module inserted in the downcomer represents hydrodynamic resistance to the liquid flow. Higher resistance in the ELAR leads to a decrease in the liquid velocity and an increase in the gas holdup (Bendjaballah et al., 1999Bendjaballah, N., Dhaouadi, H., Poncin, S., Midoux, N., Hornut, J. M., Wild, G., Hydrodynamics and flow regimes in external loop airlift reactors. Chem. Eng. Sci., 54, 5211-5221 (1999).; Cao et al., 2008Cao, C., Dong, S., Geng, Q., Guo, Q., Hydrodynamics and Axial Dispersion in a Gas−Liquid−(Solid) EL-ALR with Different Sparger Designs. Ind. Eng. Chem. Res., 47, 4008-4017 (2008).; Merchuk and Stein, 1981Merchuk, J. C., Stein, Y., Local hold-up and liquid velocity in air-lift reactors. AIChE J., 27, 377-388 (1981).; Pošarac, 1988Pošarac, D. Investigation of hydrodynamics and mass-transfer in a three phase external-loop airlift reactor. PhD Thesis, University of Novi Sad, Novi Sad, Serbia (1988).; Verlaan et al., 1989Verlaan, P., Vos, J.-C., Van T Riet, K., Hydrodynamics of the flow transition from a bubble column to an airlift-loop reactor. J. Chem. Technol. Biotechnol., 45, 109-121 (1989).; Vial et al., 2002Vial, C., Poncin, S., Wild, G., Midoux, N., Experimental and theoretical analysis of the hydrodynamics in the riser of an external loop airlift reactor. Chem. Eng. Sci., 57, 4745-4762 (2002).).

The liquid height in the gas separator has an effect on the gas holdup and the downcomer liquid velocity (Al-Masry, 1999Al-Masry, W. A., Effect of liquid volume in the gas-separator on the hydrodynamics of airlift reactors. J. Chem. Technol. Biotechnol., 74, 931-936 (1999).). The critical level of the liquid in the gas separator prevents the gas bubbles from entering into the downcomer. If the liquid level is below the critical level, the downcomer drags in gas slugs. This situation is preferable in the ELAMR, because of a bigger shear stress in the multiphase flow and minimization of membrane fouling (Bérubé and Lei, 2006Bérubé, P. R., Lei, E., The effect of hydrodynamic conditions and system configurations on the permeate flux in a submerged hollow fiber membrane system. J. Membr. Sci., 271, 29-37 (2006).; Böhm et al., 2012Böhm, L., Drews, A., Prieske, H., Bérubé, P. R., Kraume, M., The importance of fluid dynamics for MBR fouling mitigation. Bioresour. Technol., 122, 50-61 (2012).; Ratkovich et al., 2009Ratkovich, N., Chan, C. C. V., Berube, P. R., Nopens, I., Experimental study and CFD modelling of a two-phase slug flow for an airlift tubular membrane. Chem. Eng. Sci., 64, 3576-3584 (2009).).

Physical properties of liquids also influence both the gas holdup and the liquid velocity. Alcohol solutions were used as a model liquid phase of non-coalescing organic mixtures in coal liquefaction and bioreactors (Kelkar et al., 1983Kelkar, B. G., Godbole, S. P., Honath, M. F., Shah, Y. T., Carr, N. L., Deckwer, W. D., Effect of addition of alcohols on gas holdup and backmixing in bubble columns. AIChE J., 29, 361-369 (1983).). Addition of 2-propanol to an air-water system induces a behavior similar to the fermentation media used in aerobic bioprocesses (McClure et al., 2014McClure, D. D., Deligny, J., Kavanagh, J. M., Fletcher, D. F., Barton, G. W., Impact of Surfactant Chemistry on Bubble Column Systems. Chem. Eng. Technol., 37, 652-658 (2014).). In bubble columns and airlift bioreactors inorganic salts, sugars and metabolic products, such as alcohols and organic acids, were present in significant quantities in the culture medium (Jamialahmadi and Müller-Steinhagen, 1992Jamialahmadi, M., Müller-Steinhagen, H., Effect of alcohol, organic acid and potassium chloride concentration on bubble size, bubble rise velocity and gas hold-up in bubble columns. Chem. Eng. J., 50, 47-56 (1992).; Schügerl et al., 1977Schügerl, K., Lücke, J., Oels, U., Bubble column bioreactors. In Advances in Biochemical Engineering , Volume 7, Springer Berlin Heidelberg: 1977; Vol. 7, pp 1-84.). A small amount of alcohol (below 1%) remarkably decreased the surface tension of the aqueous solution, thus changing hydrodynamic properties of the airlift reactors. The surface tension of the dilute alcohol solution was the only physical property that differs from water (Freitas and Teixeira, 1998Freitas, C., Teixeira, J. A., Effect of liquid-phase surface tension on hydrodynamics of a three-phase airlift reactor with an enlarged degassing zone. Bioprocess. Eng., 19, 451-457 (1998).). In ELAR the presence of alcohols caused an increase in the gas holdup (Al-Masry and Dukkan, 1997Al-Masry, W. A., Dukkan, A. R., The role of gas disengagement and surface active agents on hydrodynamic and mass transfer characteristics of airlift reactors. Chem. Eng. J., 65, 263-271 (1997).; Gharib et al., 2013Gharib, J., Keshavarz Moraveji, M., Davarnejad, R., Malool, M. E., Hydrodynamics and mass transfer study of aliphatic alcohols in airlift reactors. Chem. Eng. Res. Des., 91, 925-932 (2013).; Miyahara and Nagatani, 2009Miyahara, T., Nagatani, N., Influence of Alcohol Addition on Liquid-Phase Volumetric Mass Transfer Coefficient in an External-Loop Airlift Reactor with a Porous Plate. J. Chem. Eng. Jpn., 42, 713-719 (2009).; Pošarac, 1988Pošarac, D. Investigation of hydrodynamics and mass-transfer in a three phase external-loop airlift reactor. PhD Thesis, University of Novi Sad, Novi Sad, Serbia (1988).; Weiland and Onken, 1981Weiland, P., Onken, U., Fluid Dynamics and Mass Transfer in an Airlift Fermenter with External Loop. Ger.Chem.Eng., 4, 42-50 (1981).). Also, Weiland and Onken (1981)Weiland, P., Onken, U., Fluid Dynamics and Mass Transfer in an Airlift Fermenter with External Loop. Ger.Chem.Eng., 4, 42-50 (1981). and Pošarac (1988)Pošarac, D. Investigation of hydrodynamics and mass-transfer in a three phase external-loop airlift reactor. PhD Thesis, University of Novi Sad, Novi Sad, Serbia (1988). reported that the addition of alcohol in an ELAR increased the liquid velocity in the downcomer. On the contrary, Al-Masry and Dukkan (1997)Al-Masry, W. A., Dukkan, A. R., The role of gas disengagement and surface active agents on hydrodynamic and mass transfer characteristics of airlift reactors. Chem. Eng. J., 65, 263-271 (1997). and Miyahara and Nagatani (2009)Miyahara, T., Nagatani, N., Influence of Alcohol Addition on Liquid-Phase Volumetric Mass Transfer Coefficient in an External-Loop Airlift Reactor with a Porous Plate. J. Chem. Eng. Jpn., 42, 713-719 (2009). found a marginal effect of alcohol solutions on the liquid velocity.

Gas distributor design has a major effect on the initial bubble size and, hence, on the hydrodynamics of the airlift reactor (Bendjaballah et al., 1999Bendjaballah, N., Dhaouadi, H., Poncin, S., Midoux, N., Hornut, J. M., Wild, G., Hydrodynamics and flow regimes in external loop airlift reactors. Chem. Eng. Sci., 54, 5211-5221 (1999).; Cao et al., 2008Cao, C., Dong, S., Geng, Q., Guo, Q., Hydrodynamics and Axial Dispersion in a Gas−Liquid−(Solid) EL-ALR with Different Sparger Designs. Ind. Eng. Chem. Res., 47, 4008-4017 (2008).; Lin et al., 2004Lin, J., Han, M., Wang, T., Zhang, T., Wang, J., Jin, Y., Influence of the gas distributor on the local hydrodynamic behavior of an external loop airlift reactor. Chem. Eng. J., 102, 51-59 (2004).; Snape et al., 1995Snape, J. B., Zahradník, J., Fialová, M., Thomas, N. H., Liquid-phase properties and sparger design effects in an external-loop airlift reactor. Chem. Eng. Sci., 50, 3175-3186 (1995).; Vial et al., 2000Vial, C., Camarasa, E., Poncin, S., Wild, G., Midoux, N., Bouillard, J., Study of hydrodynamic behaviour in bubble columns and external loop airlift reactors through analysis of pressure fluctuations. Chem. Eng. Sci., 55, 2957-2973 (2000).). Cao et al. (2008)Cao, C., Dong, S., Geng, Q., Guo, Q., Hydrodynamics and Axial Dispersion in a Gas−Liquid−(Solid) EL-ALR with Different Sparger Designs. Ind. Eng. Chem. Res., 47, 4008-4017 (2008). performed the most comprehensive study about the influence of the gas distributor on the gas holdup and liquid velocity in the ELAR. They observed that the gas distributor had a noticeable effect on the gas holdup up to the superficial gas velocity of 0.25 m/s. At higher gas inputs, the distributor’s influence was negligible. However, the effect of the gas distributor on the liquid velocity was evident in the range of gas velocities from 0.025 m/s to 0.045 m/s.

In this paper, the hydrodynamics of an ELAMR with a short multichannel ceramic membrane inserted at the bottom of the downcomer was investigated. Such a configuration enables high hydrostatic head pressure, which decreases the power necessary for permeate removal. Furthermore, by changing the liquid level in the gas separator, the set of experiments were done with gas slug entrainment in the downcomer. This operation mode is useful for preventing membrane fouling. Also, the influences of the gas distributor’s geometry and the addition of aliphatic alcohols on the riser gas holdup and downcomer liquid velocity of the ELAMR were studied. Also, this reactor has been designed for possible use as a high efficiency equipment for removal of organic and inorganic pollutants from wastewater.

MATERIALS AND METHODS

Experimental setup

A schematic illustration of the setup is shown in Figure 1. The cylindrical external-loop airlift reactor made of Plexiglas consisted of a riser (54 mm i.d. and 2000 mm in height), downcomer (25 mm i.d. and 1950 mm in height) and rectangular gas separator (400×310×300 mm). The distance between the riser and the downcomer was 100 mm. When the ELAMR was operated without gas bubbles in the downcomer (mode A), the unaerated liquid level in the gas separator was 4 cm. Lowering the liquid level to 3 cm, gas bubbles were dragged into the downcomer (mode B). The air, sparged into the riser, was used as the gas phase. Three different gas distributors were tested: single orifice (4 mm i.d.), perforated plate (7 holes of 1 mm i.d., triangular pitch) and sinter plate (100-160 µm, average pore size 115 µm, porosity 8%). Porosity and average pore size of the sinter plate were obtained with a porosimeter (Porosimeter 2000 with Macropore Unit 120). The gas flow rates were controlled and measured by an electronic mass flow controller (model Bronkhorst High Tech F 201AV). The superficial gas velocity, based on the riser cross-sectional area, was varied in the range 0.02 to 0.22 m/s for mode A. In mode B, the entrainment of gas bubbles in the downcomer started at UG = 0.15 m/s, so the investigated range of UG was 0.15 to 0.22 m/s. Two eDAQ (Australia) conductivity isoPods with miniature dip-in conductivity electrodes were used to determine downcomer liquid velocity. The tubular ceramic membrane (ZrO2/TiO2, Novasep, France) 20 cm in length and 2.3 cm in diameter, with 7 channels (6 mm i.d.), was installed in the downcomer. The filtration was disabled in the membrane module since our aim was to investigate only the influence of the membrane acting as additional resistance on the hydrodynamics in the reactor. Also, it was considered that the permeate flux would not affect the hydrodynamics (Böhm et al., 2012Böhm, L., Drews, A., Prieske, H., Bérubé, P. R., Kraume, M., The importance of fluid dynamics for MBR fouling mitigation. Bioresour. Technol., 122, 50-61 (2012).).

Figure 1
Experimental setup: 1-riser, 2-downcomer, 3-gas separator, 4-piezometric tubes, 5-gas distributor, 6-manometer, 7-mass flow controller, 8-air compressor, 9-conductivity electrode, 10-conductivity isopod, 11-membrane module, 12-membrane.

Gas-liquid systems

Tap water and dilute alcohol solutions (0.046 wt% ethanol, 0.011 wt% n-butanol and 0.0051 wt% n-hexanol) were used as the liquid phase. Added amounts of each alcohol correspond to their critical concentration reported by Keitel (1978)Keitel, G. Untersuchungen zum Stoffaustausch in Gas-Flüssig-Dispersionen in Rührschlaufenreaktor und Blasensäule. PhD Thesis, Universität Dortmund, Dortmund, Germany (1978).. Increasing the alcohol concentration above the upper limiting concentration value, only enhances the liquid phase frothing and bubble coalescence (Camarasa et al., 1999Camarasa, E., Vial, C., Poncin, S., Wild, G., Midoux, N., Bouillard, J., Influence of coalescence behaviour of the liquid and of gas sparging on hydrodynamics and bubble characteristics in a bubble column. Chem. Eng. Process., 38, 329-344 (1999).; Freitas and Teixeira, 1998Freitas, C., Teixeira, J. A., Effect of liquid-phase surface tension on hydrodynamics of a three-phase airlift reactor with an enlarged degassing zone. Bioprocess. Eng., 19, 451-457 (1998).). Surface tensions of liquid phases and the surface tension gradient (-dσ/dCA) data were taken from Šijački et al. (2011)Šijački, I. M., Tokić, M. S., Kojić, P. S., Petrović, D. L., Tekić, M. N., Djurić, M. S., Milovančev, S. S., Sparger Type Influence on the Hydrodynamics of the Draft Tube Airlift Reactor with Diluted Alcohol Solutions. Ind. Eng. Chem. Res., 50, 3580-3591 (2011)..

Measurement of hydrodynamic characteristics

Gas holdup

The gas holdup values in the riser (εGR) and the downcomer (εGD) were obtained by measuring the pressures at the bottom and the top of the riser and downcomer using piezometric tubes, and calculated from the equation:

ε G = z H (1)

In order to reduce the liquid surface fluctuations in the piezometric tubes, capillaries (50 mm in length and 0.7 mm i.d.) were inserted at the entrance of the piezometric tubes. Therefore, the relative average error of these measurements was reduced to max ±2%.This experimental method was adopted from Zahradník et al. (1974)Zahradník, J., Kaštánek, F., Rylek, M., Porosity of the heterogeneous bed and liquid circulation in multistage bubble-type column reactors. Collect. Czech. Chem. Commun., 39, 1403-1418 (1974).. Also, it was visually observed that the gas bubbles did not enter into the piezometric tube due to frothing in the riser and the downcomer.

Liquid velocity in the downcomer

The liquid velocity in the downcomer (WLD) was determined by the tracer response method. Two conductivity probes were placed in the downcomer section at a distance (L) of 1.4 m from each other. A volume of 25 cm3 of 4 M NaCl, used as a tracer, was instantaneously injected 15 cm above the top electrode. Liquid velocity in the downcomer was calculated from the measured time interval between the tracer peaks from the two conductivity probes and the known vertical distance between them, by the following equation:

W L D = L t 2 - t 1 (2)

The cross flow velocity inside the membrane could be calculated using the continuity equations by knowing WLD. Signals were recorded at a frequency of 0.1 s. For each value of a gas flow rate two measurements were performed and the average value of WLD was calculated. The relative average error of this method was ±3%.

Friction coefficient

To quantify the hydrodynamic resistance of the membrane itself, it was necessary to calculate the overall friction coefficient (Kf) in both reactors. The Kf was derived according to Verlaan (1987)Verlaan, P. Modelling and characterization of an airlift-loop bioreactor. PhD Thesis, Wageningen University, Wageningen, Netherlands (1987). by plotting the square of the measured superficial liquid velocity as a function of the difference between gas holdups in the riser and the downcomer:

W L D 2 = 2 g H K f ( ε G R - ε G D ) (3)

The commercial software pipe flow expert was used to calculate Kf(calc). Kf(calc) was estimated according to Garcia-Calvo (1992)Garcia-Calvo, E., Fluid dynamics of airlift reactors: Two-phase friction factors. AIChE J., 38, 1662-1666 (1992). and Milivojević (2011)Milivojević, M. Brzina tečnosti u dvofaznim i trofaznim pneumatskim reaktorima sa spoljašnjom cirkulacijom. PhD Thesis, University of Belgrade, Belgrade, Serbia (2011). as the sum of specific fittings in each reactor separately. In the estimation of Kf(calc) the fitting that represented the resistance of the membrane was taken as a partially open valve for the ELAMR and as a fully open valve for the ELAR. The assumption was made that the type of distributor and alcohol do not influence Kf(calc). A good agreement between calculated and experimental Kf values, with maximal relative error of 15%, was achieved.

RESULTS AND DISCUSSION

Hydrodynamic regimes

It is well known that the hydrodynamics of a bubble column (riser of the ELAR) in the gas-liquid co-current operation mode is characterized by different flow patterns depending on the gas flow rate: homogenous (bubble flow), transition, heterogeneous (churn-turbulent flow) regimes and slug flow(Govier, 1972Govier, G. W. A. K., The flow of complex mixtures in pipes. Van Nostrand Reinhold Co., New York (1972).; Hatch, 1975Hatch, R. T., Fermenter design, in: Single Cell Protein II (S. R. Tannenbaum and D. I. C. Wang, eds.). MIT Press, Cambridge (1975); p 46-68.; Wallis, 1969Wallis, G. B., One-dimensional two-phase flow. McGraw-Hill, New York (1969).).

In both reactors, ELAR and ELAMR, for operating mode A the existence of all regimes was confirmed (Figures 2a-c and 3a-c). The transition between regimes was identified by the change of the slope of the gas holdup vs. the gas velocity curves. A more detailed analysis of hydrodynamic regimes and transition superficial gas velocities was described in our earlier study (Kojić et al., 2015Kojić, P. S., Tokić, M. S., Šijački, I. M., Lukić, N. L., Petrović, D. L., Jovičević, D. Z., Popović, S. S., Influence of the Sparger Type and Added Alcohol on the Gas Holdup of an External-Loop Airlift Reactor. Chem. Eng. Technol., 38, 701-708 (2015).).The transition points in the ELAMR appeared at a slightly lower UG compared to the ELAR. This is because the larger hydrodynamic resistance in the ELAMR caused lower WLD that induced more intensive bubble coalescence. Joshi et al. (1990)Joshi, J. B., Ranade, V. V., Gharat, S. D., Lele, S. S., Sparged loop reactors. Can. J. Chem. Eng., 68, 705-741 (1990). and Bendjaballah et al. (1999)Bendjaballah, N., Dhaouadi, H., Poncin, S., Midoux, N., Hornut, J. M., Wild, G., Hydrodynamics and flow regimes in external loop airlift reactors. Chem. Eng. Sci., 54, 5211-5221 (1999). also noted that the resistance in the ELAR, and thereby WLD, and the geometry of the downcomer influenced the regime transitions. On the other hand, Vial et al. (2002)Vial, C., Poncin, S., Wild, G., Midoux, N., Experimental and theoretical analysis of the hydrodynamics in the riser of an external loop airlift reactor. Chem. Eng. Sci., 57, 4745-4762 (2002). reported that the hydrodynamic resistance had a minor influence on the regime transitions. As can be seen in Figures 2 and 3, added alcohol had a noticeable influence on the transition between regimes in both ELAR and ELAMR for the sinter plate and perforated plate. But, for a single orifice this transition was the same for all alcohol solutions. The presence of alcohols delayed the heterogeneous regime, due to their coalescence inhibiting nature (Pošarac, 1988Pošarac, D. Investigation of hydrodynamics and mass-transfer in a three phase external-loop airlift reactor. PhD Thesis, University of Novi Sad, Novi Sad, Serbia (1988).; Weiland and Onken, 1981Weiland, P., Onken, U., Fluid Dynamics and Mass Transfer in an Airlift Fermenter with External Loop. Ger.Chem.Eng., 4, 42-50 (1981).). The effect of the gas distributor type on the regime transition in the ELAR and ELAMR was also observed. When a single orifice was used, the absence of bubble flow was noted even at the lowest UG. Both reactors operated instantly in the transition regime, that is in agreement with the observations of Bendjaballah et al. (1999)Bendjaballah, N., Dhaouadi, H., Poncin, S., Midoux, N., Hornut, J. M., Wild, G., Hydrodynamics and flow regimes in external loop airlift reactors. Chem. Eng. Sci., 54, 5211-5221 (1999). and Vial et al. (2001)Vial, C., Poncin, S., Wild, G., Midoux, N., A simple method for regime identification and flow characterisation in bubble columns and airlift reactors. Chem. Eng. Process., 40, 135-151 (2001).. In mode B, at UG ≥ 0.15 m/s, gas slugs started to be dragged into the downcomer. The length of the slugs was ~5 cm in water and ~ 3 cm in alcohol solutions. When the alcohols were added smaller number of gas slugs were created in the downcomer, in comparison to water. Mode B is significant for the mitigation of membrane fouling. Large bubbles or slugs are more beneficial as they have larger wake regions, create stronger secondary flows and are more effective in promoting local mixing than smaller bubbles. Also, for a fixed bubble volume, an increase in bubbling frequency means more falling films and bubble wakes per unit time (Cui et al., 2003Cui, Z. F., Chang, S., Fane, A. G., The use of gas bubbling to enhance membrane processes. J. Membr. Sci., 221, 1-35 (2003).). According to Ratkovich et al. (2009)Ratkovich, N., Chan, C. C. V., Berube, P. R., Nopens, I., Experimental study and CFD modelling of a two-phase slug flow for an airlift tubular membrane. Chem. Eng. Sci., 64, 3576-3584 (2009). gas slugs increase the shear stress across the membrane and prevent membrane fouling.

Gas Holdup

Influence of the superficial gas velocity on the gas holdup

The influence of the superficial gas velocity on the riser gas holdup is presented in Figures 2 and 3 for the ELAR and ELAMR, respectively. For all the gas-liquid systems and both reactors (ELAR and ELAMR), the gas holdup increased with an increase in the superficial gas velocity, and it was higher in the ELAMR than in the ELAR. The increase in εGR was the highest for the single orifice (about 16%), while for the perforated plate and the sinter plate it was lower (about 8%). When the ELAMR was operated in mode B, εGD was up to 0.03, while εGR was unchanged. εGD values were about 10 times lower, compared to εGR (Figure 3).

To quantify the influence of the membrane as the resistance to the flow in the ELAMR, the overall friction coefficient was calculated from equation 3 by fitting the experimental data. The insertion of the membrane in the downcomer of our ELAR increased the overall friction coefficient by 90% while εGR was slightly increased. Table 1 illustrates that increasing the resistance in the downcomer of the ELAR results in an increase in εGR. This effect is the most pronounced when the cross section of the downcomer is reduced by more than 50%. Furthermore, this contributed to an increase of Kf by more than 100%. Reduced liquid velocity decreased the rise velocity of the bubbles and consequently increased εGR.

Table 1
List of the influence of hydrodynamic resistance on the riser gas holdup and downcomer liquid velocity in the external-loop airlift reactors.

Influence of added alcohol on the gas holdup

The effect of alcohols on the gas holdup in both reactors, without and with membrane, was the same (Figures 2 and 3). The addition of alcohols led to an increase in εGR, in comparison to water for all distributors, but more intensively for the sinter plate (SP) and perforated plate (PP). At UG ≤ 0.05 m/s only slight differences in εGR were observed, regardless of the geometry of the distributor and type of alcohol. In the bubble flow no coalescence was observed, even in the water. However, in the transition and churn-turbulent flow εGR was about 22, 17.5, and 10% higher for ethanol, n-butanol and n-hexanol solutions, respectively for SP and PP, in comparison to water. At high gas velocities, corresponding to the transition and heterogeneous flow, coalescence was much stronger with frequent bubble collisions. The inertial forces dominated over the surface forces and the effect of coalescence inhibition due to added alcohol was more pronounced. The effect of added alcohol on εGR for the single orifice (SO) (Figures 2 and 3) was lower than for the PP and the SP, so in the transition and the churn-turbulent regime alcohol solutions had only about a 7% higher εGR compared to water. The SO produced large bubbles and the gas-liquid system was instantly in the transition regime; therefore, it was difficult for alcohols to form monolayer around the bubbles and prevent coalescence. When the ELAMR was operated in mode B, added alcohols slightly decreased εGD, compared to water.

Influence of distributor type on the gas holdup

The effect of gas distributor: SO, PP and SP is shown in Figures 2 and 3. It can be clearly seen, that the SP for all used gas-liquid systems produced the highest εGR. It is obvious that the SP and PP are more efficient distributors than the SO. The membrane module reduced the effect of the SP and PP on εGR in comparison to SO (by 25% to 50% depending on the gas-liquid systems). The efficiency of both the SP and PP started to decrease at lower UG in the ELAMR, because of the fact that the churn-turbulent flow appeared earlier than in the ELAR. The SP was more efficient than the PP (UG ≤ 0.09 m/s) until the inertial forces started to be dominant. Therefore, at higher UG differences in εGR were less than 3%. Alcohols improved the efficiency of the SP and PP compared to SO (which remained the least efficient distributor), especially in churn-turbulent flow. For instance, in the region of surface forces domination εGR was higher up to 40%. However, at UG> 0.09 m/s, when coalescence was intensive, εGR produced by the SP and PP was only 4% higher for water and about 20% for alcohol solutions. When the ELAMR was operated in mode B, the SO produced slightly smaller εGD compared to the SP and PP.

Figure 2
Influence of gas distributor type and added alcohol on the gas holdup in the ELAR. Hydrodynamic regime legend: open symbols, homogeneous regime; half solid symbols, transition regime; solid symbols, heterogeneous regime and slug flow.

Figure 3
Influence of gas distributor type, operation mode and added alcohol on the gas holdup in the ELAMR. Hydrodynamic regime legend: open symbols, homogeneous regime; half solid symbols, transition regime; solid symbols, heterogeneous regime and slug flow.

Downcomer liquid velocity

Influence of superficial gas velocity on the downcomer liquid velocity

Figures 4 and 5 present the effect of UG on the liquid velocity in the downcomer for both operating modes and all distributors in the ELAR and the ELAMR, respectively. For all the gas-liquid systems WLD increased with increasing UG. Insertion of a membrane module in the downcomer of the ELAR increased frictional energy losses and led to a decrease of WLD in the range of 36-49% for both SP and PP, while for SO the decrease was in the range of 19-27%, depending on the gas-liquid system. In order to reduce membrane fouling, it is very important to know the liquid velocity through the channel of the membrane (cross-flow velocity). A continuity equation was used for its estimation. This velocity was about 2.5 times higher than WLD. Our ELAMR at higher UG can achieve the cross flow velocity for microfiltration and ultrafiltration (2-3 m/s), which was proposed by Rossignol et al. (1999)Rossignol, N., Vandanjon, L., Jaouen, P., Quéméneur, F., Membrane technology for the continuous separation microalgae/culture medium: compared performances of cross-flow microfiltration and ultrafiltration. Aquacult. Eng., 20, 191-208 (1999). and Choi et al. (2005)Choi, H., Zhang, K., Dionysiou, D. D., Oerther, D. B., Sorial, G. A., Influence of cross-flow velocity on membrane performance during filtration of biological suspension. J. Membr. Sci., 248, 189-199 (2005).. When the ELAMR was operated in mode B, the slugs in the downcomer reduced WLD by 5-15%, depending on the gas-liquid system and type of gas distributor. The velocities of the gas slugs in the downcomer were 35-46 cm/s, in the range of applied UG.

Influence of added alcohol on the downcomer liquid velocity

The results in Figures 4 and 5 showed that the downcomer liquid velocity in the majority of the investigated gas-liquid systems and gas distributors changed with added alcohol. The reason for this is a higher driving force for the liquid circulation. Namely, the driving force is proportional to the hydrostatic pressure difference between the two vertical columns in the ELAR. Added alcohol increased the riser gas holdup, while the downcomer gas holdup remained unchanged because there was no bubble entrainment in the downcomer (mode A). Therefore, WLD was highest in the ethanol and n-butanol solutions for all distributors. The membrane, inserted in the downcomer of the ELAR, diminished the effect of alcohol on WLD. For instance, added alcohols increased WLD from 10% to 15.5% in comparison to water in the ELAR with SP, while in the ELAMR the increase was only 3.7% to 10%, depending on the alcohol solution. In both operating modes, the effect of added alcohols on WLD was the same.

Influence of gas distributor on the downcomer liquid velocity

Figures 4 and 5 show the effect of UG and gas distributor type on WLD. In the ELAR, SP and PP were equally efficient. They gave higher WLD (12-20% depending on the gas-liquid system) than SO. In the ELAMR the effect of gas distributor type was reduced because of more intensive coalescence that appeared at lower UG compared to the ELAR. In the region of surface forces domination (bubble and transition flow) in the ELAMR the highest WLD were achieved using SP, while the lowest ones were with SO. However, in churn-turbulent flow (UG> 0.09 m/s), when inertial forces dominated, WLD tended to be equal for all three distributors, as the differences in the WLD between distributors were less than 3%. In both operating modes, the same influence of the gas distributor on WLD was noticed.

Figure 4
Influence of gas distributor type and added alcohol on the downcomer liquid velocity in the ELAR. Hydrodynamic regime legend: open symbols, homogeneous regime; half solid symbols, transition regime; solid symbols, heterogeneous regime and slug flow.

Figure 5
Influence of gas distributor type and added alcohol on the downcomer liquid velocity in the ELAMR. Hydrodynamic regime legend: open symbols, homogeneous regime; half solid symbols, transition regime; solid symbols, heterogeneous regime and slug flow; crossed open symbols, operation mode B.

CORRELATIONS

Empirical power law correlations

According to experimental results in both operating modes examined the general comment is that εGR and WLD in the ELAR and ELAMR depend on the superficial gas velocity, distributor type, surface properties of the gas-liquid system and overall friction coefficient. Distributor type has a strong influence on the primary bubble dispersion and therefore on the global hydrodynamics. The main parameter that describes the gas distributor, i.e., initial bubble size, is the orifice diameter (do) (Šijački et al., 2011Šijački, I. M., Tokić, M. S., Kojić, P. S., Petrović, D. L., Tekić, M. N., Djurić, M. S., Milovančev, S. S., Sparger Type Influence on the Hydrodynamics of the Draft Tube Airlift Reactor with Diluted Alcohol Solutions. Ind. Eng. Chem. Res., 50, 3580-3591 (2011).). The influence of alcohols on εGR and subsequently on WLD, should be linked to changes in the surface tension gradient, i.e., to the coalescence suppression strength of individual alcohols (Albijanić et al., 2007Albijanić, B., Havran, V., Petrović, D. L., Đurić, M., Tekić, M. N., Hydrodynamics and mass transfer in a draft tube airlift reactor with dilute alcohol solutions. AlChE J., 53, 2897-2904 (2007).; Camarasa et al., 1999Camarasa, E., Vial, C., Poncin, S., Wild, G., Midoux, N., Bouillard, J., Influence of coalescence behaviour of the liquid and of gas sparging on hydrodynamics and bubble characteristics in a bubble column. Chem. Eng. Process., 38, 329-344 (1999).). The effect of hydrodynamic resistance on the liquid velocity in the downcomer depends on the friction coefficient (Kf) in the ELAR and ELAMR, in both operating modes. Therefore, the general form of the correlations we applied was:

(y) calc = AU G B ( 1+ ( - dC A ) ) C d o D K f(calc) E (4)

The correlations were fitted for the following ranges of independent variables: 0.022 < UG< 0.218 m/s, 0.027 < (dσ/dCA) < 1.985 mNm2/mol, 0.115 < do< 4 mm, 9 < Kf(calc)< 30. The coefficients were determined by minimizing the least-square sums over nonlinear correlations. These calculations were performed in Mathcad software. Table 2 contains the values of estimated parameters in the proposed correlations (eq. 4) and the goodness of fit. The parity plot of calculated values versus experimental ones is presented in Figure 6.

Table 2
Values of correlations parameters for the riser gas holdup and downcomer liquid velocity and fit statistics

Figure 6
Comparison of experimental and predicted values for the riser gas holdup and downcomer liquid velocity using the proposed correlations (4).

Neural network simulations

In this study the feed-forward artificial neural network model (ANN) was used, beside the previously proposed empirical power law correlations, to predict εGR and WLD. Al-Masry (2006)Al-Masry, W. A., Analysis of Hydrodynamics of External Loop Circulating Bubble Columns with Open Channel Gas Separators Using Neural Networks. Chem. Eng. Res. Des., 84, 483-486 (2006)., and Chen et al. (2013)Chen, Z., Liu, H., Zhang, H., Ying, W., Fang, D., Oxygen mass transfer coefficient in bubble column slurry reactor with ultrafine suspended particles and neural network prediction. Can. J. Chem. Eng., 91, 532-541 (2013). pointed out that an ANN reasonably predicted experimental values for liquid velocities, gas holdup and volumetric gas-liquid mass transfer coefficient in the ELAR. Neural network inputs were UG, -dσ/dCA, do and Kf(calc), the ranges of network inputs were the same as in the correlations (eq. 4) mentioned above. The number of neurons was 20 and 2 in the hidden and output layer, respectively. The network was trained with the Levenberg-Marquard back propagations algorithm. The transfer function was the log-sigmoid. All 450 data points were used to train and develop the ANN: 70% of the data points for training, 15% of the data for validations and 15% of the data for the testing of the process. Figure 7 shows the parity plot of the experimental and predicted values for εGR and WLD using the ANN on the whole database. The statistical analysis of prediction with the ANN is shown in Table 2. Based on Table 2, the ANN predicts the experimental results more accurately than power law correlations.

Figure 7
Comparison of experimental and predicted values for the riser gas holdup and downcomer liquid velocity using an artificial neural network.

CONCLUSION

The following conclusions could be drawn from the present study for both external-loop airlift reactors, with and without inserted membrane in the downcomer:

The insertion of the membrane in the downcomer caused a rise of the overall friction coefficient by 90% and a decrease of the downcomer liquid velocity up to 50%. However, the cross-flow velocity remained in the recommended range of velocities for ultrafiltration and microfiltration given in the literature. Decreased downcomer liquid velocity resulted in up to a 16% increase of the gas holdup, while the regime transitions shifted toward the lower superficial gas velocities. The presence of bubbles in the downcomer decreased the downcomer liquid velocity up to 15% and resulted in a downcomer gas holdup up to 0.03. An artificial neural network gave a good prediction of riser gas holdup and downcomer liquid velocity for both reactor configurations studied. Future research should be oriented toward enabling filtration within an external-loop airlift reactor with membrane in the downcomer.

    NOMENCLATURE
  • do  diameter of orifice (mm)
  • g  gravitational acceleration rate (m/s2)
  • H  column height (m)
  • Kf  friction coefficient
  • L  distance between conductivity electrodes (m)
  • t  time (s)
  • UG  superficial gas velocity (m/s)
  • WLD  downcomer liquid velocity (m/s)
  • .   Greek letters
  • ΔH  distance between two measuring points (mm)
  • Δz  pressure drop between two measuring points (mm)
  • δ  mean relative error (%)
  • εGR  riser gas holdup
  • εGD  downcomer gas holdup
  • -dσ/dCA  surface tension gradient (N m2/mol)
  • .   Abbreviations
  • ANN  artificial neural network
  • ELAR  external-loop airlift reactor
  • ELAMR  external-loop airlift membrane reactor
  • MBR  membrane bioreactor

ACKNOWLEDGMENT

This research was supported by the Ministry of Education and Science of the Republic of Serbia. (Project No. 172025).

REFERENCES

  • Al-Masry, W. A., Analysis of Hydrodynamics of External Loop Circulating Bubble Columns with Open Channel Gas Separators Using Neural Networks. Chem. Eng. Res. Des., 84, 483-486 (2006).
  • Al-Masry, W. A., Effect of liquid volume in the gas-separator on the hydrodynamics of airlift reactors. J. Chem. Technol. Biotechnol., 74, 931-936 (1999).
  • Al-Masry, W. A., Dukkan, A. R., The role of gas disengagement and surface active agents on hydrodynamic and mass transfer characteristics of airlift reactors. Chem. Eng. J., 65, 263-271 (1997).
  • Albijanić, B., Havran, V., Petrović, D. L., Đurić, M., Tekić, M. N., Hydrodynamics and mass transfer in a draft tube airlift reactor with dilute alcohol solutions. AlChE J., 53, 2897-2904 (2007).
  • Bello, R. A., Robinson, C. W., Moo-Young, M., Liquid circulation and mixing characteristics of airlift contactors. Can. J. Chem. Eng., 62, 573-577 (1984).
  • Bendjaballah, N., Dhaouadi, H., Poncin, S., Midoux, N., Hornut, J. M., Wild, G., Hydrodynamics and flow regimes in external loop airlift reactors. Chem. Eng. Sci., 54, 5211-5221 (1999).
  • Bentifraouine, C., Xuereb, C., Riba, J.-P., An Experimental Study of the Hydrodynamic Characteristics of External Loop Airlift Contactors. J. Chem. Technol. Biotechnol., 69, 345-349 (1997).
  • Bérubé, P. R., Lei, E., The effect of hydrodynamic conditions and system configurations on the permeate flux in a submerged hollow fiber membrane system. J. Membr. Sci., 271, 29-37 (2006).
  • Böhm, L., Drews, A., Prieske, H., Bérubé, P. R., Kraume, M., The importance of fluid dynamics for MBR fouling mitigation. Bioresour. Technol., 122, 50-61 (2012).
  • Camarasa, E., Vial, C., Poncin, S., Wild, G., Midoux, N., Bouillard, J., Influence of coalescence behaviour of the liquid and of gas sparging on hydrodynamics and bubble characteristics in a bubble column. Chem. Eng. Process., 38, 329-344 (1999).
  • Cao, C., Dong, S., Geng, Q., Guo, Q., Hydrodynamics and Axial Dispersion in a Gas−Liquid−(Solid) EL-ALR with Different Sparger Designs. Ind. Eng. Chem. Res., 47, 4008-4017 (2008).
  • Carstensen, F., Apel, A., Wessling, M., In situ product recovery: Submerged membranes vs. external loop membranes. J. Membr. Sci., 394-395, 1-36 (2012).
  • Chen, Z., Liu, H., Zhang, H., Ying, W., Fang, D., Oxygen mass transfer coefficient in bubble column slurry reactor with ultrafine suspended particles and neural network prediction. Can. J. Chem. Eng., 91, 532-541 (2013).
  • Chisti, M. Y., Airlift bioreactors. Elsevier Applied Science, London (1988).
  • Choi, H., Zhang, K., Dionysiou, D. D., Oerther, D. B., Sorial, G. A., Influence of cross-flow velocity on membrane performance during filtration of biological suspension. J. Membr. Sci., 248, 189-199 (2005).
  • Cui, Z. F., Chang, S., Fane, A. G., The use of gas bubbling to enhance membrane processes. J. Membr. Sci., 221, 1-35 (2003).
  • Fan, Y., Li, G., Wu, L., Yang, W., Dong, C., Xu, H., Fan, W., Treatment and reuse of toilet wastewater by an airlift external circulation membrane bioreactor. Process Biochem., 41, 1364-1370 (2006).
  • Freitas, C., Teixeira, J. A., Effect of liquid-phase surface tension on hydrodynamics of a three-phase airlift reactor with an enlarged degassing zone. Bioprocess. Eng., 19, 451-457 (1998).
  • Futselaar, H., Schonewille, H., de Vente, D., Broens, L., NORIT AirLift MBR: side-stream system for municipal waste water treatment. Desalination, 204, 1-7 (2007).
  • Garcia-Calvo, E., Fluid dynamics of airlift reactors: Two-phase friction factors. AIChE J., 38, 1662-1666 (1992).
  • Gharib, J., Keshavarz Moraveji, M., Davarnejad, R., Malool, M. E., Hydrodynamics and mass transfer study of aliphatic alcohols in airlift reactors. Chem. Eng. Res. Des., 91, 925-932 (2013).
  • Gianetto, A., Ruggeri, B., Specchia, V., Sassi, G., Forna, R., Continuous extraction loop reactor (CELR): Alcoholic fermentation by fluidized entrapped biomass. Chem. Eng. Sci., 43, 1891-1896 (1988).
  • Govier, G. W. A. K., The flow of complex mixtures in pipes. Van Nostrand Reinhold Co., New York (1972).
  • Hatch, R. T., Fermenter design, in: Single Cell Protein II (S. R. Tannenbaum and D. I. C. Wang, eds.). MIT Press, Cambridge (1975); p 46-68.
  • Jajuee, B., Margaritis, A., Karamanev, D., Bergougnou, M. A., Mass transfer characteristics of a novel three-phase airlift contactor with a semipermeable membrane. Chem. Eng. J., 125, 119-126 (2006).
  • Jamialahmadi, M., Müller-Steinhagen, H., Effect of alcohol, organic acid and potassium chloride concentration on bubble size, bubble rise velocity and gas hold-up in bubble columns. Chem. Eng. J., 50, 47-56 (1992).
  • Joshi, J. B., Ranade, V. V., Gharat, S. D., Lele, S. S., Sparged loop reactors. Can. J. Chem. Eng., 68, 705-741 (1990).
  • Keitel, G. Untersuchungen zum Stoffaustausch in Gas-Flüssig-Dispersionen in Rührschlaufenreaktor und Blasensäule. PhD Thesis, Universität Dortmund, Dortmund, Germany (1978).
  • Kelkar, B. G., Godbole, S. P., Honath, M. F., Shah, Y. T., Carr, N. L., Deckwer, W. D., Effect of addition of alcohols on gas holdup and backmixing in bubble columns. AIChE J., 29, 361-369 (1983).
  • Kojić, P. S., Tokić, M. S., Šijački, I. M., Lukić, N. L., Petrović, D. L., Jovičević, D. Z., Popović, S. S., Influence of the Sparger Type and Added Alcohol on the Gas Holdup of an External-Loop Airlift Reactor. Chem. Eng. Technol., 38, 701-708 (2015).
  • Li, Y. Z., He, Y. L., Ohandja, D. G., Ji, J., Li, J. F., Zhou, T., Simultaneous nitrification-denitrification achieved by an innovative internal-loop airlift MBR: Comparative study. Bioresour. Technol., 99, 5867-5872 (2008).
  • Lin, J., Han, M., Wang, T., Zhang, T., Wang, J., Jin, Y., Influence of the gas distributor on the local hydrodynamic behavior of an external loop airlift reactor. Chem. Eng. J., 102, 51-59 (2004).
  • Liu, R., Huang, X., Wang, C., Chen, L., Qian, Y., Study on hydraulic characteristics in a submerged membrane bioreactor process. Process Biochem., 36, 249-254 (2000).
  • McClure, D. D., Deligny, J., Kavanagh, J. M., Fletcher, D. F., Barton, G. W., Impact of Surfactant Chemistry on Bubble Column Systems. Chem. Eng. Technol., 37, 652-658 (2014).
  • Merchuk, J. C., Stein, Y., Local hold-up and liquid velocity in air-lift reactors. AIChE J., 27, 377-388 (1981).
  • Mihaľ, M., Gavin, S., Markoš, J., Airlift reactor - membrane extraction hybrid system for aroma production. Chemical Papers, 67, 1485-1494 (2013).
  • Milivojević, M. Brzina tečnosti u dvofaznim i trofaznim pneumatskim reaktorima sa spoljašnjom cirkulacijom. PhD Thesis, University of Belgrade, Belgrade, Serbia (2011).
  • Miyahara, T., Nagatani, N., Influence of Alcohol Addition on Liquid-Phase Volumetric Mass Transfer Coefficient in an External-Loop Airlift Reactor with a Porous Plate. J. Chem. Eng. Jpn., 42, 713-719 (2009).
  • Moraveji, M. K., Sajjadi, B., Davarnejad, R., CFD Simulation of hold-up and liquid circulation velocity in a membrane airlift reactor. Theor. Found. Chem. Eng., 46, 266-273 (2012).
  • Pošarac, D. Investigation of hydrodynamics and mass-transfer in a three phase external-loop airlift reactor. PhD Thesis, University of Novi Sad, Novi Sad, Serbia (1988).
  • Prieske, H., Drews, A., Kraume, M., Prediction of the circulation velocity in a membrane bioreactor. Desalination, 231, 219-226 (2008).
  • Ratkovich, N., Chan, C. C. V., Berube, P. R., Nopens, I., Experimental study and CFD modelling of a two-phase slug flow for an airlift tubular membrane. Chem. Eng. Sci., 64, 3576-3584 (2009).
  • Rossignol, N., Vandanjon, L., Jaouen, P., Quéméneur, F., Membrane technology for the continuous separation microalgae/culture medium: compared performances of cross-flow microfiltration and ultrafiltration. Aquacult. Eng., 20, 191-208 (1999).
  • Rujiruttanakul, Y., Pavasant, P., Influence of configuration on the performance of external loop airlift contactors. Chem. Eng. Res. Des., 89, 2254-2261 (2011).
  • Schügerl, K., Lücke, J., Oels, U., Bubble column bioreactors. In Advances in Biochemical Engineering , Volume 7, Springer Berlin Heidelberg: 1977; Vol. 7, pp 1-84.
  • Shariati, F. P., Mehrnia, M. R., Salmasi, B. M., Heran, M., Wisniewski, C., Sarrafzadeh, M. H., Membrane bioreactor for treatment of pharmaceutical wastewater containing acetaminophen. Desalination, 250, 798-800 (2010).
  • Šijački, I. M., Tokić, M. S., Kojić, P. S., Petrović, D. L., Tekić, M. N., Djurić, M. S., Milovančev, S. S., Sparger Type Influence on the Hydrodynamics of the Draft Tube Airlift Reactor with Diluted Alcohol Solutions. Ind. Eng. Chem. Res., 50, 3580-3591 (2011).
  • Snape, J. B., Zahradník, J., Fialová, M., Thomas, N. H., Liquid-phase properties and sparger design effects in an external-loop airlift reactor. Chem. Eng. Sci., 50, 3175-3186 (1995).
  • Sun, Y., Li, Y. L., Bai, S., Hu, Z. D., Modeling and Simulation of an In Situ Product Removal Process for Lactic Acid Production in an Airlift Bioreactor. Ind. Eng. Chem. Res., 38, 3290-3295 (1999).
  • Verlaan, P. Modelling and characterization of an airlift-loop bioreactor. PhD Thesis, Wageningen University, Wageningen, Netherlands (1987).
  • Verlaan, P., Vos, J.-C., Van T Riet, K., Hydrodynamics of the flow transition from a bubble column to an airlift-loop reactor. J. Chem. Technol. Biotechnol., 45, 109-121 (1989).
  • Vial, C., Camarasa, E., Poncin, S., Wild, G., Midoux, N., Bouillard, J., Study of hydrodynamic behaviour in bubble columns and external loop airlift reactors through analysis of pressure fluctuations. Chem. Eng. Sci., 55, 2957-2973 (2000).
  • Vial, C., Poncin, S., Wild, G., Midoux, N., Experimental and theoretical analysis of the hydrodynamics in the riser of an external loop airlift reactor. Chem. Eng. Sci., 57, 4745-4762 (2002).
  • Vial, C., Poncin, S., Wild, G., Midoux, N., A simple method for regime identification and flow characterisation in bubble columns and airlift reactors. Chem. Eng. Process., 40, 135-151 (2001).
  • Wallis, G. B., One-dimensional two-phase flow. McGraw-Hill, New York (1969).
  • Weiland, P., Onken, U., Fluid Dynamics and Mass Transfer in an Airlift Fermenter with External Loop. Ger.Chem.Eng., 4, 42-50 (1981).
  • Xu, Z., Yu, J., Hydrodynamics and mass transfer in a novel multi-airlifting membrane bioreactor. Chem. Eng. Sci., 63, 1941-1949 (2008).
  • Yuan, Y.-J., Wei, Z.-J., Wu, Z.-L., Wu, J.-C., Improved Taxol production in suspension cultures of Taxus chinensis var. mairei by in situ extraction combined with precursor feeding and additional carbon source introduction in an airlift loop reactor. Biotechnol. Lett, 23, 1659-1662 (2001).
  • Zahradník, J., Kaštánek, F., Rylek, M., Porosity of the heterogeneous bed and liquid circulation in multistage bubble-type column reactors. Collect. Czech. Chem. Commun., 39, 1403-1418 (1974).

Publication Dates

  • Publication in this collection
    Apr 2017

History

  • Received
    24 June 2015
  • Reviewed
    25 Nov 2015
  • Accepted
    15 Jan 2016
Brazilian Society of Chemical Engineering Rua Líbero Badaró, 152 , 11. and., 01008-903 São Paulo SP Brazil, Tel.: +55 11 3107-8747, Fax.: +55 11 3104-4649, Fax: +55 11 3104-4649 - São Paulo - SP - Brazil
E-mail: rgiudici@usp.br