‘wgr. $5.32.? E." 3%.ku a 5 “5.1.1 .1... .1. z :...;...:1 — THESlS 7~ J I ' I . I This is to certify that the dissertation entitled The Enhancement of Synthesis Gas Fermentations Using Microbubble Dispersions presented by Marshall Dean Bredwell has been accepted towards fulfillment ofthe requirements for Ph . D . degree in Chemical Engineering A3565 11/ 77/116 (/3 Zb/IJ/Léec V'\ Major professor Date 7/3/00 MSU is an Affirmative Action /Equal Opportunity Institution 0-12771 LIBRARY Michigan State University PLACE IN RETURN BOX to remove this checkout from your record. TO AVOID FINES return on or before date due. MAY BE RECALLED with earlier due date if requested. DATE DUE DATE DUE DATE DUE 11/00 cJCIRC/DatoDuopGS-p.“ THE ENHANCEMENT OF SYNTHESIS GAS FERMENTATIONS USING MICROBUBBLE DISPERSIONS By Marshall Dean Bredwell A DISSERTATION Submitted to Michigan State University In partial fulfillment of the requirements For the degree of DOCTOR OF PPHLOSOPHY Department of Chemical Engineering 2000 ABSTRACT THE ENHANCEMENT OF SYNTHESIS GAS F ERMENTATIONS USING MICROBUBBLE DISPERSIONS By Marshall Dean Bredwell Synthesis gas, a mixture of primarily carbon monoxide and hydrogen, can be used in a variety of bioconversions. It is well suited as a carbon and electron source for production of liquid fiiels and chemicals or as a source of reducing equivalents for biological sulfur reductions. Synthesis gas ferrnentations have typically been gas-to- liquid mass transfer limited due to the low aqueous solubility of hydrogen and carbon monoxide, the primary components of synthesis gas. A novel technique to increase mass transport without increased power expenditure is to use microbubbles, also known colloidal gas aphrons. Microbubbles are small, surfactant-coated gas bubbles with diameters on the order 60 pm that have colloidal pr0perties, such as an electric double layer, that impart stability and useful properties such as the ability to be pumped without collapse. Surfactants that are biocompatible with the microbial biocatalysts used in synthesis gas fennentations and that are capable of forming stable microbubbles have been identified. The effects of these surfactants on the metabolic activity of Butyribacterium methylotrophicum were characterized. Non-ionic surfactants were relatively non-toxic, while ionic surfactants were toxic at concentrations necessary to form microbubbles. The mass transfer properties of microbubbles were determined in an abiotic bubble column. The local mass transfer coefficient, KL, and the volumetric mass transfer coefficient, KLa were determined as a function of surfactant concentration in the surrounding bulk liquid. Synthesis-gas F ermentations were run using the biocompatible surfactants for both the production of liquid fuels and chemicals and for the biological reduction of sulfur. The KLa value for carbon monoxide transport in a Butyribacterium methylotrophicum fermentation was measured with both conventional and microbubble sparging. A six-fold increase in KLa was obtained upon switching from conventional bubbles to microbubbles. A sulfate reducing bacteria culture was used to convert sulfur dioxide to hydrogen sulfide using synthesis gas as a source of reducing equivalents. The productivity using microbubble sparging was two-fold higher than that using conventional sparging. Dynamic microbubble coalescence was measured using video microscopy in a stirred vessel. Coalescence was modeled using a population balance approach with a film drainage model to describe the coalescence efficiency. The model predicted that the reduced coalescence rate observed at higher surfactant concentrations is due to a thickening of the liquid film between adjacent bubbles. ACKNOWLEDGMENTS I would like to first thank my wife for her support, love, and confidence in me. Without her, none of my work would have been successful. I wish to thank my advisor, Dr. Mark Worden, without whom none of this work would have been done, for his guidance and advice throughout these years. I also wish to Dr. Eric Kaufman and Dr. Punjai Selvaraj at Oak Ridge National Laboratory for their help, guidance and support on this project. I also wish to thank Dr. Costas Tsouris, also at ORNL, for giving me his time and advice in the development of both the coalescence experiments and models. I would also like to thank the other members of my graduate committee, Dr. Daina Briedis, Dr. Estelle McGroarty, Dr. Mahendra Jain, Dr. Robert Ofoli, and Dr. Andrew Grethlein. I would like to thank the colleagues with whom I spent most of graduate studies, Mark Widman and Tyler Ames. I would also like to thank a number of other people with whom I have worked with over the years. I have had the good luck in working with skilled people such as Michael Telgenhoff at Michigan State and Mark Little and Miguel Rodriguez at ORNL. I would also like to thank Chemical Engineering Department office staff of Julie Caywood, Faith Peterson, and Candy McMaster for their constant assistance. Finally, I would like to thank my parents, Dean and Patricia, for all of their support and love throughout the years. The National Science Foundation, the Department of Energy at ORNL, and the MSU Biotechnology Training Program has provided funding for this work. iv TABLE OF CONTENTS LIST OF TABLES ................................................................................. ix LIST OF FIGURES ............................................................................... x 1. INTRODUCTION .......................................................................... 1 2. OBJECTIVES AND SIGNIFICANCE .................................................. 4 2.1 Surfactant Biocompatibility ........................................................ 4 2.2 Microbubble Mass Transfer Coefficients ......................................... 4 2.3 Synthesis Gas Fermentations ....................................................... 5 2.4 Microbubble Coalescence ........................................................... 6 3. LITERATURE REVIEW .................................................................. 7 3.1 Synthesis Gas ......................................................................... 7 3.1 .1 Coal Gasification .............................................................. 7 3.1.2 Biomass Gasification ......................................................... 8 3.2 Synthesis Gas F ermentations ....................................................... 9 3.2.1 Butyribacterium methylotrophicum ......................................... 11 3.2.2 Clostrz'dium ljungdahlii ....................................................... 14 3.2.3 Sulfate Reducers ............................................................... 17 3.2.4 Other Bacteria .................................................................. 18 3.2.5 Bioreactors for Synthesis Gas Fermentations .............................. 20 3.3 Aphrons ................................................................................ 22 3.3.1 Colloidal Gas Aphrons ........................................................ 23 3.3.2 Colloidal Liquid Aphrons .................................................... 26 3.4 Coalescence ........................................................................... 27 3.4.1 Breakage Frequency ........................................................... 28 3.4.2 Collision Frequency ........................................................... 29 3.4.3 Coalescence Efficiency ....................................................... 29 4. FORMATION AND STABILITY OF MICROBUBBLES ........................... 31 4.1 Introduction ........................................................................... 31 4.2 Materials and Methods ............................................................... 32 4.2.1 Microbubble Generator ....................................................... 32 4.2.2 Bubble Size Distribution ...................................................... 32 4.2.3 Formation Studies ............................................................. 32 4.2.4 Stability Studies ............................................................... 33 4.2.5 Immobilization Studies ....................................................... 33 4.3 Results and Discussion .............................................................. 33 4.3.1 Particle Size Distribution ..................................................... 33 4.3.2 Formation Studies ............................................................. 34 V 4.3.3 Stability Studies ............................................................... 38 4.3.4 Immobilization Studies ....................................................... 43 4.4 Conclusions ........................................................................... 44 5. SURFACTANT BIOCOMPATIBLITY ................................................. 46 5.1 Introduction ........................................................................... 46 5.2 Material and Methods ............................................................... 47 5.2.1 Culture Technique ............................................................. 47 5.2.2 Culture Analysis .............................................................. 50 5.2.2.1 Gas Chromatography ................................................... 50 5.2.2.2 Spectroscopy ............................................................ 51 5.2.2.3 pH Measurements ...................................................... 51 5.2.3 Surfactant Studies ............................................................. 51 5.3 Results and Discussion .............................................................. 54 5.3.1 Dry Weight Calibration Curve .............................................. 54 5.3.2 Tween Surfactants ............................................................ 54 5.3.3 Brij Surfactants ................................................................ 57 5.3.4 Products and pH ............................................................... 60 5.4 Conclusions ........................................................................... 65 6. MASS TRANSFER PROPERTIES OF MICROBUBBLES ......................... 66 6.1 Introduction ........................................................................... 66 6.2 Materials and Methods .............................................................. 68 6.2.1 Microbubble Generation ..................................................... 68 6.2.2 Axial Dispersion .............................................................. 68 6.2.3 Mass Transfer ................................................................ 70 6.2.4 Power Consumption .......................................................... 70 6.2.5 Dynamic Gassing Out ........................................................ 71 6.2.6 Mathematical Models ........................................................ 72 6.2.6.1 Axial Dispersion ....................................................... 72 6.2.6.2 Mass Transfer .......................................................... 73 6.2.6.3 Dynamic Mass Transfer .............................................. 74 6.3 Results and Discussion ............................................................. 75 6.3.1 Axial Dispersion .............................................................. 75 6.3.2 Mass Transfer ................................................................. 76 6.3.2.1 Plug Flow Model ...................................................... 76 6.3.2.2 Void Fraction and Surfactant Effects ............................... 77 6.3.3 Power Consumption ......................................................... 81 6.3.3.1 Power Consumption in F ermentations .............................. 81 6.3.3.2 Comparison of Reactor Type ......................................... 82 6.3.4 Dynamic Gassing Out ........................................................ 84 6.4 Conclusions .......................................................................... 89 6.5 Nomenclature ........................................................................ 90 6.5.1 Symbols ........................................................................ 90 vi —¥—_—L 6.5.2 Greek Symbols ................................................................ 93 Butyribacterium methylotrophcium FERMENTATIONS ............................ 94 7.1 Introductions ......................................................................... 94 7.2 Materials and Methods ............................................................. 94 7.2.1 Media and Cultures ........................................................... 94 7.2.2 Reactor ........................................................................ 95 7.2.3 Analytical Techniques ....................................................... 96 7.2.4 Models ......................................................................... 97 7.2.5 Carbon and Electron Balances .............................................. 97 7.3 Results and Discussion ............................................................. 98 7.3.1 Products ........................................................................ 98 7.3.2 Mass Transfer Coefficients .................................................. 101 7.4 Conclusions ........................................................................... 102 7.5 Nomenclature ......................................................................... 102 7.5.1 Symbols ........................................................................ 102 7.5.2 Subscripts and Superscripts ................................................. 102 SULFATE REDUCING BACTERIA FERMENTATIONS ......................... 104 8.1 Introduction ........................................................................... 104 8.2 Materials and Methods .............................................................. 105 8.2.1 Microbubble Generation ..................................................... 105 8.2.2 Media and Cultures ........................................................... 106 8.2.3 Mass Transfer Limitation Studies .......................................... 110 8.2.4 Fermentations .................................................................. 11 1 8.2.4.1 Conventional Sparging ................................................ 111 8.2.4.2 Microbubble Sparging ................................................. 112 8.2.4.3 Mass Transfer Studies ................................................. 112 8.2.5 Analytical Techniques ........................................................ 1 13 8.2.5.1 Gas Chromatography .................................................. 113 8.2.5.2 Sulfur Chemistry ....................................................... 114 8.2.5.3 Most Probable Number (MPN) and Misc. Analytical ............ 115 8.2.6 Carbon and Electron Balances .............................................. 115 8.3 Results and Discussion .............................................................. 116 8.3.1 Confirmation of Mass Transfer Limitations .............................. 116 8.3.2 Conventional Sparging ....................................................... 1 19 8.3.3 Microbubble Sparging ........................................................ 123 8.3.4 Carbon and Electron Balances .............................................. 125 8.4 Conclusions ........................................................................... 126 8.5 Nomenclature ......................................................................... 127 8.5.1 Symbols ........................................................................ 127 8.5.2 Greek Symbols ................................................................ 127 8.5.3 Subscripts and Superscripts ................................................ 127 vii 10. 11. 12. COALESCENCE ........................................................................... 129 9.1 Introduction ........................................................................... 129 9.2 Mathematical Model ................................................................. 131 9.2.1 Population Balance Equation Model ....................................... 131 9.2.2 Collision Frequency ........................................................... 134 9.2.3 Coalescence Efficiency ....................................................... 136 9.2.4 Solution Techniques .......................................................... 140 9.3 Materials and Methods .............................................................. 140 9.3.1 Coalescence Measurements ................................................. 140 9.3.2 Particle Size Distributions ................................................... 141 9.3.3 Experimental Variables ...................................................... 142 9.3.4 Surface Tension ............................................................... 143 9.4 Results and Discussion .............................................................. 143 9.4.1 Data Quality .................................................................... 143 9.4.2 Deformable Drop Efficiency Model ........................................ 146 9.4.3 Film Drainage Efficiency Model for Bubbles ............................. 149 9.4.3.1 Model Results ........................................................... 149 9.4.3.2 Physical Significance of Fitted Parameters ......................... 154 9.5 Conclusions ........................................................................... 159 9.6 Nomenclature ......................................................................... 159 9.6.1 Symbols ........................................................................ 159 9.6.2 Greek Symbols ................................................................. 161 9.6.3 Subscripts and Superscripts .................................................. 162 SUMMARY AND CONCLUSIONS .................................................... 163 APPENDICIES - FORTRAN PROGRAMS .......................................... 165 11.1 Axial Dispersion ................................................................... 166 11.2 Bubble Coalescence ............................................................... 172 BIBLIOGRAPHY ........................................................................... 178 viii Table 1: Table 2: Table 3: Table 4: Table 5: Table 6: Table 7: Table 8: Table 9: Table 10: Table 11: Table 12: Table 13: Table 14: Table 15: Table 16: Table 17: LIST OF TABLES Effects of disk distance on bubble diameter .................................... 34 Phosphate buffered basal salts medium composition ......................... 48 Composition of trace mineral solution .......................................... 49 Composition of Vitamin solution ................................................ 50 Composition of phosphate stock solution ...................................... 50 Surfactant critical micelle concentration ....................................... 53 Carbon and electron balance results for Tween surfactants .................. 63 Mass transfer rates ................................................................. 80 Composition of the enhance PBB medium ..................................... 95 Reductance degrees ............................................................... 98 Lactic acid media .................................................................. 107 Minimal salts media ............................................................... 107 Batch vitamin solution ............................................................ 108 Heavy metal solution............................., ............................... 108 Trace elements 11 .................................................................. 109 Synthesis gas uptake from serum bottles ....................................... 119 Surface tension ..................................................................... 143 ix Figure 1: Figure 2: Figure 3: Figure 4: Figure 5: Figure 6: Figure 7: Figure 8: Figure 9: Figure 10: Figure 11: Figure 12: Figure 13: Figure 14: Figure 15: Figure 16: Figure 17: Figure 18: Figure 19: Figure 20: Figure 21: Figure 22: LIST OF FIGURES Microbubble generator ............................................................ 23 Schematic of a microbubble ...................................................... 25 Bubble size distribution using dynamic light scattering ...................... 34 Equilibrium formation time for microbubbles ................................. 35 Equilibrium formation time in the presence of salts .......................... 36 Foam and foam-liquid interface level in drainage experiments ............. 38 Initial foam gas void fractions for microbubbles .............................. 39 Microbubble stability ............................................................. 40 Drainage experiment using alginate microbubbles ........................... 44 Chemical structure of Tween and Brij surfactants ............................ 52 Dry weight calibration curve ..................................................... 54 Growth of B. methylotrophicum in batch bottles with Tween .............. 55 Specific growth rate of B. methylotrophicum with Tween .................. 56 Growth of B. methylotrophicum in batch bottles with Brij .................. 57 Specific growth rate of B. methylotrophicum with Brij ...................... 58 Acetate production ................................................................ 60 Ethanol production ................................................................ 61 Butyrate production ............................................................... 62 pH profile of a Tween 80 fermentation ......................................... 64 Experimental apparatus used in axial dispersion and mass transfer. . . . . 69 Axial dispersion coefficients ..................................................... 75 Optimized fit for KL using a plug-flow model ................................. 77 Figure 23: Figure 24: Figure 25: Figure 26: Figure 27: Figure 28: Figure 29: Figure 30: Figure 31: Figure 32: Figure 33: Figure 34: Figure 35: Figure 36: Figure 37: Figure 38: Figure 39: Figure 40: Figure 41: Figure 42: Figure 43: Figure 44: Figure 45: Mass transfer coefficients from microbubbles ................................. 78 Comparative KLa values for varying reactor types ............................ 83 Experimental data from a dynamic gassing out run .......................... 84 Linear form of Equation 19 ...................................................... 85 Percent oxygen transferred to liquid phase .................................... 86 Dynamic mass transfer coefficients of microbubbles ........................ 88 Flow diagram for a B. methylotrophicum fermentation ...................... 96 Fermentation product profile ..................................................... 99 Mass transfer coefficients from fermentation .................................. 101 802 reducing fermentation flow diagram ...................................... 112 Potentiometric titration of sulfide ............................................... 114 CO uptake by SRB in a batch serum bottle .................................... 117 H2 uptake by SRB in a batch serum bottle ..................................... 118 Mass transfer coefficient for CO ................................................ 121 Mass transfer coefficient for H2 .................................................. 122 SRB reactor productivity ......................................................... 124 Captured video image of microbubbles ......................................... 144 Bubble size distribution histogram .............................................. 145 Coalescence using efficiency modeled with Equation 53 .................... 147 Average bubble size using efficiency modeled with Equation 53 ......... 148 Coalescence in 0 mg/L Tween 20 using Equation 63 ........................ 149 Coalescence in 180 mg/L Tween 20 using Equation 63 ..................... 150 Coalescence in 300 mg/L Tween 20 using Equation 63 ..................... 151 xi Figure 46: Dynamic average bubble size .................................................... 152 Figure 47: Coalescence efficiency ............................................................ 153 Figure 48: Number of bubbles in vessel ..................................................... 154 Figure 49: Effect of surfactant on C1 ......................................................... 156 xii 1. INTRODUCTION Synthesis gas, a mixture containing carbon monoxide (CO), hydrogen (H2), carbon dioxide (CO2) and trace sulfur compounds can be readily obtained from the gasification of coal or biomass (Worden, 1997). Synthesis gas can be anaerobically converted into alcohols such as ethanol (Grethlein, 1990) and butanol (Grethlein, 1991a) which have been typically made from petroleum. The advantages of using bioprocessing to convert synthesis gas to liquid fuels rather than typical catalytic techniques are that coal-derived synthesis gas is a potentially less—expensive feedstock than agricultural products such as starch, CO2 production can be limited with the addition of small amounts of secondary electron donors (Grethlein, 1991b), and many bacteria that are capable of converting synthesis gas (such as Butyribacterium methylotrophicum and Clostridium ljungdahlii) have high tolerances to sulfur compounds (Grethlein, 1992b), (Vega, 1990a). Biological processes also exhibit high substrate and product specificity, which minimizes formation of by-products and can operate at relatively low temperatures and pressures. Typical problems that occur with the bioprocessing of synthesis gas are that product concentrations are typically low and that the transport of synthesis gas components to the liquid phase are slow due to inherently low aqueous solubilities. The conventional method to increase gas-to-liquid mass transport in stirred tanks is to increase the power input to enhance bubble breakup and thereby increase the effective area for mass transport. This approach is not economically feasible in large-scale reactors because power consumption is proportional to the impeller rate to the third power and impeller diameter to the fifih power (McCabe, 1985). Additionally, the high shear rates created by the impellers can be detrimental to cellular metabolism (Bailey, 1986). A potential way to overcome the gas-to-liquid mass transport limitations of synthesis gas fermentations is the use of microbubble dispersions. Microbubbles (or colloidal gas aphrons) are small, surfactant coated bubbles with a potentially a multi- layered shell stabilized by surfactant (Sebba, 1984). Microbubbles typically have diameters on the order of 50 pm in comparison to 2 to 5 mm for conventional bubbles found in sparged fennenters (Kaster, 1990). The surfactant shell imparts an electrical charge to the bubbles’ surface and helps to prevent coalescence, giving the bubbles colloidal properties such as the ability to be pumped without collapse unlike conventional foams (Longe, 1989). The mass transfer rate, in mass transport limited situations, is directly proportional to the bubbles’ interfacial area per unit gas volume. Therefore increasing the interfacial area available for mass transport without high power input would be beneficial. Another advantage to using microbubbles with synthesis gas fermentations is that typically synthesis gas consists of over 95% consumable substrate, and as the components are transferred out of the bubble, the bubble will shrink. This shrinkage will further increase the area per unit gas volume and increase the efficiency of mass transfer. Additionally, the bubbles’ internal pressure increases as the bubbles shrink due to curvature and surface tension effects (Rosen, 1989), increasing the gas component’s liquid phase equilibrium concentration. Therefore, higher dissolved CO tensions can be obtained without increasing the overall pressure in the system. This research is presented in eight chapters. Chapter 2, the Objectives and Significance, contains the goals and requirements of this project as well as the importance of this work. Chapter 3 contains a literature review, which gives pertinent background information into this work. Identification of surfactants that can form microbubbles is given in Chapter 4, Formation and Stability of Microbubbles, and are biocompatible with microorganisms in Chapter 5, Surfactant Biocompatibility. The mass transfer coefficients and power requirements of microbubbles are examined in Chapter 6, Mass Transfer Properties of Microbubbles. The results of fermentations that have benefited using microbubble sparging are presented in Chapter 7, Butyrz‘bacterium methylotrophicum Fermentations and Chapter 8, Sulfate Reducing Bacteria Fermentations. Finally, the results of coalescence studies on microbubbles are given in Chapter 9, Coalescence. 2. OBJECTIVES AND SIGNIFICANCE 2.1 Surfactant Biocompatibility Surfactants increase the stability of microbubbles and decrease the average diameter (Kaster, 1990), (Longe, 1989). The addition of external surfactant significantly reduces their rate of drainage (Amiri, 1990) and thereby decreases their rate of coalescence. Most previous research on microbubbles used an ionic surfactant to supply the microbubble surface with a significant charge for use with applications such as floatation or harvesting. In most cases, ionic surfactants are toxic to the cellular metabolism due to their ability to form ionic bonds with the cellular proteins, while non- ionic surfactants tend to form much weaker hydrogen bonds and are therefore less toxic. Therefore, the first objective of this study was to identify surfactants that allow formation of microbubble dispersions while not being detrimental to cellular metabolism. 2.2 Microbubble Mass Transfer Coefficients Little information exists in the literature on the mass transfer rate from microbubbles. The mass transfer characteristics of microbubbles, with their potentially complex outer shell, have not been well characterized. Kaster et a1. (1990) obtained values for the volumetric mass transfer coefficient in a Saccharomyces cerevisiae fermentation using microbubble sparging but without the addition of an external surfactant. Walso and Gal-Or (1971) have suggested empirical equations to describe the mass transfer coefficient from swarms of small rigid bubbles, but these equations do not account for the possible presence of a thick liquid shell. The determination of the mass transfer properties of microbubbles enables a more accurate description of the processes occurring during fermentations using microbubbles. Therefore, the second objective was to study the mass transport characteristics of microbubble. 2.3 Synthesis Gas Fermentations Continuous cell-recycle fermentations have been demonstrated for B. methylotrophicum (Grethlein, 1990), Clostridium ljundahlii (Klasson, 1993), and for the class of bacteria known as sulfate reducing bacteria (SRB) (van Houten, 1994). In order to use microbubble sparging in these fermentations, modifications to handle increased liquid filtration and potential equipment difficulties will have to be made. Mass balances around the reactors can be used to calculate the volumetric mass transfer coefficient as has been done in previous, conventionally sparged fermentations (Klasson, 1992). Important parameters such as bioreactor productivity, mass transfer coefficients, and substrate conversion will be compared to conventionally sparged fermentations. Subsequently, the third objective of this study is to show improved reactor productivity and increased mass transfer in continuous, microbubble-sparged, synthesis-gas fermentations. 2.4 Microbubble Coalescence Microbubbles coalesce through drainage of the film between two adjacent bubbles. The rate of this coalescence could be related to the thickness of the surrounding film around a microbubble, the rate at which the liquid drains, and the steric hindrance effects of the surfactant. The rate of coalescence could also be related to the conditions that prevail in the vessel with microbubbles, i.e. turbulent flows giving rise to a greater number of collisions and thereby increasing the likelihood of coalescence. The rate of coalescence is an important factor in understanding the phenomena occurring in a stirred tank during fermentation. A fundamental understanding of the coalescence dynamics occurring in a stirred vessel would allow for accurate prediction in average bubble size during fermentations. Design criteria for microbubble—sparged fermentation processes could also be developed using this information. Therefore, the final objective of this work is to conduct a fundamental study into the coalescence phenomena of microbubbles. 3. BACKGROUND AND LITERATURE REVIEW 3.1 Synthesis Gas Synthesis gas, a mixture of carbon monoxide (CO) and hydrogen (H2) gases, is produced by the partial oxidation of an organic feedstock at high temperature in the presence of steam. Presently, synthesis gas for fuels and chemicals is produced primarily by catalytic steam reformation of natural gas or by partial oxidation of heavy oil fractions. The future of synthesis gas production lies in coal gasification, due to its availability and abundance, and in biomass gasification, due to its renewability and cost. The composition of synthesis gases can vary significantly depending on the material being gasified and the type of gasifier used. Typically, most synthesis gases have, other than CO and H2, methane (CH4), carbon dioxide (CO2), water (H20) and a variety of contaminants (Simbeck, 1983). Typical contaminants also depend on the material being gasified, but commonly are sulfur compounds such as carbonyl sulfide (COS) and hydrogen sulfide (H28) and nitrogen compounds such as ammonia (NH3). These contaminants are costly to remove either pre or post burn and can lead to the formation of SOx and NOX which are causes of acid rain and ozone layer degradation. 3.1.1 Coal Gasification Coal is a readily available, low-cost source for synthesis gas for chemicals and fuels. The estimated coal surplus in the United States as of 1989 was 250 million tons of coal, which corresponds to a 220-year surplus at current usage rates (Mallory, 1979). Coal gasification was a heavily used technology in the 1950’s until it was replaced by lower cost raw materials such as natural gas and petroleum. Typically, synthesis gas is produced from coal through steam gasification, where at high temperatures and pressures, coal, water and oxygen (02) are combined to form a mixture that is primarily CO and H2. The direct composition of the synthesis gas is dependent on the type of gasifier, the gasification conditions, and the source of the gasification coal. Also, the final composition of the synthesis gas depends on the application for which the synthesis gas is to be used. When used as a substitute for natural gas, the formation of CO is minimized and emphasis is placed on CH4 and H2 compositions. When used as a source of electric power, a high CO:H2 ratio is desired. Finally, for chemical fuels production, a specific CO:H2 ratio is desired that is based on the product formation stoichiometry. There are several different types of coal gasifiers that are currently in use and depend primarily on different techniques for contacting the coal with O2 and H20. These included fixed or moving bed gasifiers, which typically operate at lower temperatures with high solid residence times, entrained-bed gasifiers, which typically operate at high temperatures, and fluidized-bed gasifiers, which operate at moderate temperatures and have rapid and complete gas-solids mixing. The type of gasifier used strongly influences the composition of the resulting synthesis gas. An in depth review of existing gasifier technology is given in Grethlein (1991b). 3.1.2 Biomass Gasification Biomass gasification offers several advantages of over conventional feedstocks such as coal and petroleum. Most biomass has a much lower sulfur content than most coals. Synthesis gas that was derived from wood chips at a GE gasification plant contained 28 ppm of H28 (Worden, 1997) compared to up to 2% in coal-derived synthesis gas. The purification steps to remove sulfur compounds from synthesis gas are costly in both sorbent recycling and disposal (Kaufman, 1996a) and in energy and equipment costs. Biomass materials are also more reactive than coal, and therefore require lower gasification temperatures and pressures as well as shorter residence times. Fluidized-bed coal gasifiers are typically run at 1000 °C with residence times of up to 3 hours while biomass gasification require temperatures of 850 °C with residence times from 30 s to 5 min (Worden, 1997). Finally, the gasification of biomass provides for the disposal of wood and yard wastes as well as utilizing a renewable resource. The economics favor the gasification of biomass over the burning of biomass because the basic conversion efficiency of gasification is about 32% compared with 15%- 20% for the combustion route (Parkinson, 1996). The final composition of the gas obtained from biomass gasification is dependent on the operating conditions, gasifier temperature, and whether it is air-blown or oxygen-blown. An investigation of the how gasifier conditions effect the final gas composition are given in Maschio (1994) and in Kinoshita (1991). The first biomass gasification facility came on line in Fall of 1995 in Vamamo, Sweden that uses a pressurized, air-blown gasifier. This plant is in the 30 to 60 MW range. Other ventures, using a variety of feedstocks, are being brought online. In Hawaii, the Pacific Center for High Technology is using IGT’s Renugas process to generate a low BTU gas from sugar cane bagasse and in Minnesota, the Renugas process is being used to produce gas from alfalfa stems (Parkinson, 1996). 3.2 Synthesis Gas Fermentations Synthesis gas can be catalytically converted into chemical products such as methanol in reactors operated at high temperatures and pressures (Kinoshita, 1991) which can then be used as a fuel additive or as a precursor to other chemicals. A major catalytic route for synthesis gas is through Fischer-Tropsch synthesis. Fischer-Tropsch is a catalytic approach to convert CO and H2 to a variety of paraffinic, olefinic, and oxygenated compounds, which can then be used as either fuel additives or desirable chemicals. A key disadvantage to the use of Fischer-Tropsch synthesis is the lack of catalyst specificity, generating a wide range of fractions with a low molecular weight fraction, a gasoline fraction, and a diesel oil fraction (Grethlein, 1991b). The only products that can be formed with 100% selectivity are the Cl compounds. This creates difficulty in generating higher molecular weight alcohols and acids (e.g. butanol and butyrate) because the existing catalysts give a wide range of products. The existing catalysts are also easily poisoned by sulfur compounds, which must be removed prior to undergoing Fischer-Tropsch synthesis. Finally, conventional catalytic techniques usually require a strict ratio of CO to H2, which is typically adjusted prior to the catalytic reaction through separate gas-shifting reactors to modify the synthesis gas composition through the water-gas shift reaction. The biological conversion of synthesis gas has many advantages of conventional catalytic techniques. First, the reaction specificity is high with fermentation conditions able to be optimized to produce specific products (e.g. ethanol or acetate) (Phillips, 1993), (Grethlein, 1991b). Second, biological conversion happens at temperatures and pressures around room temperature and atmospheric pressure. Third, most biological catalysts used are highly tolerant to sulfur species (Grethlein, 1992b), (Reis, 1992) and can in many cases be used without prior removal. It has been estimated that the operating costs for the production of clean synthesis gas can amount to over half of the total 10 operating costs for coal processes such as SASOL (Grethlein, 1991b). Finally, biological conversion does not have a set CO/I-I2 ratio in respect to the metabolic reactions. Some of the key disadvantages to biological processing result in low reactor productivies. Work has been done on this issue in attempts to increase the productivies from fermentations. With some high productivity bacteria, such as Clostridium ljungdahlii, under development this issue is becoming no longer as significant as it was. Development in the separations of dilute aqueous fermentation products has lead to the use of extractive distillations to remove ethanol. A company in Fayetteville, AR (Bioengineering Resources, Inc.) have made significant process developments in the production of ethanol from synthesis gas and have given cost estimates that are competitive with the costs of producing ethanol from fermentation of agricultural materials. A key drawback in the fermentation of synthesis gas components is the low aqueous solubilities of CO and H2, being 60% and 4% by mass as soluble as oxygen, a sparingly soluble gas (Worden, 1997). This results in synthesis gas fermentations being gas-to-liquid mass transport limited (Worden, 1989), (Klasson, 1992). 3.2.1 Butyribacterium methylotrophicum B. methylotrophicum is an obligate anaerobe that is capable of growth on a variety of substrates including glucose (Moench, 1983), forrnate and methanol (Kerby, 1987), H2 and CO2 (Lynd, 1982), and CO (Lynd, 1982). When grown on methanol in the presence of acetate and CO2, a nine-hour doubling time was observed (Lynd, 1983) while the CO adapted strain was able to grow on 100% CO gas with a doubling time of 12 hours (Lynd, 1982). The CO adapted strain has been shown to produce a variety of substrates that is linked to the fermentation pH (Grethlein, 1990). At pH 6.8, the primary fermentation 11 product is acetate with a molar ratio of acetate:butyrate of 32:1 (Grethlein, 1990). At pH 6.0, acetate and butyrate are produced in nearly equimolar amounts. Ethanol and butanol are minor products with butanol concentrations up to 2.7 g/L being produced at low pH (Grethlein, 1991a). The acetogenic metabolism of B. methylotrophicum has been elucidated for CO and CO2 utilization for the formation of acetyl-CoA and butyryl—CoA , the formation of ethanol and butanol, and the formation of acetate and butyrate (Shen, 1996). A detailed description of the enzymatic pathways involved in the product formation from B. methylotrophicum can be found in Grethlein et a1. (Grethlein, 1992a). The primary steps involved in the CO metabolism involve the reaction of CO and H20 with CO dehydrogenase to yield CO2, the production of formate from CO2 via formate dehydrogenase, tetrahydrofolate-linked transformations resulting in a methyl-corrinoid complex, the reaction of CO with CO dehydrogenase to form an enzyme-bound carbonyl moiety, and the synthesis of acetyl-CoA from both the methyl and carbonyl groups bound to the CO dehydrogenase complex. The consumption of H2 and CO2 occur via the same mechanism with H2 being used to reduce CO2 to bound CO (Grethlein, 1992a). Continuous fermentations have been developed to exploit the CO metabolism of B. methylotrophicum to produce products such as acetate, butyrate, ethanol, and butanol (Grethlein, 1990). The stoichiometry of the unicarbonotrophic metabolism of CO is given in Equation 1 below (Grethlein, 1991b). 4C0 —> 2.17C02 + 0.74CH3C00H + 0.45Cell (1) Similarly, the carbon-balanced reaction for growth on H2 and CO2 is given in Equation 2 below (Grethlein, 1991b). 12 2H, +1.03C0, —> 0.43CH,C00H + 0.13Cell (2) The products of these fermentations have been found to be linked to the pH as mentioned previously, with more than half of the incoming carbon was converted to CO2 and acetate the primary liquid-phase product, accounting for 67% of the reduced-product carbon. One trend with continuous fermentation using slow-growing anaerobes such as B. methylotrophicum is that the biomass concentration in the reactor is low and subsequently product concentrations are lower. The average biomass concentration in previous fermentations was approximately 0.28 g/L (Grethlein, 1990). The low specific growth rate of the organism, a 12 hour doubling time, requires very low dilution rates in order to prevent cell washout. The low dilution rates limit the reactor’s volumetric productivity, which was approximately 0.001 g/L*h for alcohols and 0.015 g/L*h for acids (Grethlein, 1990). A total cell-recycle can be used to increase the biomass and product concentrations and therefore increase volumetric productivities. In the case of B. methylotrophicum, increases of 830% for acetate, 850% for butyrate, 820% for ethanol, 1481% for butanol, and 2440% for cell mass have been obtained using a total cell recycle over a continuous fermentation without cell recycle (Grethlein, 1991b). This represents an 8-9-fold increase in acid production and an 8-14 fold increase in alcohol production. The inhibitory effects of a number of products and potential synthesis gas components have been investigated. The effect of the product alcohols, ethanol and butanol, have been investigated up to approximately 12 g/L (Grethlein, 1991b). It was shown that for ethanol, significant inhibitory effects start to occur at concentrations above 5 g/L and for butanol these effects started at concentrations of 2.8 g/L. This indicates that butanol is the more toxic product (Grethlein, 1991b). The effect of the alcohols on the 13 bacteria is through damage to the cellular membrane, where the alcohol can increase fluidity of the membrane, causing a number of detrimental effects such as the disruption of the nutrient transport and energy generation systems. The effect of the produced acids, acetate and butyrate, were also studied in concentrations up to 12 g/L. There did not seem to be significant inhibition due to these acids, but the fermentations did give evidence for co-metabolism of CO gas and either acetate or butyrate (Grethlein, 1991b). The effect of sulfide on the metabolism was examined through the addition of sodium sulfide (Na2S) and hydrogen sulfide gas (H2S) (Grethlein, 1992b). The addition of H28 gas to the headspace severely inhibited the both cell growth and acetate production in concentrations higher than 3.2%. With the addition of Na2S to the liquid phase, inhibition was seen to occur in the growth rate at concentrations between 1.4% and 3.3%, but the specific and molar yields of acetate and butyrate were not significantly effected. The results from this data indicate that below 2% H28 in the headspace, there was little inhibition from sulfide and therefore should not affect the metabolism of B. methylotrophicum during coal-derived synthesis gas conversion (Grethlein, 1992b). 3.2.2 Clostridium ljungdahlii Clostridz'um ljungdahlz'z', a gram-positive, motile, rod-shaped anaerobic bacteria was isolated in 1987 from chicken waste at the University of Arkansas. It was shown to be capable of converting CO, H2, and CO2 to a mixture of acetate and ethanol (Barik, 1988). In addition to synthesis gas components, C. ljungdahlii is also capable of growth on xylose, arabinose, and fructose (Klasson, 1992). It is expected that the ethanol and acetate are formed through an acetyl—CoA intermediate. It has been shown to produce 14 ethanol in concentrations up to 48 g/L in a CSTR with cell recycle (Klasson, 1993). The production of ethanol and acetate from C0 and H2 is shown stoichiometrically below. 6C0 +3H, —2 C, H5011 +4C0, (3) 2C0, +611, —> CZHSOH +3H,0 (4) 4C0 +2H, 0 —> CH,C00H +2C02 (5) 2C0, +4H, —> CH,CO0H +2H,0 (6) The ‘wild’ type strain of C. ljungdahlii showed an ethanol/acetate ratio of only 0.05 with a maximum production of ethanol of 0.1 g/L (Vega, 1989a). Work with the fermentation parameters, such as lowering the pH to 4.0, has brought about a significant increase in the ethanol/acetate ratio up to 3 (Gaddy, 1995). The development of a designed medium, without yeast extract, has been able to virtually invert the ethanol and acetate production (Phillips, 1993) and nearly eliminate acetate as a product. These studies also found that both the product concentration and cell concentration, which ranged from 1 to 2 g/L using a total cell recycle, were independent of inlet gas flow rate (Phillips, 1993). Using the designed medium, after 25 days, the ethanol concentration was as high as 48 g/L (Phillips, 1993). Research with C. ljungdahlii has shown that the presence of reducing agents in liquid media has brought about an increase in solvent formation (Klasson, 1991a). The reducing agents cause altered electron flow which directs the flow of carbon to acid and alcohol production. The reducing equivalents are transferred in the form of NADH, which results in increased alcohol formation. The addition of small quantities of reducing agents such as methyl and benzyl viologen, sodium thioglycollate, and ascorbic acid, have been able to double and quadrupole the ethanol/acetate ratio (Klasson, 1992). Other 15 research investigating the connection between sporulation and increased solventogenesis has identified that a shift of the bacteria into a sporulation state is accompanied by morphological changes and increased solventogensis (Klasson, 1992). Studies with C. ljungdahlz'i have been done using complex nutrients such as cellobiose, rhamnose, and galactose, which were found to promote sporulation in other Clostrz’dium species (Klasson, 1992). The increase in ethanol concentration was the‘highest in the presence of cellobiose and galactose, where ethanol concentrations were four times the concentration obtained in the presence of yeast extract and cell concentrations were 20% higher (Klasson, 1992). Since ethanol production seems to be linked to a phase of the bacteria that is on the edge of sporulation, studies using two continuous reactors in series have been done (Klasson, 1992). The scheme uses the first reactor to promote cell growth, while the second reactor is used for the production of ethanol. The conditions and media constituents in the first reactor are optimized for cell growth, while in the second reactor, the pH is dropped from 4.5 to 4.0, the dilution rate is shifted to control the growth rate, and media additions such as cellobiose are added. This two-reactor scheme was capable of a 30-fold improvement in specific productivies over a single reactor (Klasson, 1992). Reactor development using C. ljungdahlii that is currently under investigation include the use of increased pressure in fermentations in order to overcome gas-to-liquid mass transport limitations that are prevalent in synthesis gas fermentations. High-pressure fermentations can be useful in minimizing reactor volume requirements. Pressure fermentations have been successful for some synthesis gas bacteria, such as Peptostreptococcus productus, to a point where the liquid phase carbon monoxide 16 concentration has become high enough to cause inhibition of metabolism (Vega, 1990b). Therefore, it has become necessary to develop conditions were the carbon monoxide tension remains low despite increased pressure. This has been done through high cell concentrations, obtained through feeding alternative substrates and gradual stepwise increasing of the cell concentration (Vega, 1990b). Pressure fermentations using C. ljungdahlii have been done at pressures up to 2.5 atm (Vega, 1989a) and work is underway to obtain a pressure adapted strain for higher pressures. Another technique to enhance gas-to-liquid mass transfer using C. ljungdahliz' is the use of solvents. Synthesis gas components are more soluble in a munber of organic compounds such as toluene and heptane (Halling, 1994), and their introduction into the fermentation media may result in an increase in gas-to-liquid mass transfer rates. An increase of 2.2 fold was seen for oxygen gas in a 15% by volume media with the prefluorocarbon, FC40, in waste gas treatment (Cesario, 1997). Recently, work by Rodriguez (M. Rodriguez, unpublished results) have seen a 4-fold increase in the mass transfer coefficient in serum bottle fermentations of C. ljungdahlii using a 20% by volume media with hexadecane. 3.2.3 Sulfur Reducers The class of bacteria known as sulfate reducing bacteria (SRB), is capable of using a variety of sources for reducing equivalents in the reduction of sulfur compounds. Microbial processes utilizing sulfate-reducing bacteria (SRB) have found potential application in treatment processes of sulfur laden wastes. In particular, processes for flue gas desulfurization (Selvaraj, 1995), gypsum recovery (Kaufinan, 1996b), sulfur recovery from sulfite/sulfate wastewater from pulp and paper, chemical and mining industries 17 (Hammack, 1994), and degrading explosive material. The cost of feedstock and bioreactor productivity are critical parameters in the economic viability of the microbial processes (Selvaraj, 1996b). Lactic acid, ethanol, hydrogen, synthesis gas, and sewage digest have all been proposed as electron donors for SRB. Lactic acid and ethanol are most likely too expensive of a feedstock for the process. In specific locations, sewage digest could be available at low or negative cost (Selvaraj, 1996b). Synthesis gas is an attractive feedstock for the SRB process due to its wide availability from coal or biomass gasification, its zero COD discharge, and it possible availability on site. Various groups have demonstrated that SRB could be supported by carbon dioxide (CO2) and/or CO as their sole carbon source and hydrogen as their energy source (van Houten, 1994), (du Preez, 1992). Recently, van Houten et al. (1994) operated a gas-lift, sulfate-reducing reactor with a feedstock of CO/I-I2 with CO concentrations up to 20% that was capable of a maximum sulfate conversion rate of 10 g SO42'/(d*L) and du Preez et al. (1994) operated a pilot-scale packed bed reactor with a feedstock of CO that was capable of a maximum rate of 2.4 g SO42'/(d*L). 3.2.4 Other Synthesis Gas Bacteria A number of other bacteria have been shown to be capable of using synthesis gas components for the production of a variety of compounds, in most cases acetate. None of the other bacteria have been to shown to produce ethanol at the same level as C. ljungdahlii or as many different products as B. methylotrohicum. A strain of Peptostreptococcus productus that was capable of utilizing CO as the energy source at mesophillic temperatures was isolated from an anaerobic sewage digester in 1984. This 18 bacteria grew quickly, td = 1.5 h, with up to 50% CO in the headspace with acetate and CO2 as the major products. The reaction stoichiometry is presented below (Vega, 1989b). 4C0+2H20—>2C02 +CH3C00H (7) It was also seen that P. productus was able to grow at higher concentrations of CO (up to 90%), but an increase in the concentration of yeast extraction was required. Significant work has been done with P. productus. Through batch culture, key kinetic parameters, such as specific uptake rate and yield coefficients that are usable in the prediction of reactor performance (Vega, 1989d) have been determined. A model has been developed using the parameters estimated from batch culture to evaluate the conversion and levels of CO as a function of inlet gas flow rate under mass transfer- limited conditions (Vega, 1989c). Since the primary product from P. productus is acetate rather than ethanol, work has drifted from P. productus to higher ethanol producers such as C. ljungdahlii. Methane (CH4) gas can also be generated from synthesis gas through fermentation. A mixed culture of Rhodospirillum rubrum, a photosynthetic bacterium, and two methanogens, Methanosarcina barkerz', and Methanobacterium formicum has been used to convert synthesis gas to CH4 (Klasson, 1990). The triculture carries out the same reactions as the water gas-shift reaction and methanation. These reactions are given below. C0+H,0—>H,+C02 (8) 4H2+C0,—>CH4+2H20 (9) In the triculture, R. rubrum carries out reaction 8 while M. barkeri and M. formicum carry out reaction 9. The reason that two different organisms were used to convert H2 to CH4 19 was that while M. formicum has a high rate of H2 uptake, it is inhibited in the presence of CO, and while M. barkeri has a tolerance to CO, it has a lower rate of H2 uptake (Klasson, 1990). This triculture has been used in many different reactor configurations, continuous stirred tank reactors (CSTR), bubble column reactors, and trickle-bed reactors (TBR) (Vega, 1990b) to convert synthesis gas into methane. The yield of methane from H2 were estimated to be 83% of theoretical and 100% CO conversion was obtained in these reactors (Vega, 1990b). 3.2.5 Bioreactors for Synthesis Gas Fermentations Due to the mass transfer limitations that occur in synthesis gas fermentations, the design of such reactors and fermentation systems centers around increasing the mass transfer coefficient (Klasson, 1991b). The mass transfer capabilities of the reactor must be sufficient enough to generate high cell densities and supply enough substrate to the liquid phase. Many different types of reactors have been examined for possible use in synthesis gas fermentations: CSTR, packed—bed, bubble column, gas-lift and trickle-bed. Free-cell reactors such as the CSTR have used a variety of techniques to increase both biocatalyst density and to help overcome mass transfer limitations. Two stage reactor schemes for C. ljungdahlii have been used where one reactor vessel is used to promote growth and a second reactor in series has environmental conditions, such as low pH and a shift in the dilution rate, optimized for product formation (Klasson, 1991b). The Michigan Biotechnology Institute (MBI) has used two stage reactor schemes for the conversion of synthesis gas components to butanol using one reactor to convert synthesis gas to butyrate using B. methylotrophz'cum and the second reactor to convert butyrate to 20 butanol using Clostridium acetobutylicum (Grethlein, 1992a). Grethlein et al. (1991b) used headspace and cell recycling in tandem to ensure mass transfer limitation operation and maintain high cell density for the production of n-butanol with B. methylotrophcium. Stirred-tank reactors are capable of achieving high mass transfer rates but require substantial energy input for agitation. An increase in impeller speed from 300 rpm to 450 rpm increased the volumetric mass transfer coefficient from 28.1 h’1 to 101.1 h'1 using a mixed triculture fermentation (Klasson, 1992). This approach, while effective, results in dramatic increases in power consumption. This increase is especially dramatic for large- scale systems because power consumption is proportional to the impeller rate to the third power and impeller diameter to the fifth power (McCabe, 1985). Immobilized cell reactors such as a trickle-bed reactor have been used as a possible alternative to the free-cell reactors. Immobilized cell reactors eliminate the need for mechanical agitation and expensive cell-recycle systems and usually give high interfacial area and high mass transfer coefficients (Charpentier, 1981). Klasson et. a1. (Klasson, 1992) found for a triculture fermentation that CO conversion for a given gas loading rate was two-fold greater in a trickle-bed reactor with the cells immobilized on ceramic saddles than in a CSTR. Selvaraj et a1. (Selvaraj, 1997a) found an eight-fold increase in productivity for the reduction of 802 to S?" by SRB in a trickle-bed reactor with the cells immobilized on BIO-Sep beads over a CSTR on a sewage digest media. Similar results were also obtained using coal synthesis gas as a feedstock (Selvaraj, 1996a). A packed-bed reactor was used in a pilot plant for the reduction of sulfate and nitrate waste using SRB and synthesis gas as a feedstock (du Preez, 1994). Similarly, air- lift reactors have been used with SRB fed a synthesis gas substrate (van Houten, 1994). 21 Multiple groups have increased pressure in reactor schemes to increase the carbon monoxide and hydrogen tensions (du Preez, 1994), (Klasson, 1992). Recently, Soni et al. (1998) have tried increased pressure on methanogens. The idea is that higher pressure will allow for higher levels of dissolved CO and H2 which will lead to higher concentrations of biomass (du Preez, 1994). Theoretical results using fermentation parameters for P. productus indicate that for a 10 fold increase in total pressure, a 16 fold increase in the CO uptake by the bacteria could be obtained (Klasson, 1991b). One drawback to the use of increased pressure is the inhibitory effects of higher levels of dissolved CO tension. Thus, maintaining low dissolved CO tensions in high pressure fermentations is of great importance and can be achieved by having high reactor cell densities (Vega, 1990b). High reactor cell densities in high-pressure reactors for P. productus have been obtained by stepwise increases in pressure (Vega, 1990b). This approach allows the cells to gradually acclimate to the increased pressure. Work is ongoing to develop a high-pressure adapted strain of C. ljungdahlii (Klasson, 1992). 3.3 Aphrons The term aphron comes from the Greek word aphros, which means foam. The existence of aphrons was first reported by Sebba (1987). They typically are below 100 pm in diameter and are encapsulated by a surrounding liquid film and encased in surfactant, which acts to stabilize them. There are two different types of aphrons, colloidal gas aphrons (CGA) where the dispersed phase consists of a gas, and colloidal liquid aphrons (CLA) where the dispersed phase consists of an organic liquid or oil with a soluble surfactant dissolved within. While being similar in structure, CLA and CGA have significantly different properties and generation techniques. 22 3.3.1 Colloidal Gas Aphrons Collodial gas aphrons, also known as microbubbles are made by creating a high- shear zone at a gas-liquid interface. Devices such as a venturi throat (Sebba, 1987) and a spinning disk apparatus (Sebba, 1985) (Figure 1) have previously been used to generate dispersions. When a gas bubble becomes trapped in the liquid, it can be pulled into the high-shear zone and further broken down into smaller and smaller bubbles. The surfactant present in the liquid is absorbed at the interface stabilizing the smaller bubbles and reducing the rate of coalescence in the system. Microbubbles also exhibit colloidal properties that allow them to be pumped, unlike conventional foams (Longe, 1989). Stirring Motor 5 Liter Fermenter Spinning Disk Figure 1: Microbubble generator It has been suggested that microbubbles have a complex outer shell arrangement, having both a surfactant bilayer at the liquid shell-bulk liquid interface and a surfactant 23 monolayer at the gas-liquid shell interface as shown in Figure 2 (Sebba, 1987). This liquid shell has been likened to the liquid film that surrounds a soap bubble blown into air (Sebba, 1987). The surfactant molecules in the outer bilayer impart an electrical double layer that reduces bubble coalescence through electrical repulsion (Sebba, 1987). The electric double layer of bubbles created in aqueous solutions of surfactant has been shown to be quite significant, ranging from —250 mV with a strong, ionic surfactant (Usui, 1981), to 60 mV with a weak, non-ionic surfactant (Laskowski, 1989), (Collins, 1978). The zeta potential, a measurement of the surface charge on a bubble or particle, has also been shown to decrease with increasing bubble size (Usui, 1981). The significance of the zeta potential in microbubble dispersions is the stability provided to the microbubbles through the electric charge, and the shorter range van der Waals forces. It can be shown that the repulsive forces between colloids is proportional to the surface potential (Schramm, 1992). 24 Figure 2: Schematic of a microbubble Microbubbles have been used in aerobic fermentations of S. cerevisiae that were generated with only the natural surfactants produced by the bacteria. The mass transfer coefficients obtained from this microbubble-sparged fermentation were found to be independent of impeller rate and 4 times that of conventional sparging at low impeller rates (Kaster, 1990). The independence of power input on the mass transfer is an added benefit of microbubbles in that a minimal impeller rate is needed to only maintain adequate mixing and high power input is not necessary to promote bubble breakup. Microbubbles have also been used for the bioremediation of soil to increase the amount of oxygen retained in the soil and to promote increased mass transfer to flowing groundwater (Michelsen, 1991). Jenkins et al. (1993) found increased oxygen utilization using microbubbles over conventional oxygenation methods during in situ biodegradation of xylene in soil columns. 25 Another major use for microbubbles has been in floatation. Microbubbles have been used to harvest S. cerevisiae from a suspension (Save, 1995). CGA offers a rapid and efficient method for the recovery of microorganisms, greater than that obtained through conventional foam fractionation. It was also found that surfactant type played a significant role in developing attractive forces between the cell and CGA. Anionic surfactants, such as sodium dodecyl sulfate (SDS), did not promote good attachment to cells while hydrophobic cationic surfactants, such as cetyl pyridinium chloride (CPC) gave high separation factors (Save, 1995). Models have been developed to describe the separation of microorganisms by CGA (Save, 1994). CGA have also been proposed to be used for the recovery of proteins through interaction between the aphron and the protein due to electrostatic and hydrophobic forces. The aphron phase would then rise due to buoyancy and thereby separate proteins (J auregi, 1996). The floatation process has also been used extensively in clarification and washing. A review of microbubble floatation processes has been written by Solari et a1. (1992). CGA have been used to clarify pahn oil mill effluent, suspensions of algae, and suspensions of inorganic minerals, primarily through electrostatic or hydrophobic interactions (Subramaniam, 1990). CGA have extensively been used for the washing of soils to remove oily wastes (Roy, 1992a); (Roy, 1994), the separation of heavy metals (Cabezon, 1994), the separation of organic dyes (Roy, 1992b), and solvent sublation (Caballero, 1989). 3.3.2 Colloidal Liquid Aphrons A similar type of dispersion is a colloidal liquid aphron (CLA). CLAs have an organic or oil phase in an aqueous environment. CLAs have the same structure as 26 microbubbles (Sebba, 1984), and their size ranges from submicron to 50 um in diameter (Zhang, 1996a). In order to produce CLA, the oil (or internal) phase has to be brought to a small size, and these dr0plets must be surrounded by the soapy film. This can be accomplished by spreading the oil phase on a dilute surfactant solution. For the oil to spread on the surfactant solution, the oil must have an oil soluble surfactant present at a concentration that is just sufficient to produce a spreading coefficient greater than the surface pressure produced by the surfactant dissolved in the aqueous phase. CLA have been found to be extremely stabile. Zhang et al. (1996a) found that after 230 days of storage, the maximum in the peak size shifted from 11 pm to 18 pm for kerosene in water CLAs. CLAs allow for very high concentrations of the oil phase to be dispersed in water. Phase volume ratios (PVR) up to 20 have been prepared without phase inversion or coalescence from kerosene (Sebba, 1984). CLAs have been used for predispresed solvent extraction in a kerosene-water system (Matsushita, 1992) and in enzyme extraction (Save, 1993). They have also been used to separate a hydrophobic organic dye from water through a combination CGA/CLA floatation process (Zhang, 1996b). Finally, in liquid-liquid extraction, an increase of 16 fold in the mass transfer rate over conventional extraction was obtained using CLA (Matsushita, 1992). 3.4 Coalescence Bubble size in a stirred vessel is dependent on a balance between the coalescence and break-up rates in the vessel. Coalescence is the process by which two (or more) bubbles, particles, drops, etc come together to with sufficient energy and remain together for a long enough time for a single unit to form. Breakage (or break-up) is the process by which sufficient energy is applied to shear the unit into multiple daughter units. The 27 coalescence and breakage in complex systems, such as fermentation broths are functions of many parameters, such as surface characteristics of the bubbles or drops, dynamic surface tension, bulk and dispersed phase properties, energy dissipation, size, etc. In the analysis of dispersive systems, the population-balance-equation (pbe) approach has been used extensively to model coalescence and breakage (Coulaloglou and Tavlarides (1977), Das et al. (1987), and Prince and Blanch (1990)). This approach describes the history of a population such as size, age, and concentration, during the course of interaction events between particles, bubbles, or drops, and with the surrounding environment. A pbe model is also effective in that it can accommodate continuous mass transfer between phases, flow in and out of the systems, as well as coalescence and breakage events. This model consists of three major parameters, the breakage frequency, the collision frequency and the collision efficiency. 3.4.1 Breakage Frequency The breakage frequency is how often bubbles and drops undergo a breakage event. Breakage has been proposed to occur through a number of mechanisms such as drop elongation in a shear flow field (Taylor, 1934), turbulent pressure fluctuations (Hinze, 1955), relative velocity fluctuations (Narismhan et al., 1980) and drop-eddy collisions (Coulaloglou and Tavlarides, 1977). A number of mechanistic models have been developed that provide simple mathematical expressions for drop breakage rates. These models are based on the physical properties of the system such as drop diameter, interfacial tension, and dispersed phase density, the geometry of the vessel, and the energy provided to the dispersion by agitation. 28 3.4.2 Collision Frequency The collision frequency is the rate at which bubbles or drops collide. Commonly the bubble-bubble (or drop-drop) collision frequency is modeled assuming that the bubbles behave like gas molecules in a turbulent flow (Coulaloglou and Tavlarides, 1977) and that their collision processes can be described by an analogy to the kinetic theory of gasses (Kennard, 1938). Making these assumptions, the collision frequency becomes a function of the bubbles' size, their velocities (and therefore the energy dissipation in the system) and the number of bubbles in the system. 3.4.3 Coalescence Efficiency The coalescence efficiency is the percentage of collision events in a system that will result in a coalescence event. The coalescence efficiency is a function of two characteristic parameters: the contact time and the coalescence time. Once bubbles have collided, they will stay in contact for some time period, known as the contact time, in which they will either coalesce or separate. The contact time between bubbles has typically been determined to be dependent on the energy dissipation in the system and the bubbles' sizes (Coulaloglou and Tavlarides, 1977). Coalescence occurs if the contact time is long enough that the liquid film between the two bubbles drains away until a critical film thickness where rupture occurs. This is the coalescence time (Tsouris and Tavlarides, 1994). Therefore, coalescence will occur when the contact time is greater than the coalescence time. A number of different models for the coalescence efficiency have been proposed. Coulaloglou and Tavlarides (1977) used an expression that was developed from models for film drainage between pairs of drops immersed in an incompressible stagnant viscous 29 liquid. Sovova (1981) suggested that the model used by Coulaloglou and Tavlarides (1977) favored the coalescence of smaller drops and replaced it with an expression that describes the ratio of the interfacial energy in the system to the energy of collision. It has been suggested by Tsouris and Tavlarides (1994) that this form of the coalescence efficiency favors the coalescence of larger drops. Das et al. (1987) introduced a white— noise model, which considers the film drainage between two colliding drops as a stochastic process driven by a suitably idealized random process for the fluctuating force. According to this model, the drops are considered nondeformable, and the critical thickness of the film before rupture is assumed to be specific for a given system. Muralidhar and Ramkrishna (1986) employed a time-scale analysis in order to understand the significance of factors affecting drop coalescence. They considered both deformable and nondeformable drops and under which conditions the white-noise model and a band- limited model are valid. Muralidhar et al. (1988) studied the coalescence of rigid drops in a stirred dispersion and developed a colored-noise model. The white noise, band-limited, and colored-noise models have a large number of parameters that are difficult to measure or estimate (Tsouris and Tavlarides, 1994) and are computationally difficult. Tobin et al. (1990) studied the effect of drop size on coalescence frequencies and concluded that the coalescence of small drops is not as frequent as the model of Coulaloglou and Tavlarides (1977) predicts. Tsouris and Tavlarides (1994) suggest a kinetic collision mechanistic model to describe the coalescence. 30 4. Formation and Stability of Microbubbles 4.1 Introduction Fermentations that involve a gaseous substrate are inherently rate-limited by mass transfer from the gaseous to the liquid phase. Synthesis gas fermentations are a primary example of such mass-transfer-limited fermentations. The traditional approach to enhance gas-to-liquid mass transfer is to increase the impeller rate, thereby increasing the interfacial area available for mass transfer. However, this approach results in a dramatic increase in power consumption, especially for large-scale systems, because power consumption is proportional to the impeller rate to the third power and the impeller diameter to the fifth power (McCabe et al., 1985). Microbubble dispersions offer an alternative, providing high interfacial area with the potential for low power input. In aerobic fermentations with yeast, the transport of oxygen from a gas to a liquid, reported in terms of kLa, was found have a 4-fold increase in using microbubble sparging over conventional air sparging (Kaster, 1990). In order to study microbubbles in fermentations, it is first necessary to identify criteria that define parameters such as stability and formation. The stability parameters have been suggested by Amiri and Woodburn (1990) to be dependent on the drainage rate of the liquid film between microbubbles. The formation rate of microbubbles, and the size distribution are suggested to be dependent on the salt concentration or ionic strength of the solution (J auregi, 1996) 31 4.2 Materials and Methods 4.2.1 Microbubble Generator Microbubble dispersions were formed using a spinning-disc device first described by Sebba (1985) and shown schematically in Figure 1. This device employs a high—speed electric motor (Talboys #37830, Cole Partner, Chicago, IL) to spin a 5-cm-diameter by l- cm-thick disk at speeds above 4000 rpm. Stationary baffles located within 5 mm of the spinning disk result in a localized high-shear zone. The stainless-steel disk and baffles are mounted in a 6 L fermentation vessel with the motor attached to the headplate. The generator was adjusted to determine the optimal configuration in terms of disk placement based on visual observation of microbubble formation. 4.2.2 Bubble Size Distributions Bubble-size distributions were measured with a Malvem Mastersizer (Malvem, Inc.) particle size analyzer, which uses a light scattering technique (Chaphalkar, 1993). Bubbles were loaded into the recycle flow 100p containing water with surfactant until an obscuration of 8-12% was obtained. Data were taken for 60 seconds. The Mastersizer uses a light-scattering technique that gives accurate size distribution data within seconds of injecting the sample. In contrast, microscope-based image-analysis methods are relatively slow and require many individual measurements to give statistically meaningful results. 4.2.3 Formation Studies In the formation-rate experiments, the dispersion was considered fully formed when an equilibrium foam volume was obtained. Dispersions were formed at 7000 rpm. Experiments were done with three types of surfactant: non-ionic Triton X-100 (Sigma 32 Chemical Co., St. Louis, MO), anionic sodium dodecyl sulfate (SDS) (Sigma Chemical Co.) and cationic cetyl pyridinium chloride (CPC) (Sigma Chemical Co.). Salt (NaCl) was added to the microbubble generator (MBG) in runs with Triton to determine its effects on microbubble formation. 4.2.4 Stability Studies In the drainage and stability experiments, the microbubble dispersion was generated, poured into a 1 L graduated cylinder, and allowed to drain. The heights of the air-foam and foam-liquid interfaces were measured as a firnction of time. Runs with salt (NaCl) dissolved in the liquid phase were done in a similar fashion. 4.2.5 Immobilization Studies Immobilized microbubbles were made by generating by using a 2.0% (w/v) solution of alginic acid with the addition of 240 mg/L of Tween 20 surfactant. The alginic acid solution was prepared by allowing the material to mix for 24 hours on a magnetic stirrer until homogeneous. Microbubbles were then formed from the alginic acid solution in the MEG. The expected 3-fold volume expansion was not obtained; rather a 2-fold expansion was obtained. The solution was then dripped into a 25 g/L calcium chloride (CaCl2) solution that was agitated by a magnetic stirrer. The immobilized microbubbles were allowed to cure for 30 nrinutes prior to removal from the CaCl2 solution. 4.3 Results and Discussion 4.3.1 Particle Size Distribution The number-averaged bubble diameter was 58.9 pm for Triton X-100 at a concentration of twice that of the CMC (135 mg/L) and a temperature of 25°C. The average bubble size measured by the Mastersizer is in the range of literature values of 40 33 to 80 um (Longe, 1989). A sample of the data obtained from the particle size analyzer is shown in Figure 3 below. The bubble size distribution was unaffected by the distance the spinning disk was to the baffles within 5 mm. The results from those experiments are summarized in Table 1. 12. 10. Percentage O) 0 50 100 150 200 Bubble Size (um) Figure 3: Bubble size distribution using dynamic light scattering Table 1: Effect of disk distance on bubble diameter Disk Distance from Baffles Bubble Diameter 1 mm 57.50 um 3 mm 68.55 pm 5 mm 54.45 pm 34 4.3.2 Formation Studies Figure 4 shows the equilibrium formation time for microbubble dispersions as function of surfactant type and dimensionless surfactant concentration (DSC), which is defined as the ratio of surfactant concentration to its critical micelle concentration. Time of Microbubble Formation (min 16 . 14 _ -a—Triton, non-ionic +808, anionic 12 .. +CPC, cationic 10 . 8 . 6 . 4 . \Z~\ . ‘0 2 . 0 : : : : : : : : : O 0.5 1 1.5 2 2.5 3 3.5 4 4.5 Dimensionless Surfactant Concentration Figure 4: Equilibrium formation time for microbubbles 35 Figure 5 shows the effects of the addition of NaCl to the generation media on the equilibrium formation time. 16.. 14 .. 12 u- 10., -5- RO Water Time of Microbubble Formation (min CO 6 .. —a—0.08 g/L NaCl +0.8 g/L NaCl 4 .. 2 "r 0 : : : : : : : i : o 0.5 1 1.5 2 2.5 3 3.5 4 4.5 Dimensionless Surfactant Concentration Figure 5: Equilibrium formation time in the presence of salts Experiments were conducted at DSC values down to 0.25, but stable dispersions did not always form. If a stable dispersion did not form after 20 minutes, no data are shown for that concentration. A stable dispersion was indicated by a three-fold volume expansion during foam formation. The surfactant concentration is an important variable in the formation of stable dispersions. There is believed to be a double layer of surfactant around each nricrobubble (Sebba, 1987); however, when the surfactant concentration is much below the CMC, stable microbubble dispersions do not form. In such cases, there is insufficient surfactant present to stabilize the large amount of interfacial area (Longe, 1989). Figure 4 indicates that as the surfactant concentration was increased above the CMC, the dispersion 36 formation time decreased asymptotically to a value that is characteristic of the surfactant used. This trend is thought to be due to saturation of the surfactant-adsorption capacity of the nricrobubbles. Further addition of surfactant has a negligible effect because the additional surfactant molecules remain suspended in the fluid film surrounding the microbubbles. For ionic surfactants such as CPC and SDS, the amount of surfactant necessary to achieve this asymptotic formation time was approximately the CMC. For non-ionic surfactants, the asymptote occurred at a higher concentration. This result may be due in part to CMCs of ionic surfactants typically being higher than that of non-ionic surfactants. In this study the CMCs were 0.24 mM for Triton, 0.90 mM for CPC, and 8.27 mM for SDS. Ionic surfactants need to overcome the electrical repulsion of charged groups to form micelles (Rosen, 1989). The number of surfactant molecules present at the CMC for ionic surfactants is therefore greater than that for non-ionic surfactants. The addition of an electrolyte changes the CMC for ionic surfactants (Rosen, 1989), presumably because charge shielding reduces intermolecular repulsion. However, salts have little effect on the CMC of non-ionic surfactants (Sebba, 1987). This trend is reflected in Figure 5, in which dispersion-formation time was found to be independent of salt concentration. 37 4.3.3 Stability Studies Figure 6 shows a typical time course for the drainage and stability experiments for a SDS foam with a DSC of 2. 1200 .. *Foam 1000 " ““““““ . . , ‘ ‘ . +|nterface .. 800 .. ““““ .l .5. E, 600 w 2 >° 400 .. 200 .. 0 " : i : : : : 0 10 20 30 4O 5O 60 Time (min) Figure 6: Foam and Foam-Liquid interface level in drainage experiments 38 Figure 7 shows the effects of surfactant concentration on initial gas void fraction (eg°). The value of ago was unaffected by NaCl in the range of 0.2 to 2.4 g/L for all surfactant concentrations tested (data not shown). 1 .- 0.9 .- 0.8 u. c .9 *5 0.7 .. E W0 U- 0.6 -- 1' E o > 0-5 .- 3 o 4 (D ' " -a-Triton, non-ionic E 0.3 .. —A-SDS, anionic “ 'E -o-CPC, cationic _ 0.2 .- Oo1 .- 0 : 'fi : 1'. F i 0 1 2 3 4 5 6 Dimensionless Surfactant Concentration Figure 7: Initial foam gas void fractions of microbubbles 39 Figure 8 shows foam stability, defined as the ratio of the volume of liquid drained after four minutes to the initial volume of foam, as a function of surfactant type and concentration. 0.7 .- 0.6 .. +Triton, non-ionic 0-5 -- +Triton, with 0.08 g/L NaCl ,2 +CPC, cationic :73 0.4 .. -e-SDS, anionic E (D E 0.3 .. as o u. 0.2 .. 0.1 J- A # fl 0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 Dimensionless Surfactant Concentration Figure 8: Microbubble Stability Microbubble dispersions are an aggregate of spherical bubbles surrounded by a thin liquid film called the lamella. When the dispersion is first formed, the microbubbles are spherical, and the lamella is thick. Such dispersions are termed 'wet' and have good stability (Sebba, 1987). As the dispersion is allowed to drain, the liquid in the lamella drains into the plateau borders, regions where several bubbles meet. The liquid in the lamella drains in two different stages. Drainage occurs by gravity while the lamella are still thick and the microbubbles are still spherical. As the lamella thins, the microbubbles lose their spherical shape and begin to become polyhedral (Sebba, 1987). At this stage, 40 drainage of the dispersion is a result of surface-tension differences that depend on pressure differences throughout the lamella. The pressure differences arise from differences in the radius of curvature of the lamella. The greater the bubble size in the dispersion, the greater the drainage rate as a result of surface tensions differences (Vold, 1983), (Rosen, 1989). At some critical thickness of the lamella (50-100 1:1), the surrounding film will rupture (Rosen, 1989). When foam is poured into the graduated cylinder, buoyancy also contributes to foam settling. Microbubble dispersions will "cream" (i.e. faster rising bubbles will migrate to the surface). The time required for crearning is depends on the rise velocity of the bubbles, which, according to Stokes law is proportional to the bubble radius squared (Vold, 1983). The crearning process leads to a gradient in bubble size across the height of the column. Amiri and Woodbum (1990) have shown that buoyancy effects result in a gradual increase in gas void fraction and bubble diameter from the bottom to the top in a column of microbubble dispersion. Another factor contributing to the degradation of microbubble dispersions is "bubble ripening" (i.e. the grth of larger bubbles at the expense of smaller ones). The difference in the interior pressures (AP) between two bubbles having radii R1 and R2, is given by the Laplace equation, AP=20(—1—-—1—) (10) R1 R2 where 0' is the surface tension. Because the pressure is greater inside smaller bubbles, there will be net transport of gas molecules through the liquid phase fiom smaller to larger bubbles. As the larger bubbles grow, they rise more rapidly and thus migrate to the top of the foam column. 41 The initial gas void fraction, ego, is a measure of the amount of gas initially trapped in the microbubble dispersion. Figure 7 shows ego to be independent of surfactant type and concentration and approximately constant. The value of eg", between 0.65 and 0.70, approaches 0.741, the tightest packing arrangement for monosized spheres in a face-centered cubic cell (Amiri, 1990), (Brown, 1981). The theoretical void fraction based on a body-centered cubic cell is 0.68 (Brown, 1981), (Weaire, 1993). For the non- ionic surfactant, ago did not change with the addition of sodium chloride. The addition of salt also has little effect on the CMC for non-ionic surfactants, as mentioned earlier (Sebba, 1987). Spherical bubble foams (also known as early or wet foams) are very stable, and will break only when the bubbles rise to the surface under the influence of gravity. As the foam begins to cream and drain, the bubbles grow larger and change their shape from spherical to polyhedral. Polyhedral bubbles can achieve a tighter packing; thus the gas void fraction increases with not only height along the column but also with time. Polyhedral foams (also known as late or dry foams) are much less stable. The lamellae of these foams thin as they drain and break when they can no longer resist the upward movement of the less dense gas trapped in the cells. The gas void fraction at which foams typically burst is 0.89 (Weaire, 1993). Foam stability depends primarily on two factors: the rate of drainage of liquid from the lamella and strength of the fihn that encapsulates each bubble (Sebba, 1987). At concentrations below the CMC, foam stability was found to increase with increasing surfactant concentration. At concentrations above the CMC, foam stability did not change appreciably with surfactant concentration. A possible explanation for this observation is that as the concentration of surfactant in the liquid 42 phase increased, the density of surfactant adsorbed on the interface also increased until saturation was reached. 4.3.4 Immobilization Studies Successful immobilization of microbubbles with a size range of 50 to 80 um (observed by electron microscopy) was obtained through the usage of calcium alginate. A large number of easily distinguishable microbubbles was seen in each sphere. These immobilized microbubbles have potential for future studies where changes in bubble size could be observed. They also have potential uses in the efforts to probe the outer structure of a microbubble. The alginate microbubble foam drainage (without curing) was measured and is presented in Figure 9 below. The lamella drains at a more constant and slower rate than microbubbles formed without alginate most likely due to increased viscosity in the surrounding lamella. Increased viscosity slows the rate of coalescence in the system by increasing the coalescence time of microbubbles (see Chapter 9). 43 1200 .. 1000 1N v v v v : t 3 Cw 800 .. Z? .5. g 500 __ +Foam Volume 2 + Liquid Volume § 400 .. 200 . 0 : i i : i 0 5 10 15 20 25 30 35 40 45 Time (min) Figure 9: Drainage experiment using alginate microbubbles 4.4 Conclusions The size distribution of microbubble dispersions can be rapidly measured using a light-scattering method. The number-averaged diameter for the microbubbles used in this study was 60 um. Formation times of microbubble dispersions decreased with increasing surfactant concentration to an asymptotic value that varied with surfactant type. This asymptotic value was usually achieved at surfactant concentrations around the CMC. Foam stability also exhibited saturation behavior; stability increased asymptotically with surfactant concentration to a constant value near the CMC. Microbubble foams were found to have a constant initial gas void fraction that was independent of surfactant concentration and type. The experimental eg" value of 0.66 approaches the theoretical packing limit for mono-sized spheres. Salt concentrations from 44 0 to 2.4 g/L had virtually no effect on either the void fraction or the formation time for the non-ionic surfactant. Finally, by increasing the viscosity of the continuous phase, a decreased drainage rate was seen, and therefore increased bubble stability. It was also possible to immobilize microbubble in a calcium alginate matrix for later future studies. 45 5. Surfactant Biocompatibility 5.1 Introduction The obligate anaerobe Butyribacterium methylotrophicum consumes carbon monoxide, CO, to produce acetate, ethanol, butyrate and butanol (Worden, 1989) as described previously in Chapter 2. Its broad range of useful products and its ability to grow on 100% CO make B. methylotrophz’cum one of the most versatile unicarbonotrophs. One of the key disadvantages of using this microorganism is low product concentration and slow growth rates (Grethlein, 1991b). Another key disadvantage to fermentation is that most fermentations are gas-to-liquid mass transport limited (Worden, 1989). Microbubbles have the potential to significantly increase the rate of gas-to-liquid mass transfer. Microbubbles are made in the presence of a surfactant that acts to stabilize microbubbles and reduce the rate of coalescence through imparting an electric charge to the bubble’s surface. As determined in Chapter 4, the addition of a surfactant is necessary to generate a microbubble dispersion. Results fiom Chapter 4 indicated that the concentration of the surfactant should be near or above its critical micelle concentration (CMC) in order for a stable dispersion to form. To be of use for fermentation, the surfactant should not be toxic to cellular growth, inhibit the formation of key products, or significantly change fermentation parameters (such as pH) This chapter examines the growth and product formation of B. methylotrophicum during CO fermentations in the presence of surfactants that are suitable for generating microbubble dispersions. This information is needed to select surfactants for use in microbubble-sparged synthesis gas fermentations. 46 5.2 Materials and Methods 5.2.1 Culture Techniques All chemicals and vitamins were obtained from Sigma Chemical Company (St. Louis, MO), except for sodium dodecyl sulfate which was obtained from Boehringer Mannheim (Mannheim, Germany). The nitrogen (N2) and CO gases were purchased from AGA Gas and Welding (Lansing, MI). Butyribacterium methylotrophicum was obtained from the Michigan Biotechnology Institute (Lansing, MI) and was grown anaerobically at 37 °C in a phosphate buffered (PB), sulfide-reduced medium prepared as previously described (Kerby, 1987) on 100% carbon monoxide. Table 2 gives the media components and their concentration. Resazaurin was used as an oxygen indicator in maintaining stock cultures but was not used in the experimental fermentations due to interference with measuring optical density. The medium was dispersed into 125 mL serum bottles (Wheaton, Millville, NJ) at 50 mL per bottle and sealed with butyl rubber stoppers and aluminum crimps (Wheaton, Millville, NJ). For the experimental runs with surfactant present in the media, the surfactant was added to the media prior to autoclaving. A 1% innoculum of an actively growing culture of B. methylotrophicum was used. The cultures were incubated in an Innova 4000 shaker (New Brunswick Scientific, New Brunswick, NJ) at 100 rpm in the dark. 47 Table 2: Phosphate buffered basal salts media composition Component Concentration NaCl 0.9 g/L MgCl2-2H2O 0.2 g/L CaCl2-2H2O 0.1 g/L NH4C1 1.0 g/L Yeast Extract 1.0 g/L Trace Mineral Solution 10 mL/L Vitamin Solution 10 mL/L Phosphate Stock 25 mL/L 2.5% Na2S-9H2O Stock 25 mL/L 0.2% Resazurin 1.0 mL/L The composition of the trace mineral solution is given in Table 3, the composition of the Vitamin solution is given in Table 4, and the composition of the phosphate stock is given in Table 5. 48 Table 3: Composition of trace mineral solution Component Concentration Nitrilotriacetic acid 12.8 g/L FeSO407H2O 0.10 g/L MnCl204H2O 0.10 g/L CoCl206H2O 0.17 g/L CaCl202H2O 0.10 g/L ZnCl2 0.10 g/L CuCl20H2O 0.02 g/L H3303 0.01 g/L NaMbO4 0.01 g/L NaCl 1.00 g/L Na2Se03 0.017 g/L NiSO4o6H2O 0.026 g/L 49 Table 4: Composition of vitamin solution Component Concentration Biotin 2.0 mg/L Folic acid 2.0 mg/L Pyridoxine HCl 10.0 mg/L Thiamine HCl 5.0 mg/L Nicotinic acid 5.0 mg/L Pantothenic acid 5.0 mg/L Cyanocobalamin 0.1 mg/L p—aminobenzoic acid 5.0 mg/L Thioctic acid 5.0 mg/L Riboflavin 5.0 mg/L Table 5: Composition of phosphate stock solution Component Composition KH2PO4 150 g/L K2HPO4 290 g/L 5.2.2 Culture Analysis 5.2.2.1 Gas Chromatography Liquid samples (2 mL) were taken throughout the duration of each experiment. To prepare samples for product analysis, 10% (v/v) of 1 M phosphoric acid was added. After a ten-minute incubation at 37 °C, the samples were centrifuged and analyzed by gas 50 chromatography for acetate, ethanol, butyrate, and n-butanol. The separation was done with a Perkin-Elmer Autosystems GC (Perkin Ehner, Norwalk, CN) with a Hayesep R, 6’ x 'A” x 2 mm deactiglass column (Alltech, Waukeegen, WI) and a flame-ionization detector. The injector, column, and detector were 220 °C, 170 °C, and 220 0C respectively. The mobile phase was helium at 25 mL/min. 5.2.2.2 Spectroscopy Growth was analyzed using a Lambda 3 spectrophotometer (Perkin Elmer, Norwalk, CN) at 660 nm immediately after withdrawing the sample. A calibration curve of optical density versus dry weight was constructed. After measuring optical density, the samples were vacuum filtered through a tared 0.22 pm filter disk (Millipore). The samples were then dried at 85 °C for 24 hours. The filter was allowed to cool in a desicator for 30 nrinutes prior to weighing. 5.2.2.3 pH Measurements The pH was monitored through a phenol red spectroscopic method. A 1 mL sample of cells was centrifuged in a Fisher Microcentrifuge 235C (Fisher Scientific, Chicago, IL) at 15,000 rpm for 10 minutes. Phenol red was then added to the sample to 20 uM, and the optical density was measured at 560 nm to determine the pH of the sample. The phenol red assay was calibrated using larger samples with a pH electrode. 5.2.3 Surfactant Studies The primary surfactants used for this study were Tween (polyoxyethylene sorbitans) and Brij (polyoxyethylene alcohols). The structures of these surfactants are shown in Figure 10. The number of ethylene oxide adducts is indicated by EC. 51 (OCHZCH2)n40H CH E0adducts=20=n1+n2+n3+n4 0 /\ . __ / \ Tween 20. R — monolaurate HZC CH CHZ(OCH2CHZ)TI30CR Tween 40: R = monopalmitate \ / Tween 80: R = monooleate HC— CiH H0(CH2CH20>n1 (OCHZCHanZOH CH CH O CH CH O H 3'7752: "=2 3( 2)is l 2 2 ln Brij56: ":10 321758: n=20 Figure 10: Chemical Structure for Tween and Brij surfactant (Ash, 1981). DSC values up to 3 were used because previous studies (Chapter 4) indicated that microbubble dispersions needed DSC values of at least 1 for the formation of stable dispersions. Increasing DSC values above 3 had no effect on the dispersion’s stability, presumably because the surfactant absorption saturation value of the available gas-liquid interface had been reached. The critical micelle concentrations and aggregation numbers for the surfactants used in these experiments are listed in Table 6 (Ash, 1981). 52 Table 6: Surfactant critical micelle concentration Surfactant Critical Micelle Concentration Brij 52 < 7 uM Brij 56 7 uM Brij 58 77 uM Tween 20 60 mg/L Tween 40 29 mg/L Tween 80 13 mg/L 53 5.3 Results and Discussion 5.3.1 Dry Weight Calibration Curve The dry weight-optical density calibration curve is given in Figure 11. 0.0005 '1' 0.00045 .. 0.0004 .. 0.00035 -. 0.0003 .. 0.00025 .. 0.0002 .. 0.00015 .. 0.0001 .. 0.00005 . y = 0.000742X + 0.000013 R2 = 0.973290 Cell Dry Weight (gImL 0 0.1 0:2 0:3 0:4 0:5 0.6 0.7 Optical Density (at 660 nm) Figure 11: Dry weight calibration curve 5.3.2 Tween Surfactants Figure 12 shows growth curves for B. methylotrophicum with Tween surfactants as a function of surfactant concentration. The control culture containing no surfactant was prepared and inoculated at the same time and under the same conditions as the experiment bottles. The medium for the entire experiment was made from the same batch. The data in Figure 12 are the average of triplicate runs. Figure 13 shows the specific growth rate of B. methylotrophicum determined from the exponential growth phase data in Figure 12. 54 Cell Concentration (gIL) 0.9 .. 0.8 u 0.7 .. 0.6 .. 0.5 .. 0.4 .. 0.3 .. 0.2 .. 0.1 .. —e—Control —o—Tween 20, DSC=1 —e—Tween 20, DSC=2 +Tween 20, DSC=3 - o- .Tween 40, DSC=1 - 0- .Tween 40, DSC=2 . A. .Tween 40, DSC=3 —o—Tween 80, DSC=1 +Tween 80, DSC=2 +Tween 80, DSC=3 : ,G"O - = " /\"'°~ r. = ‘ 3 150 200 250 300 350 Time (hours) Figure 12: Growth of B. methylotrophicum in batch bottles with Tween 55 0.015 .. -e—Tween 20 0.01 .1. _a—Tween 40 +Tween 80 Specific Growth Rate (h '1) 0.005 .. I l I I I I I I I 0 0.5 1 1.5 2 2.5 3 3.5 Dimensionless Surfactant Concentration Figure 13: Specific growth rate of B. methylotrophicum with Tween None of the Tween surfactants strongly inhibited growth of B. methylotrophicum in batch bottle fermentations. However there did seem to be an effect of the length of the hydrophobic end of the surfactant on the growth curves. Tween 20 has a lauric acid group (C12), Tween 40 has a pahnitic acid group (C16), and Tween 80 has an oleic acid group (C18-1). In general, the shorter the chain length, the lower the growth rate (see Figure 10). However, for the Tween surfactants, most cultures obtained approximately the same final cell density. It has been suggested that longer chain lengths slow surfactant diffusion into the cellular membrane and thereby result in lower toxicity (Schwuber, 1980). Tween surfactants have also been found to be non-toxic to other types of culture systems (e.g. soy and carrot suspension cultures) and in some cases Tween 20 actually enhanced the 56 growth of the culture (T.T. Ames, unpublished data). Higher surfactant concentrations (10 to 20 DSC) were found to be toxic to the plant suspension cultures. There was no inhibition from the products formed at the concentrations seen in this study (Grethlein, 1990). The deviations from typical exponential growth are believed to be due to mass-transfer limitations of the gaseous substrate. 5.3.3 Brij Surfactants Figure 14 shows the growth of B. methylotrophicum in the presence of Brij surfactants. The control experiments were performed in the same manner as the Tween run. Similarly, the data in Figure 14 are an average of triplicate runs. Figure 15 shows the specific growth rate as a function of Brij surfactant concentration. 1.. —a—Control +Bri, 52, DSC=1 e 0-9 " +36, 52, DSC=2 o 0 8 +36, 52, DSC=3 - ,. .~ 4’ . .. . o. .36, 56, DSC=1 Ir - - - o- .Brij 56, DSC=2 3 B: 0.7.. .x..Bri, 56, DSC=3 .' V +36, 58, DSC=1 _ ,' S 05 +Bri, 58, DSC=2 ”a g ' .. +Bri, 58, DSC=3 " h *5 0.5.. 0 2 04 o . 'P o = 0.3.. 0 o 0.2 .. '1 A ,, ___.-o 0.1.. , z; A .--'c-> _ . MA.:‘2:'.A """" fi---A.‘-A-"A-— A O l . ‘ l l 1 I 0 50 100 150 200 250 300 Time (hours) Figure 14: Growth of B. methylotrophicum in batch bottles with Brij 57 0.05 .. 0.045 .. 0.04 .. 0.035 ‘4 0.03 .. 0.025 .. 0.02 .. 0.015 .. Specific Growth Rate (h") -e- Brij 52 -¢— Brij 56 0.01 .. 0.005 4 +30 58 l I I I 1.5 2 2.5 3 3.5 Dimensionless Surfactant Concentration 0 i e 0 0.5 1 Figure 15: Specific grth rates of B. methylotrophicum with Brij The Brij surfactants did strongly inhibit growth in some cases, but the inhibitory effect again depended on the chain length. The Brij surfactants are polyoxyethylene (polyethylene glycol -PEG) alcohols (see Figure 10) and had the following characteristics (in order of increasing chain length): Brij 52 - PEG(2) Cetyl alcohol, Brij 56 - PEG(10) Cetyl alcohol, and Brij 58 - PEG(20) Cetyl alcohol. As shown in Figures 14 and 15, the specific growth rate and final cell densities are much lower than those for the control for increasing amounts of the two longer chain surfactants, Brij 56 and Brij 58. In general, the higher the surfactant concentration the lower the specific growth rate and final cell density. The Brij surfactants are apparently more inhibitory to the growth of B. methylotrophicum in concentrations necessary to form microbubbles than are the Tween surfactants. 58 Growth experiments were also done with a few ionic surfactants (e.g. sodium dodecyl sulfate, cetyl pyridinium choride, and sodium dodecyl benzene sulfonate). The growth of B. methylotrophz'cum was strongly inhibited by these surfactants even at the lowest concentrations tested (0.5 DSC) (data not shown). Both anionic and cationic surfactants strongly bind proteins. The charged groups of ionic surfactant form relatively strong ionic bonds with charged groups on the proteins. Non-ionic surfactants, on the other hand, lacking the charged groups, bind through much weaker hydrophobic interactions with the protein chain and thus are less likely to inactivate or denature enzymes and proteins (Schwuber, 1980). Non-ionic surfactants would thus be expected to have the least influence on cellular metabolism in fermentation. 59 5.3.4 Products and pH The acetate, ethanol and butyrate fermentation profiles are shown in Figures 16, 17 and 18 respectively as a function of surfactant type and chain length. 1.6 .. —a—Dontrol —o—Tween 20, DSC=1 1'4 " —e—Tween 20, DSC=2 +Tween 20, DSC=3 1.2 + - o- .Tween 40, DSC=1 - o. .Tween 40, DSC=2 - A- .Tween 40, DSC=3 1 -- +Tween 80, DSC=1 —o—Tween 80, DSC=2 +Tween 80, DSC=3 0.8 .. 0.6 -r 0.4 .. Acetate Concentration (glL) 0.2 . db - 0 50 100 150 200 250 300 350 Time (hours) Figure 16: Acetate production 60 Ethanol Concentration (glL 0.18 0.16 4. 0.14 .. 0.12 .. 0.1 .. 0.08 .. 0.06 . 0.04 . 0.02 . —e—Control +Tween 20, DSC=1 —e—Tween 20, DSC=2 +Tween 20, DSC=3 - o- Tween 40, DSC=1 - o- .Tween 40, DSC=2 - a, Tween 40, DSC=3 —o—Tween 80, DSC=1 +Tween 80, DSC=2 +Tween 80, DSC=3 150 200 250 Time (hours) Figure 17: Ethanol production 61 0.35 1F —a—Control —e—Tween 20, DSC=1 0.3 .. —e—Tween 20, DSC=2 —¢—Tween 20, DSC=3 - o- -Tween 40, DSC=1 0.25 .. - o- -Tween 40, DSC=2 - A. .Tween 40, DSC=3 —o—Tween 80, DSC=1 G 0.2 i. +Tween 80, DSC=2 +Tween 80, DSC=3 0.15 .. 0.1 HP Butyrate Concentration (gIL 0.05 . 150 Time (hours) Figure 18: Butyrate production The remainder of the product concentration data (carbon dioxide) was calculated using carbon and electron balances for each run (Erickson, 1978). The data shown in Table 9 for Tween surfactants, represent the average of triplicate values. The abbreviations in used in Table 9 are summarized as the following; Ace indicates acetate, Et indicates ethanol, Bu indicates butyrate, CM indicates cell mass, and DSC indicated the dimensionless surfactant concentration. 62 Table 7: Carbon and electron balance results for Tween surfactants Control 4 CO —> 0.40 Ace + 0.076 Et + 0.065 Bu + 0.61 CM + 2.17 CO2 Tween 20 4 CO —> 0.48 Ace + 0.050 Et + 0.028 Bu + 0.70 CM + 2.12 C02 (1 DSC) Tween 20 4 CO —> 0.43 Ace + 0.114 Et + 0.039 Bu + 0.58 CM + 2.18 C02 (2 DSC) Tween 20 4 CO ——> 0.45 Ace + 0.163 Et + 0.012 Bu + 0.52 CM + 2.20 C02 (3 DSC) Tween 40 4 C0 —+ 0.41 Ace + 0.083 Et + 0.057 Bu + 0.61 CM + 2.17 C02 (1 DSC) Tween 40 4 CO —> 0.34 Ace + 0.076 Et + 0.079 Bu + 0.66 CM + 2.19 C02 (2 DSC) Tween 40 4 CO —) 0.37 Ace + 0.068 Et + 0.082 Bu + 0.62 CM + 2.18 C02 (3 DSC) Tween 80 4 CO ——> 0.43 Ace + 0.085 Et + 0.047 Bu + 0.62 CM + 2.16 C02 (1 DSC) Tween 80 4 CO ——> 0.43 Ace + 0.063 Et + 0.031 Bu + 0.77 CM + 2.13 C02 (2 DSC) Tween 80 4 C0 —+ 0.37 Ace + 0.096 Et + 0.078 Bu + 0.55 CM + 2.20 C02 (3 DSC) A typical plot of pH as a function of time during the experiment is shown in Figure 19. The pH profiles for other runs (data not shown) had the similar profiles to the one presented here. 63 7.2 -. 7 6.8 . 6.6 - I a. 6.4 . 6.2 .. -S—Control +Tween 20, DSC=1 6 .. -e—Tween 20, DSC=2 —¢—Tween 20, DSC=3 5.8 : i ' ' i i : 0 50 100 150 200 250 300 350 Time (hours) Figure 19: pH profile of a Tween 80 fermentation The pH decrease as the fermentation proceeded is due to the production of acetic and butyric acids. The metabolic pathways for acid production have been well- characterized (Grethlein, 1992b). A shift in the product ratios for B. methylotrophicum has been observed, in which less acetate and more butyrate and alcohols are produced at lower pH (Grethlein, 1992b); (Grethlein, 1990). Moreover, a pH shift from 6.8 to 6.0 as the cells entered the stationary phase has been shown to induce formation of butyrate as the primary product (Grethlein, 1991a); (Worden, 1989). Consistent with this trend, Figures 12 and 16 show that acetate is produced throughout the fermentation, but the butyrate production starts after 70 hours. The acetate, ethanol and butyrate profiles in Figures 16, 17 and 18 are the averages of triplicate data sets. In the averaged sets of data the variability was as great as +/- 25% between bottles. Due to the post-autoclave 64 additions done to each bottle and individually inoculating each bottle, these variations are not unexpected. The carbon and electron balance results for the Tween surfactant experiments, shown in Table 7, show the ratio of the primary products, acetate, ethanol, and butyrate, were not affected by the presence of the surfactant in the media. This result suggests that the Tween surfactants are a good choice for microbubble-sparged CO fermentations. 5.4 Conclusions The effect of Tween surfactants on growth of B. methylotrophicum on C0 varied With the chain length. Longer chain lengths had negligible effects on growth in a DSC range of 0 to 3. Shorter chain lengths slowed growth somewhat, but did not affect the final cell density. None of the Tween surfactants significantly affected the product stoichiometry. Brij surfactants were more inhibitory over the same concentration range, in some cases reducing both the growth rate and the final cell density. An increase in the ratio of butyrate to acetate observed late in the fermentations coincided with a decrease in the fermentation pH and the onset of the stationary phase. 65 6. Mass Transfer Properties of Microbubbles 6.1 Introduction The commercial feasibility of synthesis-gas fermentations is limited by the: relatively low volumetric productivities. Several studies have indicated that the rate limiting step in these fermentations is the inherently slow transfer of synthesis gas int the liquid phase (Worden, 1997). The driving force for synthesis-gas mass transfer is lovr on a mass basis, the aqueous solubilities of CO and H2 are only 60% and 4% that c oxygen, respectively. Moreover, per carbon equivalent consumed, more moles c synthesis gas must be transferred than for an aerobic fermentation. Consequently, th well-known gas mass-transfer issues that are central to bioreactor design for aerobi fermentations are even more critical for synthesis-gas fermentations. A common approach used to enhance gas-to-liquid mass transfer is to increase th impeller rate. The resulting increase in average shear rate enhances bubble breakuj increasing the interfacial area for mass transfer. Although effective, this approach i costly for large ferrnenters, because power consumption is proportional to impeller rate t the third power and impeller diameter to the fifth power (McCabe, 1985). In addition, th increased power input can be detrimental to shear-sensitive cells. Sparging with microbubbles is another approach that has the potential to enhanc mass transfer while maintaining low power consumption and shear rates. Microbubble offer the potential for orders-of-magnitude higher mass-transfer rates due to the sma‘ size and higher interfacial areas. Microbubble mass transfer should be most effective fc gases having a high consumable fiaction, such as synthesis gas, which can consist of ove 95% CO and H2 (Worden, 1998). As the consumable gas is transferred into the liqui 66 phase, the microbubble shrinks, further increasing the interfacial area per unit gas volume. Additionally, the gas pressure inside a bubble is greater than that outside due tc the surface tension. The magnitude of this pressure differential, known as Laplace pressure, increases as the microbubbles shrink (Rosen, 1989). Because the gas solubility is proportional to the intrabubble pressure, the Laplace pressure increases the driving force for mass transfer. The lack of electrodes to measure liquid-phase, carbon-monoxide concentration: suggests the use of oxygen as a model gas for the mass-transfer studies. The high rate 0: mass transfer from microbubbles makes measurement of volumetric mass-transfer coefficients (KLa) challenging. Unsteady-state methods that use oxygen electrodes tc measure changes in liquid-phase oxygen concentration as a function of time are poorl) suited, because the response times of the probes are too slow to follow the rapic dynamics of the liquid-phase oxygen concentration (V an't Riet, 1979). Sulfite-oxidatior methods are also poorly suited due to significant rate of reaction that can occur in the liquid film surrounding the slowly rising bubble (Bailey, 1986). Consequently, a methoc based on measuring the steady-state oxygen concentration profile through a bubble column was developed. A mathematical model was developed to predict changes in the bubble diameter across the bubble column, so that the mass-transfer coefficient (KL) am the interfacial area per unit volume (a) could be computed individually. The effects or gas void fraction and the concentration of surfactant in the bulk liquid on KL were determined. Power requirements to produce the microbubbles were experimentally measured and used to estimate the power required for microbubble sparging of a large- scale, synthesis-gas fermentation. This chapter describes the experimental studies tc 67 determine the mass transfer properties of microbubbles and the power consumption necessary to produce them. 6.2 Materials and Methods 6.2.1 Microbubble Generation The microbubble dispersions were formed using a spinning-disk apparatus described in Chapter 2. Microbubble size distributions were measured by laser-light scattering using a Malvem Mastersizer, also described in Chapter 2. The surfactant used in all studies was Tween 20 (Sigma Chemical Co., St. Louis, M0) at two times the critical micelle concentration (CMC) (120 mg/L). This surfactant was chosen from the studies done in Chapters 4 and 5. The cost of the surfactant was not a major consideration in the selection due to the low concentration used and the typically low cost of the common polysorbates. The microbubbles were formed with 100% oxygen for the axial- dispersion and mass-transfer studies. 6.2.2 Axial-Dispersion The degree of axial dispersion of the water/microbubble dispersion passing through the bubble column was measured to determine a suitable mixing model for calculating the Km values. The experimental system, shown schematically in Figure 20, consisted of a 60-cm-long by 2.5-cm-ID glass column. The column had ports at 7 cm, 22 cm, 37 cm, and 52 cm along its length. A 1/30 hp centrifugal pump was used to supply the carrier liquid (distilled water). A peristaltic pump delivered microbubbles from the generator to the column. Standpipes were used to reduce pressure fluctuations from the peristaltic pump. 68 OUtlet <——-]— Surfactant Solution (I. Oxygen Gas 1 L Carrier )1 ‘ Water / .6" Vi. 1: /...————:a‘ :,\ l“; K ’ \\\ :K‘rjg/ Figure 20: Experimental apparatus used in axial dispersion and mass transfer Two conductivity probes were made by mounting a 1 mm diameter brass roe inside of a 2-mm-ID brass tube with a 0.5-mm-thick plastic, non-conductive spacer ir between. Wires soldered to the ends of the inner rod and outer tube were connected te conductivity meters (Cole Partner, Chicago, IL). The ends of the probes were filed fla and polished (Thompson, 1993). The probes were mounted in ports 7 cm and 52 en above the base of the column. Five mL of a 25 g/L CaCl2 tracer was injected into the bull liquid stream just upstream of the column. The probes’ responses were recorded on the hard disk of a PC at 20 Hz using LabTech Notebook software and a Data Translation I/C 69 card. Voltage isolators were used to eliminate a current loop between the probes. Between 5 and 10 replicates were done for each set of conditions. 6.2.3 Mass Transfer A Clark-style oxygen electrode (Diamond General, Ann Arbor, MI) attached to a Chemical Microsensor II (Diamond General, Ann Arbor, MI) was used to measure the steady-state, liquid-phase oxygen concentration at several positions in the column. The electrode was calibrated at 0% oxygen with nitrogen-sparged water and at 100% with oxygen-sparged water. The water was maintained at a constant temperature of 22 °C in a water bath during the calibration. The bulk liquid phase was initially degassed to prevent stripping of nitrogen gas into the microbubbles by maintaining distilled water under a vacuum created by a water aspirator for 24 hours prior to the experiment. The degassed water gave an oxygen reading of 0% using the calibrated electrode. The surfactant was added to the bulk liquid prior to degassing. Experiments were run at several flow rates and surfactant concentrations. Between three and five replicates were performed for each set of conditions. 6.2.4 Power Consumption The power consumption of the microbubble generator was measured using a Lightnin mixer, model L1U08 (Lightnin, Rochester, NY), that gives simultaneous readouts of rpm and power input. The power input for the generator was measured at the maximum mixing rate of the mixer (1800 rpm) in air, water, and microbubble dispersions. The volumetric-uptake rate of synthesis-gas components was obtained from the continuous fermentation. 70 6.2.5 Dynamic Gassing Out Experiments were done in a 100 gallon fermenter vessel to estimate the mass transfer coefficient from microbubble sparging using a dynamic gassing out method. The fermenter was a Pfaudler vessel equipped with Rushton-bladed turbines for mixing. The vessel diameter was 3 feet with an impeller diameter of 1 foot. The vessel also had a sparger for the delivery of compressed air or nitrogen from a tank. The impeller was driven by an overhead drive and was calibrated using a strobe gun. Since typical oxygen probe response times were felt to be inadequate to handle the rapidly changing liquid- phase oxygen concentration, conditions were set so that the liquid-phase concentration would not change as rapidly. The use of a large-scale fermenter and low gas flow rate were kept the liquid-phase concentration from changing rapidly. The oxygen electrode (Ingold) was attached to an amplifier/interface and then to a data acquisition system (Camile Data Acquistion). The microbubbles were made from air and were sparged to the fermenter through silicon tubing. The microbubbles were made in an aqueous, 2 DSC solution of Triton X-100 that was refilled into the generator through a peristaltic pump. The oxygen electrode was calibrated at 0% saturation after sparging the liquid phase with nitrogen gas, and 100% saturation after sparging with air at the beginning of each day of experimentation. The liquid inside the vessel was changed between each run with tap water. The temperature was monitored on several occasions during the run and was found to remain relatively constant at 14 °C. Gas holdup was estimated by immersing a flask of known volume into the vessel, capturing a known volume of liquid/ gas dispersion, and allowing the gas phase to settle out. The experiment was performed by stripping the oxygen concentration down to 0% with nitrogen and then 71 starting the flow of microbubbles to the system. The liquid-phase oxygen concentration was monitored with time. 6.2.6 Mathematical Models 6.2.6.1 Axial Dispersion The tracer peak for each probe was normalized using a calibration curve. If the areas under the input and response peaks differed by more that 15%, that data set was discarded. An unsteady-state mass balance on the tracer in the liquid phase is shown below, along with the appropriate boundary conditions for an open vessel. ac, a"- C, ac, _ Dar 2 — u—— (11) 81‘ 52 é’z t=0 z=0 CL=C0 t>O z=0 CL=0 t> 0 z = 00 CL = 0 This model is based on the assumptions that the salt tracer exists only in the liquid phase and that the salt concentration is uniform. in the radial direction. A computer program originally developed by Wakao and Kaguei (1982) and later modified by Thompson (1993) was used to determine the optimal values of the dispersion coefficient and interstitial velocity for each data set by minimizing the least-squares error. The program converted the experimental data into Fourier series and calculated a predicted outlet tracer curve using the Fourier series from the input probe (at 7 cm) as an input function and the transfer functions from Equation 11. 72 6.2.6.2 Mass Transfer The gas phase was assumed to pass through the column in plug flow. This assumption is typically satisfactory in systems having a large length-to-diameter ratio (F ogler, 1992). The length-to-diameter ratio for the bubble column is 24. A steady-state, liquid-phase mass balance on oxygen assuming plug flow is shown below, along with the appropriate boundary conditions. dCL KLa . ———— C —C =0 dz u ( L L) (12) The interfacial area per unit gas volume is given by: 3R280 (13) a : R3 where the gas void fraction varies with the bubble radius according to the equation: 3 R 50 = 50,0(R_0] (14) The interfacial oxygen concentration varies with the atmospheric pressure, depth in the column, and bubble radius as shown below: C; = H[p(—g—](L — z) +P0 + 2—0] (15) g. R The rate of microbubble shrinkage varies significantly across the length of the column and is described by the following equation and boundary conditions. 73 3K . RT 4in— L(c,_c,) g dR gc u M g 40' 3P0 + 2(L — z)p(—) + — gc R (16) z = O R = R0 This equation accounts for hydrostatic pressure, Laplace pressure, and loss of gas via transfer into the liquid phase. Equations 12 through 16 were solved using Simusolv modeling software (Dow Chemical Company, Midland, MI) using a Gears stiffness method with a generalized reduced-gradient optimization on KL. The average value of R of 60 um was determined with the dynamic light scattering method (Chapter 4) and used, together with the flow rates of the degassed water and microbubble dispersion, to calculate the initial value of a. 6.2.6.3 Dynamic Mass Transfer An unsteady-state mass balance for oxygen on the liquid phase, assuming a well- rnixed gas phase for a CSTR is shown below. dc, (t) dt = K,a[—CI7G — C, (1)] (17) which can be integrated to give a linear form shown below. ln[l — :L j = —KLat (18) * L Equation 18 can be plotted to give a slope of —KLa. 74 6.3 Results and Discussion 6.3.1 Axial Dispersion The dispersion coefficients for the microbubble dispersions in the bubble column are shown in Figure 21 as a function of liquid velocity and microbubble gas flow rate. The error bars represent the average deviation from the mean of the replicates. The literature correlation for a conventional bubble column given below (Deckwer, 1974) is also shown. 1),, = 2.7D}-4u§-3 (19) 25 20 ul- — Deckwer et. al. a 7.22 mI/min A 12.73 ml/min o 26.5 ml/min 0 40.26 ml/min x 54.0 ml/min + 67.8 ml/min _\ (II I I _I. O I Dispersion Coefficient (cmzls) t. 4M 6 8 10 12 14 16 Velocity (cm/s) Figure 21: Axial dispersion coefficients Figure 21 indicates that axial dispersion for the microbubble dispersions was considerably lower than in conventional bubble systems. In conventional systems, the much larger bubbles rise at an appreciable rate relative to the surrounding liquid. Liquid 75 trapped in the bubble wake is transported through the column at a higher velocity than the bulk liquid, contributing to axial mixing. The very low rise velocities of microbubbles do not induce such wakes. The dispersion coefficients were used to calculate Peclet numbers. The lower bound on the Pe values was 42, which is indicative of low degrees of axial mixing (Fogler, 1992). Consequently, the plug-flow mixing model was deemed appropriate for calculating KL values from the oxygen-profile data. 6.3.2 Mass Transfer 6.3.2.1 Plug Flow Model Figure 22 shows an example of the agreement between the experimentally measured liquid-phase oxygen concentration profile across the bubble column and that predicted by the mathematical model for the optimal KL value. The error bars represent the average deviation from the mean of the replicates. The predicted change in R across the bubble column is also shown for this run. Even though the residence time of the microbubbles in the column was only a few seconds, R, CL”, 80, and a all are predicted to change significantly, demonstrating the need to include Equations 13 - 16 in the model. 76 3.50E-02 3.50E-05 A 3.00E-02 .. 3.00E-05 15 CD 5’ 2.50E'02 -- -- 2.50E‘05 A r: .5. .2 en ‘5' 2.00E-02 .. .. 2005-05 .2 J:- "o N 5 a: 8 1.50E-02 .. ..1.50E-05 2 O .D o '3 S: 1.00E-02 .. —Optimized Fit .. 1.00E-05 m >. A Experimental Data X . 0 5005-03 .. -*- Rad'us .. 5.00E-06 0.00E+00 ; - . - : 0.00E+00 0.00 0.10 0.20 0.30 0.40 0.50 0.60 Column Position (m) Figure 22: Optimized fit for KL using a plug flow model 6.3.2.2 Void Fraction and Surfactant Effects The experimentally determined KL values are shown in Figure 23 as a function of so, for different concentrations of surfactant in the bulk liquid. Also shown in this figure are KL correlations published by Waslo and Gal-Or (1971), and Calderbank and M00- Young (1961) for small, rigid—surface bubbles. The well—known theoretical correlation for diffusion from a sphere into an infinite pool of stagnant fluid (Sh=2) is also shown. 77 0.00025 0 0x Tween 20 a 1x Tween 20 0-0002 1- A 5x Tween 20 —.—Calderbank and Moo-Young —Waslo and Gal-Or 0.00015 .. 0 —--3h = 2 KL (mls) C vvvvvvrr-vv vvvvvvvtvvv vvvvvvvvivvvvvvvvvvvvvvvvv‘v’v vvvvvvvv l .- \U \ 0.00005 . 0 0.01 0.02 0.03 0.04 0.05 0.06 0.07 Gas Void Fraction, 89 / Figure 23: Mass transfer coefficients from microbubbles The experimentally determined KL values are similar in magnitude to those predicted by the correlations. However, there is a significant decrease in KL with increasing surfactant concentration. This trend suggests that adsorption of the surfactant at the gas-liquid interface creates additional resistance to gas mass transfer. Such resistance can arise through several mechanisms. First, surfactant molecules can retard surface flow by the surface tension gradient at the interface. Second, they can act as physical barriers to the passage of gaseous molecules at the interface (Koide, 1976). For soluble surfactant monolayers, Princen et al. (1967) suggests that gas transfer occurs by simple Fickian diffusion through aqueous pores between the surfactant molecules. Third, the surfactant concentration may influence the thickness (h) of the liquid shell surrounding the bubble. Amiri et al. (1990) estimated the shell of microbubbles to be 78 about 0.75 pm thick. Assuming mass transfer is limited by the rate of diffusion across the shell, h can be estimated. from KL using the following expression (Princen, 1965). DOZHO h kL,S (20) For the range of KL values shown in Figure 23, the range of h values predicted by this equation is 0.2 pm to 3 pm. This calculation assumes the diffusivity of oxygen in the shell is equal to that in pure water. The presence of soluble surfactants has been reported not to significantly affect diffusion of small molecules (Quintana, 1990). Also, the gas- side fihn resistance has been shown to be negligible for a sparingly soluble gas, such as oxygen (Motarjemi, 1978). The effect of surfactants, gases, and composition on the thickness of the water shell has not been determined to this date. The effects of these conditions could be addressed at some later date. The KL values measured when no surfactant was initially present in the bulk liquid were higher than the values predicted by the three correlations based on the data presented in Figure 23. One explanation for this trend is that bubble shrinkage enhances mass transfer relative to conventional bubbles through the Laplace pressure effect and a steepening of the oxygen concentration gradient outside the bubble. These ideas are explored more thoroughly using a dynamic model of a single microbubble (Worden and Bredwell, 1998). The KLa values for the experiments shown in Figure 23, calculated using Equation 21, ranged from 200 h'1 to 1100 h". KLa = -—gKL (21) 79 Table 8 contains the KL, a, and KLa values determined in this study using the bubble column. Table 8: Mass Transfer Results Surfactant Microbubble Liquid Flow KL a KLa (DSC) Flow (mL/min) (m/s) (m'l) (11") (mL/min) 0 39.8 1033.6 1.1426e-4 2531.8 1041.22 0 39.8 1810.4 1.4717e-4 1453.1 769.9 0 39.8 2587.1 2.1418e-4 1019.1 785.8 0 81 1033.6 5.2524e-5 5091.8 962.8 0 81 1810.4 7.7295e-5 2939.1 817.8 0 81 2587.1 1.1925e-4 2065.8 886.9 1 39.8 1033.6 6.0244e-5 2531.3 549.0 1 39.8 1810.4 7.5274e-5 1453.1 393.8 1 39.8 2587.1 1.2353e-4 1019.1 453.2 5 39.8 1033.6 2.9325e-5 2531.3 267.2 5 39.8 1810.4 4.3340e-5 1453.1 226.7 5 39.8 2587.1 5.5200e-5 1019.1 202.5 These values are considerably higher than those reported for mechanically agitated, commercial-scale fermentations, yet they were measured in an unagitated bubble column. Thus, despite the reduction in KL seen at high surfactant concentrations, the large increase in a due to the exceedingly small bubble diameter resulted in high KLa values. 80 6.3.3 Power Consumption 6.3.3.1 Power Consumption in Fermenters The measurements of power required to drive the microbubble generator in w: and in the nricrobubble dispersion at 1800 rpm were used to calculate a Power numbe: 0.036 for the microbubble generator. Power numbers are typically constant in turbulent regime (Rushton, 1950). Consequently this value could be used to calculate expected power consumption at the operating rpm (4000 rpm) from the defining equat for Power number, shown below. Pg. PM = pN 1’ D? (2: The measured rate of microbubble formation in the generator (300 mL/rnin of gas) 1 used, together with the volumetric CO consumption rate for B. methylotrophic (0.02573 mol/h*L) (Grethlein, 1991b), to calculate a power requirement of 0.01 kW m3 of fermentation broth to form the microbubble dispersions. It was assumed that power number was constant in the turbulent regime, that the power consumption in foam was 1/3 of the power consumption in the liquid, and that there was 10 consumption of synthesis gas components (CO) from the microbubble. A small amo of additional power would be required to maintain adequate liquid mixing in the reac In comparison, typical P/V values for commercial-scale, stirred-tank ferrnenters are the order of 1 kW/m3 (Van't Riet, 1979). An estimated P/V value for synthesis-. fermentations using conventional sparging in stirred-tanks is 2 kW/m3 (E.N. Kaufm personal communication). 81 6.3.3.2 Comparison of Reactor Type The dependence of KLa on P/V has been well—studied for conventional bubble For stirred tanks, a commonly used correlation is (Van't Riet, 1979) P KLa = K(?) (ug )fl (23) For bubble-coalescing media, or = 0.4 and 0:05; for non-coalescing media, 0t=0.7 and = 0.2 (V an't Riet, 1979), where P is in W/m3, V is in L, and ug is in m/h. For airli reactors a comparable correlation is (Margaritis, 1981), (Siegel, 1988). KLa = [(9 0(3)!) (24) Reported values of 0t and B are 1.0 (Siegel, 1988) and 0.9 (Bello, 1985), respectively. Tb correlations shown in Equations 23 and 24 (for a superficial gas velocity of 60 cm/mir are plotted in Figure 23, along with the range of KLa values measured in this study, t show the performance enhancement possible using microbubble sparging Comparatively, the superficial gas velocity for the microbubble study ranged from 180 t 500 cm/min. Relatively high power inputs are needed in conventional gas-sparge systems to break up bubbles and provide adequate KLa values. However fc microbubbles, only enough power is required to provide adequate liquid mixing. Thu: microbubble sparging offers the possibility of simultaneously providing both high KL values and low power inputs. The energy savings arise because the high shear levels use to provide bubble breakup are applied only to the small liquid volume contained in th microbubble generator, rather than the entire contents of the bioreactor. The only powe delivered to the liquid contents of the main bioreactor is that needed for mixing. In large scale fermenters, the microbubbles could be generated in an external vessel as in thi 82 study. Increased filtration capacity would then be required to handle the larger scai liquid entering the reactor through the microbubble foam. Alternatively, the microbu generator could be housed directly in the larger fermenter to create a local “microbubble generation zone” within the fermenter, thereby removing the cost of 12 scale filtration necessary to regenerate the solution to make microbubbles. 1 00000 10000 . 1000 .. Microbubble Range X, !? x/X’ .: x/X/ v x/ 1 1 i—+.- “h x/‘r 4' ' I -“ 100 4r/W x 10 -o—CSTR - Bubble Coalescing " _i_csrR - Bubble Non-coalescing —x—Airlift - with draft tubes (Margaritis et. al.,1981) 1 I I I I I I I I :I 100 1‘ (PN) (W/m3) Figure 24: Comparative KLa values for varying reactor types The simultaneous advantages of reduced power and increased producti demonstrated in synthesis-gas fermentations could, in principle, also be realized variety of other unit operations in which gas-to—liquid mass transfer is rate limit Examples include aerobic fermentations, soil washing, stripping of volatiles from a lie Phase, and flotation. 83 6.3.4 Dynamic Gassing Out A sample of the raw data obtained from the absorption of oxygen into nitrogen- stripped water is shown in Figure 25 below in terms of percent air saturation. The temperature remained constant at 14 OC throughout the experiment. The probe was calibrated on a daily basis and fresh tap water was used for each run. 60- 50. 40. 304 °/o Air Saturation 20. 10. '— 400 600 800 1000 1200 1400 Time (sec) 0 200 Figure 25: Experimental data from a dynamic gassing out run C . . . . The data were plotted as the ln[l — I; J versus trme, as shown 1n Equation 19 to obtain a L KLa value. A sample of these plots is given in Figure 26. 84 200 400 600 800 1000 1200 -0.1 . I '0.2 III- y = -0.000638x + 0.026814 2 _. _0.3 __ R — 0.999961 -0.4 .. ‘0.5 III- ‘0.6 II- -0.7 .- Time (sec) Figure 26: Linear form of Equation 19 To evaluate whether the assumption that the composition of the gas phase changes as function of time in the reactor, a calculation was done to compare the maximum amount of oxygen gas that could be transferred to the actual amount of gas that was transferred. The maximum amount of gas transferred was determined by calculating the total amount of oxygen coming into the reactor and assuming it was all transferred into the liquid phase. Other assumptions made include that the gas behaves as an ideal gas and the pressure inside of the bubble is atmospheric. The results from this comparison are shown for a feed of 300 mL/min gas phase from microbubbles at 50 RPM in Figure 27. 85 10-. .- 100 9 .. .. 90 d 8 .. .. 80 3’ V 7 -I- III- 70 U c o .2 t a 6 .. -- 60 £2 4: 8 5 5 . .. 50 ea 5 g I- o 4 .. 40 g 0 o S: 3 . .. 30 g Q 2 _ _ —Theoretical .. 20 0 / x Experimental 1 -- ——Percent Transferred -- 10 0 i i i 4 0 0 10 20 30 40 50 Time (minutes) Figure 27: Percent of oxygen transferred to liquid phase The results in Figure 27 indicate that for such small bubbles, there is a considerable fraction of gas transferred into the liquid. The maximum percentage of inlet oxygen transferred to the liquid phase was 86% at 300 mL/min, 60% at 600 mL/min, and 50% at 900 mL/min. This indicates that the assumption that the gas phase concentration in the reactor did not change is most likely not valid. To properly account for the changing composition of the gas phase, accurate data regarding the gas holdup of the reactor and the outlet gas composition would be needed. Using such data, gas-phase and liquid-phase mass balances could be written using the well-mixed liquid phase balance in Equation 18 and a well-mixed gas phase mass balance given below. The model could then be subsequently solved to give a KLa value from this type of experiment. 86 dCG Cl—C, . V, = —K C —C — 25 dt (VG/G) ‘al ‘ JV, ( ) where C1 and C2 are the gas-phase oxygen concentrations at the reactor inlet and outlet, respectively. It would be necessary to use an average gas flow rate if the inlet and out flows were not constant. It would also be necessary to use an average value for CL. since there will be bubbles with a wide variety of oxygen gas compositions. Van’t Reit (1979) has suggested that the data from making the assumption of a constant composition gas phase can become a significant problem if the gas phase residence time is not much smaller the l/KLa. In the experimental system used the gas phase holdup was only roughly estimated. A gas void fraction of 0.0303 was measured for a 300 mL/min flow rate of microbubbles into the reactor. The gas void fraction did not seem to vary significantly for different sparging rates or different measurement times past 15 minutes. Using the estimate of gas holdup, the gas phase residence times for the three different flow rates, 300 mL/min, 600 mL/min, and 900 mL/min were 38.4 min, 19.2 min, and 12.8 min respectively, and were obtained by dividing the reactor gas holdup by the inlet gas flowrate. Comparatively, the 1/KLa values ranged from 20.5 min to 50.3 min. These numbers verify the results seen earlier that the gas-phase dynamics are important to consider in this situation and should be included in the analysis of the data and determination of KLa. Dunn and Einsele (1975) have constructed correction charts for the determination of more accurate KLa values where only the gas void fraction and the KLa determined from the liquid-phase model are known. The correlation charts were constructed from plotting the relationship between KLa determined using only liquid-phase dynamics and the KLa determined from using both liquid-phase and gas-phase dynamics at different 87 gas-phase residence times. Extrapolating the gas-phase residence times, the correction chart given in literature (Dunn, 1975), the corrected KLa values were determined. The results are presented in Figure 28, along with the correlation describing mass transfer coefficients in a bubble non-coalescing media (Van't Riet, 197 9). 20 .- 18 .. 16.. 14. +Bubble Non CozaTes—Eingj +300 mL/min 1 +600 mL/min l KLa (h 1) (Corrected values) 8 4 J +900 mL/min 2 .. 0 i : i + i i 0 100 200 300 400 500 600 (PN) ratio (W/m3) Figure 28: Dynamic mass transfer coefficients of microbubbles The power-to-volume input was estimated by calculating the impeller Reynolds number, Re, from Equation 26 and determining the power number, Pno from correlation charts (Bailey, 1986). The impeller Reynolds number was calculated to be 700, and the subsequent power number was determined to be 3.9. The system was operated in the transition region between the laminar and turbulent flow regimes (between a Re 50 and 5000) 88 _ pIIViDi2 #1 Re. i (26) The power was then calculated from Equation 22. The effect of the aeration rate on the power requirements was determined through calculating the dimensionless aeration rate, Na. N, = G MD; (27) The effect of aeration on the power requirements was small; N,l was never higher than 104, suggesting that there was no power loss due to aeration. Figure 25 indicates that for an equivalent P/V ratio mass transfer coefficients from microbubbles are higher than those for conventional bubbles described by the bubble non-coalescing correlation at lower power to volume inputs. (The literature correlation used a gas flow rate of 300 mL/min.) This trend. is similar to that seen in Figure 24. Increasing power input generally increases bubble breakup, giving smaller average bubble sizes, and this trend is seen in the literature correlation. Microbubbles, on the other hand, have very small diameters entering the vessel and are unlikely to break up further. Thus, the KLa values determined from microbubbles should be independent of the power input to the vessel. This trend has been seen previously in fermentations (Kaster, 1990). The small increase in KLa seen in Figure 25 may be due to greater surface mixing at higher (P/V). The KLa curves for microbubbles cross the KLa value determined from the literature correlation at a similar (PN) ratio as in Figure 24. 6.4 Conclusions A steady-state method was developed to measure KL and KLa values for microbubble dispersions. The KL values obtained were comparable to those predicted by 89 correlations for small, rigid-surface bubbles. An increase in the concentration of surfactant in the bulk liquid decreased the KL value by as much as 75%, presumably due to increased mass-transfer resistance in the rrricrobubble shell. The KLa values ranged from 200 to 1100 h], verifying that microbubble dispersions can provide extremely high mass-transfer rates. The Power number of the microbubble generator was measured to be 0.036. Based on this value, an incremental power-to-volume ratio to make microbubbles for a synthesis-gas fermentation was estimated to be 0.01 kW per m3 of fermentation capacity. Thus, the projected power input to microbubble sparged fermenters is low, with little power required beyond what is necessary to keep the liquid mixed. The mass transfer coefficients from microbubbles were also measured in a dynamic gassing out method in a 100 gallon fermenter. Difficulties in the measurement of the gas holdup and exiting gas composition did not allow for accurate measurements of KLa. Correlation charts were used to account for the change in gas composition with bubble residence time. The results from this dynamic study show similar trends when compared with the KLa values versus power input obtained in the steady-state experiments. 6.5 Nomenclature 6.5.1 Symbols (1 = Interfacial area/unit gas volume (m!) b = Constant for physical properties of the fluids and reactor geometry in airlift fermenters C1 = Gas—phase oxygen concentration at inlet (mg/L) 90 C2 C0 CL CL CT Dt D02 Dax DSC gc Ho KL kL,s KLa Gas-phase oxygen concentration at outlet (mg/L) Gas-phase oxygen concentration (mg/L) Liquid-phase oxygen concentration (mg/L) Equilibrium liquid-phase oxygen concentration (mg/L) Initial salt tracer concentration (g/L) Concentration of calcium chloride tracer in liquid phase (g/L) Tube diameter (m) Bubble diameter (m) Diffusion Coefficient (mZ/s) Axial dispersion coefficient (mz/s) Impeller diameter (m) Dimensionless surfactant concentration = surfactant concentration/critical micelle concentration Gravity constant (9.81 m/sz) Conversion factor Reactor inlet gas flow rate (mL/min) Surfactant film thickness (m) Henry’s law constant (Kg/m3 Pa) Ostwald coefficient of gas solubility Constant for reactor geometry in stirred tanks Mass-transfer coefficient (In/s) Mass-transfer coefficient of surfactant layer (m/s) Volumetric mass-transfer coefficient (s'l) 91 ZZZ” O: '11 R02 RC.’ Sh UG VG Column length(m) Molecular weight (g/mol) Impeller rate (rpm) Dimensionless aeration number Power input (W) Power number Atmospheric pressure (N/mz) Reactor inlet partial pressure of CO (atm) Reactor outlet partial pressure of CO (atm) Peclet number (radius) Bubble radius (m) Initial bubble radius (m) Gas constant Reaction rate of oxygen in the liquid phase Impeller Reynolds Number Sherwood munber Liquid velocity (m/s) Gas velocity (m/s) Time (sec) Temperature (C) Volume (m3) Reactor liquid volume (L) Reactor gas volume (L) 92 7.5.2 So 80,0 Position in column (m) Greek Symbols Exponent for the (PN) term in stirred tank and airlift correlations = Exponent for the gas flow rate term in stirred tank and airlift correlations = Gas void fraction = Initial gas void fraction = Bulk liquid density (g/m3) = Continuous phase density (g/m3) = Surface tension (N/mz) 93 7. B. ME T HYLOT ROPHIC UM FERMENTATIONS 7 .1 Introduction In the previous chapter, the mass transfer coefficients from microbubbles were measured to be very high, in the range from 240 h'1 to 1100 h]. This chapter describes experiments to determine whether the use of microbubbles would enhance synthesis gas fermentations. The obligate anaerobe, B. methylotrohicum is capable of consuming CO to produce liquid fuels and chemicals such as ethanol and acetate. These fermentations have previously been shown to be mass transport limited (Worden, 1989) and could potentially benefit form microbubble sparging. A dynamic model describing the mass transfer of nricrobubbles has been developed (Worden and Bredwell, 1998). This model predicts that for a gas with a high consumable fraction, such as synthesis gas, the mass transfer rate from a microbubble increases rapidly as the bubble shrinks. This effect is less significant for gases with a 20% consumable fraction, such as air, where the amount of bubble shrinkage is not as large. The increase in mass transfer due to shrinkage is partially due to increased interfacial area per volume and partially due to increased internal bubble, which increases the saturation concentration. These results indicate that synthesis gas fermentations could benefit more than aerobic fermentations from microbubble sparging. 7.2 Materials and Methods 7.2.] Media and Cultures B. methylotrophicum cells were grown in serum bottles as described in Chapter 5, and after one week were used to inoculate the reactor (5% by volume). A phosphate- buffered, basal media was used in the reactor with half the concentration of yeast extract 94 used by Grethlein (1991b). The composition of the enhanced phosphate buffered media is given in Table 9 below. Table 9: Composition of the enhanced PBB media Component Concentration NaCl 0.9 g/L MgC12°2HzO 0.2 g/L CaCl202H2O 0.1 g/L NH4C1 1.0 g/L Yeast Extract 1.0 g/L Trace Mineral Solution 50 mL/L Vitamin Solution 100 mL/L Phosphate Stock 25 mL/L 2.5% Na2S-9H2O Stock 25 mL/L 0.2% Resazurin 1.0 mL/L (NH4)2SO4 2.0 g/L The vitamin solution, trace elements solution, and phosphate stock are the same as described in Chapter 4, Tables 3, 4, and 5. 7.2.2 Reactor A synthesis-gas fermentation using Butyribacterium methylotrophicum was run in a 1.5 L Multigen fermenter (NBS, Edison, NJ) using a Minitan stainless steel filtration unit containing eight Durapore 0.2 pm filter plates (Millipore, Bedford, MA) for total cell recycle. A schematic diagram of the apparatus is shown in Figure 29. 95 Fresh Media v __ Reactor 1 i Microbubble _, ‘ _, , Filter I Generator Figure 29: Flow diagram for a B. methylotrophicum fermentation A peristaltic pump delivered fresh media to the reactor at a rate of 0.2 mL/min. The pH was controlled at 6.6 using a pH 4000 controller (NBS, Edison, NJ) with 6 N NaOH and 6 N H3PO4. The CO was sparged into the fermentation through a stainless steel fiit having a 10 um pore size with the flow rate being controlled through a needle valve (Cole Parmer, Chicago, IL). The impeller rate was 200 rpm during the entire fermentation. Gas flow rates were measured using a wet test meter (GCA/Precision Scientific, Chicago, IL). 7.2.3 Analytical Techniques Gas compositions were measured using a Perkin-Elmer Autosystems gas chromatograph equipped with a thermal conductivity detector (Perkin—Elmer, Norwalk, CT) and a Hayesep DB 30’ by 1/8” column (Alltech Assoc, Deerfield, IL). The flow rate 0f the carrier gas (He) was 30 mL/min. The inlet and detector temperatures were 110 OC. The temperature program was 40 0C for 6 minutes, followed by a ramp at 45 °C/min to 96 275 °C for 1.5 minutes. The biomass concentration was measured using optical density (OD) at 660 nm. A calibration curve was used to convert OD into dry cell weight. Liquid- phase products were measured using gas chromatography with a flame ionization detector with a 4% Carbowax 20M on Carbograph IDA 80/120 mesh 6’ by %” deactiglass column (Alltech Assoc., Deerfield, IL). The flow rate of the carrier gas (He) was 20 mL/min. The injector and detector temperatures were 220 °C, and the column was maintained at 180 °C for 7 minutes. 7.2.4 Models A mass balance on C0 in the gas phase in the reactor assuming an ideal gas phase, mass-transfer-limited operation, and perfect mixing in both the gas and liquid phases is shown in Equation 28 (Klasson, 1992). Pi: = K.a V‘Rgf PCO HG (28) 7.2.5 Carbon and Electron Balances Carbon and electron balances have previously been done on B. methylotrophicum fermentations (Grethlein, 1991b). They were used in the product analysis in Chapter 5, and are presented in more detail here. The carbon and electron balance equations are given in Equations 29 and 30, respectively. N150 =NC0 +NC02 +2Nace +2Neth +4Nbut +Ncells (29) out out yCONCO : yCOz NC02 +7H20NH20 +yaceNace +7erhNeth +ybu’Nbu! +766”ch6113 (20) con verr prod prod The reductance degrees, y, are given in Table 10 below. 97 Table 10: Reductance degrees Component Reductance Degree Yco 2 Ycoz 0 Yace 4 Yeih 6 7601 5 Yceiis 4.2 71120 0 The reductance degrees are from Erickson (1978). 7.3 Results and Discussion 7.3.1 Products The concentration of the fermentation products for both conventional and microbubble sparging in the reactor are shown in Figure 30. 98 .I. O N 0| , ‘ -_ 20 g, .1 i .. 3’ ' W s z .. 15 E 9 .9. Cell ’5 ° 8 3 .5. Acetate c g 1,. Butyrate -- 10 8 = + Ethanol *6 ° 5 U -. 5 e LLlEEEEéa n. . i : O 60 70 Time (Days) Figure 30: Fermentation product profile Similar concentrations of acetate, ethanol, and butyrate have been reported in previous CO fermentation studies using B. methylotrophicum (Grethlein, 1991b). The volumetric productivity for acetate, the primary product, was 3.84 g/L*day prior to microbubble sparging. Butanol is typically not produced to a significant degree at the pH used. The dip in the reactor productivity at day 26 was due to the installation of new filter packets and subsequent resterilization of the reactor. Conventional sparging through the 10 um stainless steel frit produced bubbles on the order of 0.5 to 1 mm in diameter. On day 48 of the fermentation, conventional sparging was replaced by microbubble sparging. The microbubble sparging was run for 8 days prior to the termination of the experiment due to filter fouling. This problem prevented obtaining a maximum productivity from the reactor with microbubble sparging. Fogler (1992) suggests that a minimum of three residence 99 times are required for a steady-state to occur. The liquid-phase residence time in the reactor and microbubble generator combined was 10.4 days. In order for a liquid-phase steady-state to be achieved, it would require 31.3 days. The spinning disk microbubble generator and filter units were not capable of completing this task. The gas-phase on the other hand, had an approximate residence time of 30 minutes. Therefore, to obtain a gas- phase steady-state, a minimum of only 1.5 hours would be required. The carbon and electron balances were done both on the fermentation with conventional and microbubble sparging. For both conventional and microbubble sparging, between 60% and 80% of the incoming CO was converted. The carbon balance closed to 96.3% and the electron balance closed to 97.1% with conventional sparging. With microbubble sparging the closure was lower; the carbon balance closed to 88.3% and the electron balance to 91.1%. The lower closer may be due to some product loss in the microbubble generator. When started initially, and when filters were adjusted, there was some overflow of the microbubble generator. This overflow caused loss of fermentation liquid from the microbubble generator at various times, thereby losing some of the products from the fermentation. At times during the fermentations (approximately once per day), the filters were backflushed to maintain maximum filtration capacity. During those operations the liquid was allowed to accumulate in the reactor. After completing a backflush, the pumps were restarted, and the excess liquid from the reactor was filtered back into the generator. On occasion, the rapid increase in the liquid concentration would cause the microbubble foam to expand, and some liquid would overflow the generator. 100 7.3.2 Mass Transfer Coefficients Mass-transfer coefficients were obtained by varying the inlet gas flow rates and plotting the results as shown in Figure 31. 4.5 4 " 0 Conventional Sparging 3.5 __ a Microbubble Sparging 3 _- ’6 0° 2.5 -- E Kla = 90.64 h'1 8 2 .. 2 a R = 0.9944 V 1.5 -- 1 _ M O 5 KLa = 14.2 h-1 ' R2 = 0.9433 0 : . : i 0 0.05 0.1 0.15 0.2 0.25 1/G (min/mL) Figure 31: Mass transfer coefficients from fermentation By inspection of Equation 28, the data should be linear if the fermentation is mass- KLa transfer limited, with a slope of VLRgT . Results for both conventional and microbubble sparging are shown in Figure 31. The Km for microbubble sparging was four times higher than for conventional sparging, even though the gas throughput for the microbubbles was only about one half that for the conventional bubbles. A lower gas flow rate was used for microbubble sparging to avoid exceeding the liquid-handling capacity of the filters. The microbubble foam had a significant liquid void fraction, approximately 0.32 (Bredwell, 1995), which increased the liquid flow rate 101 through the filter. The KLa value measured in the fermentation is lower than that measured in the bubble column studies, probably because of the very low microbubble flow rate used in the fermentation. The low flow rate gave lower 80 values than were used in the bubble column studies. A comparable four-fold increase in the volumetric mass-transfer coefficient due to microbubble sparging was reported previously in an aerobic Saccharomyces cerevisiae fermentation by Kaster et al. (Kaster, 1990). Filter fouling prevented long—tenn operation of the reactor with microbubble sparging. Since microbubbles have a significant liquid fraction associated with them, the use of microbubbles increased the liquid flow rate into the reactor, especially at higher gas flow rates. This significantly increased the requirements of the filter used for cell retention. Other experiments using microbubbles to deliver synthesis gas to a sulfate- reducing bacteria culture have used dual hollow fiber membranes in parallel to successfully handle the increased filtration requirements (Selvaraj, 1997b). 7 .4 Conclusions The six-fold increase in the volumetric mass transfer coefficient despite lower gas flow rates, indicates that microbubbles can enhance the gas-to-liquid mass transfer in synthesis gas fermentations. The results demonstrated that microbubble technology can enhance synthesis gas fermentations. 7 .5 Nomenclature 7.5.1 Symbols Gi = Inlet gas flow rate (mL/min) H = Henry’s law constant (atrrr/mol*L) KLa = Volumetric mass transfer coefficient (h'l) 102 N = Number of moles (moles) Pico = Inlet partial pressure of CO (atm) POCQ = Outlet partial pressure of CO (atm) Rg = Gas constant T = Temperature (K) VL = Reactor volume (L) y = Reductance degree 7.5.2 Subscripts and Superscripts in = inlet out = outlet prod = produced convert= converted CO = Carbon monoxide CO2 = Carbon dioxide H2O = Water ace = Acetate eth = Ethanol but = Butyrate cells = Cell mass 103 8. SRB FERMENTATIONS 8.1 Introduction Sulfur dioxide ($02) is one of the major air pollutants in the US, with 70% of total SO2 emissions coming from the combustion of fossil fuels, primarily from coal-fired power plants. The quantity emitted in the US in 1992 was estimated to be 22.73 million tons (1992). In the atmosphere, SO2 reacts photochemically or catalytically with other constituents to form sulfuric acid which contributes to acid rain. Conventional flue gas desulfurization (F GD) processes are based on wet-scrubbing techniques that use disposable sorbents such as dolomite or limestone or regenerable sorbents such as copper oxide. Microbial processes, such as sulfate reducing bacteria (SRB), have found potential application in treatment processes of sulfur-laden wastes. Under anaerobic conditions, sulfur compounds (sulfite, sulfate, thiosulfate) can act as a terminal electron acceptor for SRB. Organic acids and alcohols (Colleran, 1995), sewage digest (Selvaraj, 1995), and hydrogen and synthesis gas (du Preez, 1994), (van Houten, 1994) have all been used as electron donors for SRB. While organic acids and alcohols (such as lactic acid and ethanol) may be utilized, they are most likely too costly to use as feedstocks for this process. Synthesis gas, therefore, is an attractive alternative feedstock for SRB fermentations. Bioreactor design for synthesis gas fermentations centers around the issue of providing high gas-to-liquid mass transport rates while trying to maintain economic feasibility. Techniques typically used to increase mass transfer rates, such as increasing the impeller rate to promote bubble breakup, are expensive, especially in large-scale fermenters because power consumption is proportional to the impeller rate to the third 104 power and impeller diameter to the fifth power (McCabe, 1985). Power consumption can be a critical cost in synthesis gas fermentations, because the bioreactors used are expected to be very large. Based on the reactor productivities using sewage digest as a feedstock, in order to biodesulfiuize the effluent from a 1000 MW coal-fired power plant, a total reactor volume of 410,000 gallons would be required (Selvaraj, 1996b). Increasing volumetric productivity can reduce the reactor size needed in biodesulfurization. This chapter describes the development of a microbubble—sparged CSTR for the biodesulfiuization of flue gases by SRB using a synthesis gas feedstock. Gas-to-liquid mass transport limitations of synthesis gas sparging SRB fermentations are also discussed. The work with SRB fermentations was done at Oak Ridge National Laboratory (ORNL) in the laboratory of Eric Kaufinan. Work was done in collaboration with Punjai Selvaraj, Mark Little, and Miguel Rodriguez. 8.2 Materials and Methods 8.2.1 Microbubble Generation A new microbubble generator was built at ORNL. It was similar in design and construction to the previous generator. The average bubble size was measured ex situ with a Coulter LS 130 particle size analyzer (Coulter, Hialeah, FL) using both a microvolume module (MVM) and a flow-through hazardous fluids module (HFM). The MVM is a vessel into which the microbubbles were loaded into water with surfactant and agitated with a magnetic stirrer. The HFM module continuously pumps liquid in a recirculation loop to maintain suspension of the sample. The microbubbles were added to the 15 mL MVM until an obscuration of 8 to 15% was obtained, and the particle size was 105 measured via dynanric light scattering. A similar procedure was done using the HFM. Between each run, the MVM was drained and cleaned, and the fluid in the HFM was filtered. Particle size distributions were obtained for 60 seconds. 8.2.2 Media and Cultures Mixed SRB cultures were isolated fiom sewage solids obtained from the dissolved air floatation (DAF) unit of a municipal sewage treatment plant in Oak Ridge, TN as described previously (Selvaraj, 1997a). The cultures were grown in lactic acid (LA) media for seed cultures and 10 L batch reactor runs. Chloroform was added to the LA media at 25 ppm to inhibit the growth of methanogens. The composition of this medium is given in Table 11. A minimal salts (MS) medium was used for mass transport limitation studies and as the primary reactor media. Its components are listed in Table 12. 106 Table 11: Lactic acid media Component Concentration Citric acid 3.6 g/L CaCl2 0.8 g/L NH4C1 1.0 g/L K2HPO4 0.5 g/L FeCl2 0.52 g/L Yeast Extract 1.0 g/L Cysteine HCl 0.05 g/L Sodium Lactate (60% Syrup) 5.8 mL/L Butyric acid 0.52 mL/L Table 12: Minimal Salts Media Component Concentration Na2HPO4 1.2 g/L KH2PO4 1.8 g/L MgClz 0.7 g/L NH4C1 0.2 g/L Fer 0.04 g/L Mineral Water 50 mL/L Batch Vitamin Solution 0.2 mL/L Heavy Metal Solution 15 mL/L 107 The batch Vitamin solution and the heavy metal solution are given in Tables 13 and 14 respectively. Table 13: Batch vitamin solution Component Concentration Biotin 2.0 mg/L Folic acid 2.0 mg/L Pyridoxine HCl 10.0 mg/L Thiamine HCl 5.0 mg/L Riboflavin 5.0 mg/L Nicotinic acid 5.0 mg/L DL-calcium pantothenate 5.0 mg/L Cyanocobalamin 0.01 mg/L p-amino benzoic acid 5.0 mg/L Thioctic acid 5.0 mg/L Table 14: Heavy metal solution Component Concentration EDTA 1.5 g/L ZnSO407H2O 0.1 g/L Trace Elements 6 mL/L 108 Table 15: Trace elements 11 Component Concentration AlCl3 0.0507 g/L KI 0.139 g/L KBr 0.139 g/L LiCl 0.139 g/L H3BO3 3.060 g/L ZnCl2 0.280 g/L CuCl202H2O 0.326 g/L NiCl206H2O 0.513 g/L CoCl206H2O 0.513 g/L SnCl202H2O 0.139 g/L BaCl202H2O 0.163 g/L Na2MoO402H2O 0.139 g/L CuSeO405H2O 0.139 g/L NaVO3 0.024 g/L In serum bottles, the sulfate source was provided by the addition of up to 4.0 g/L of Na2SO4 or MgSO4. In the CSTR, the sulfite source was provided by a 5% 802, 5% CO2, and balance N2 gas. For growth on synthesis gas, a mixture of 47% CO, 36% H2, 10% C02, 5% CH4, and the balance N2, (Air Liquide, Houston, TX) was used. This gas composition is typical of a coal-fired, oxygen blown Texaco gasifier (Simbeck, 1983). 109 For bottle studies, 100 mL of MS media in a 150 mL serum bottle with butyl rubber stopper was made anaerobic by gassing with nitrogen followed by steam sterilization. The synthesis gas was bubbled through the media after sterilization. The headspace was monitored for synthesis gas components and hydrogen sulfide and was replenished with fresh synthesis gas when needed. The bottles were usually shaken at 100 rpm at 30°C. Tween 20 surfactant was used in the reactor studies at 4 times the surfactant critical micelle concentration (240 mg/L). 8.2.3 Mass Transfer Limitation Studies A 10 L glass carboy was used for a batch reactor for biomass growth. The carboy had a stainless steel headplate that allowed for pH control, gas sparging, venting, and liquid inlet/outlet ports. A pump cycled liquid through a loop with a pH probe, and the pH in the reactor was controlled at 7.0 with Chemcadet controller (Cole Parmer Instrument Company, Niles, IL) with 6 N NaOH and 6 N H3PO4. Nitrogen gas was bubbled through the system to strip off the biologically generated H2S. LA medium was used with the addition of 25 ppm chloroform and 4 g/L of sodium sulfate as a sulfur source. The lactate concentrations were monitored by a model 2700 Dual Channel Biochemistry Analyzer (Yellow Springs Instrument Co., Yellow Springs, OH). Sulfate concentrations were also monitored. The reactor was stirred using a magnetic stirrer and stirbar. When the lactate concentration dropped below 0.5 g/L, the reactor was stopped and the biomass harvested with a IEC Chemical Centrifuge (International Equipment Co., Needham, MA) at up 5200 rpm with N2 gas to fill the basket. The centrifuge was opened in an anaerobic glove box under a nitrogen atmosphere, and the biomass was collected 110 from the centrifuge basket. The biomass was then resuspended in MS media and used for mass transfer limitation experiments. 8.2.4 Fermentations 8.2.4.1 Conventional Sparging A 2 L Virtis Omni-culture chemostat (Virtis Co., Gardiner, NY) with temperature and agitation control was used as the primary reactor vessel. The vessel headplate was modified for acid/base additions and gas and liquid inlets and outlets. The pH was controlled at 6.6 with a Chemcadet controller (Cole Parmer Instrument Company, Niles, IL) with 6 N NaOH and 6 N H3PO4. The reactor was agitated at 300 rpm. The reactor was operated in a continuous mode with a feed rate of fresh MS medium at 0.2 mL/min. A filtration system consisted of two Amicon Diaflo hollow fiber cartridges (Amicon, Inc., Beverly, MA) operated in parallel. All pumps used were Masterflex peristaltic pumps (Cole Parmer Instrument Company, Niles, IL). Nitrogen was sparged at 50 mL/min to the reactor vessel to strip the produced H2S from the reactor. Synthesis gas was added directly to the reactor through a stainless steel fiit. A flow diagram is given in Figure 32. 111 (D ./e——>Vent KT Synthesis MS Medra ——.—6 1 Gas soz/N2 H3PO4 | I I'— NaOH I. i v v i l — ' /, :1—-—>< D Figure 32: SO2 reducing fermentation flow diagram 8.2.4.2 Microbubble Sparging The reactor scheme used for microbubble sparging is similar to the apparatus used for conventional sparging. The synthesis gas was supplied directly to the microbubble generator. The synthesis gas microbubbles were delivered to the reactor through a peristaltic pump. The liquid in the reactor was filtered, with the solids being returned to the reactor and the liquid being returned to the microbubble generator. 8.2.4.3 Mass Transfer Studies The mass transfer characteristics of the fermentation were measured while the system was under continuous sparging. The inlet synthesis gas flow rate was changed, and the resulting conversion of CO and H2 was monitored. A mass balance on the two consumed gas components in the fermentation is given below (Klasson, 1991b). 112 L [552.51% = K a“) __VLRT (31) P50 P,” HG" P‘ P’ V RT i——’ =K,a”2 L _ (32) P132 P,” HG‘ 8.2.5 Analytical Techniques 8.2.5.1 Gas Chromatography Hydrogen sulfide in the effluent gas was analyzed using a gas chromatograph (Hewlett Packard HP 5890 Series II) equipped with a teflon column (30 in x 1/8 in) packed with Super Q (80/ 100 mesh) (Alltech, Waukeegen, WI). Temperatures of the column, injection port, and thermal conductivity detector were 70 0C, 125 °C, and 125 0C respectively. The canier gas was helium at 25 mL/min. The calibration was based on 1%, 5%, and 10% H23 in nitrogen standards. Synthesis gas components were measured using a gas chromatograph (Hewlett Packard HP 5890 Series II) equipped with a HP PLOT molecular sieve 5 A capillary column (30 m x 0.32 mm) with a 12 um film thickness. Temperatures on the column, injection port, and thermal conductivity detector were 55 °C, 100 °C, and 240 °C respectively. Liquid samples were filtered through a 0.22 pm syringe filter and analyzed by gas chromatography using a HP 5890 Series II with a HP WAT (crosslinked PEG) capillary column (30 m x 0.53 mm) with a 1.0 pm film thickness. The column temperature program was initially 70 °C followed by ramping to 200 °C at 25 oC/min with a 1.2 min hold then ramping to 225 °C at 25 °C/min with a 3.0 min hold. The injection port temperature was 245 °C while that of the flame ionization detector was 265 °C. 113 .____.. __..-.. _._._ .. _ .. .. ._. --....___ _.. W, 3.2.5.2 Sulfur Chemistry The sulfite in the reactor was analyzed spectrophotometrically by the reaction of fuchsin and formaldehyde in sulfuric acid (Steigrnan, 1950). The sulfide in the reactor was precipitated using zinc acetate in a basic solution. The resulting zinc sulfide was resuspended and reacted with dimethyl-p-phenylenediamine and ferric chloride to form methylene blue. The concentration of sulfide was then measured spectrophotometrically at 660 nm and compared to a sulfide calibration curve. The concentration of a sulfide solution was determined from potentiometric titration with a standard lead chlorate solution until an endpoint was reached. From this known sulfide concentration an calibration curve was constructed. A sample of the potentiometric titration is given in Figure 33 below. ~780 ' ii) 1 2 3 4 5 6 7 8 -800 -. ‘- I It ._ - - I. - -820 .. -840 .1 i- '860 u- u- -880 .. Potential (mV) -900 .. -920 .1- -940 . 4 -960 .1. mL 1000 ppm Pb(CIO4)2 added Figure 33: Potentiometric titration of sulfide 114 8.2.5.3 Most Probable Number (MPN) and Misc. Analytical In all runs the reactors were monitored for 8032' and suspended solids. Effluent gas was also monitored for H2S and synthesis gas components. Effluent gas flow rates were monitored with a wet test meter (GCA/Precision Scientific, Chicago, IL) for two hour periods. For solids determination, the liquid samples of known volume were filtered through tared Whatrnan glass microfiber filters (Grade GF/C) using a Gelman 25 mm polysulfone filter funnel. The filters were then dried in for 2 h in an oven at 60 OC followed by cooling to room temperature. The SRB in the liquid samples were enumerated by the MPN method using SRB medium from Bioindustrial Technologies, Inc. (Houston, TX) as described previously (Selvaraj, 1997a). 8.2.6 Carbon and Electron Balances Mass balances on sulfur and carbon were done with 99.6% conversion of SO2 to H2S and no liquid-phase products assumed. C0”' + C05" = C0“ + cog“ (33) S0321 : HZSout + SZ—,out + 8032-011! (34) An electron balance was done around the reactor and is shown below. CO H, .90? _ C02 HZO yCOnconvert + 7H2 ”convert + 7503* ”convert _ ySZ- nS2- + yCOanrod + 7H20nprod (35) The reductance degrees, y, for CO, H2, SO32} S2", CO2, and H20 were 2, 2, -8, -2, 0, and 0 respectively as suggested by Erickson (1978). 115 8.3 Results and Discussion 8.3.1 Confirmation of Mass Transfer Limitations Biomass was grown in a 10 L batch reactor on lactate and was harvested and resuspended in the MS. The cultures were then switched to a synthesis gas feedstock and adapted to it over a period of two or three days. At the start of the experiment, the headspace and liquid of each biomass concentration (0.45 g/L, 2.40 g/L and 5.45 g/L) was sparged with fresh synthesis gas for 5 minutes prior to monitoring by GC. The headspace in each bottle was sampled and analyzed every 30 minutes. The pressure inside of the serum bottles was monitored using an inert tracer gas, CH4. Since the bacteria did not consume or produce this gas, the amount in each sealed serum bottle would not change during the fermentation. Therefore, using a ratio of final concentration of CH4 to initial concentration of CH4, the pressure inside of the serum bottles could be adjusted. Carbon monoxide depletion is shown in Figure 34, and hydrogen depletion is shown in Figure 35. 116 [cm/[c0]o 1.2 .._.—_C_. .— 0.8 .. I o A . . _ 0.6 .. . g o 0.45 g/L solids { A . 2 2.4 g/L solids 7 g 5.45 g/L solids ; 0.4 I. . L— " hi A I 02.. ’- A A I 0 ‘ ' ' . . - iH‘LH—l—H— 0.00 1.00 2.00 3.00 4.00 5.00 6.00 7.00 8.00 Time Elapsed (hours) Figure 34: CO uptake by SRB in batch serum bottles 117 9.00 1.2 , 1 A I ‘- ‘I I I O. ‘ I t 0.8 .- r 0.6 .. '1': _ I E, ,9 0.45 g/L solids ; ° 04 .2AgAsmMsi ' 1. 5.45 g/L solids f " t I 0.2 .. o ' A 0 .L : : ' s 4 :: : 0.00 1.00 2.00 3.00 4.00 5.00 6.00 7.00 8.00 9.00 Time Elapsed (hours) Figure 35: H2 uptake by SRB in batch serum bottles The uptake pattern of synthesis gas components by SRB is typical of other synthesis gas consuming bacteria (Vega, 1989b). Typically CO is consumed first to some critical concentration before H2 uptake begins. This pattern may be due in part to inhibition of the hydrogenase enzymes by CO (Ragsdale and Ljungdahl, 1984). The inhibition of the hydrogenase enzymes by CO is through reversible binding of CO to the iron-sulfur active site of the enzyme (Adams, 1987). Also, this trend could be due in part to the production of H2 from the CO metabolism by SRB (Yagi, 1959). The uptake rates of both carbon monoxide and hydrogen did not show significant variation for the different concentrations of biomass. Vega et al. (1989b) and Worden et al. (1989) show similar results for other synthesis gas consuming bacteria where the uptake rate became independent of biomass concentration. As shown in Table 16, an 118 order of magnitude change in concentration of biomass did not cause significant changes in the uptake of the CO component of synthesis gas. Table 16. Synthesis Gas Uptake from Serum Bottles Suspended Solids CO (mmol/h) H2 (mmth) 0.45 g/L 19.0 i 9.9% 28.7 i 11.0% 2.40 g/L 17.8 i 10.4% 24.3 i 0.1% 5.45 g/L 18.8 i 2.7% 14.6 i 3.6% This indicates that the serum bottles were mass transfer limited for CO at all three of the biomass concentrations tested. However, decreased H2 uptake was observed at the higher biomass concentration of 5.45 g/L. This may be due to mechanism in the system where the hydrogenase enzymes are more inhibited by CO and higher biomass concentrations, an artifact, or analytical errors. These data show that the serum bottle was also mass transfer limited for H2 at all three of the concentrations tested. The MPN number was determined from the 0.45 g/L serum bottle and was found to be 1.10 x 10H cells/L. The average mean weight of a dry cell of SRB is 3.125 x 10'13 g/cell (Postgate, 1984). Therefore the dry weight of the SRB in the serum bottle was 3.44 x 10'2 g/L, indicating that the viable SRB concentration in the bottle was only 7.6% of the total solids. 8.3.2 Conventional Sparging A CSTR was operated for extended periods of time using synthesis gas as a feedstock. The SO2 flowrate was gradually increased until the effluent concentration of 8032' began to increase. At this point the flow of SO2 to the reactor was decreased until 100% conversion of SO2 was achieved. Sulfite, S037”, is toxic to SRB. Concentrations of 40 ppm seem to have significant inhibitory or toxic effects on SRB observed when the 119 S03 concentration in the reactor was increased above 40 ppm. Sulfite is formed when SO2 is dissolved in water. An increase in the 8032' concentration indicates that the maximum SO2 conversion rate has been obtained. The reactor was capable of a maximum throughput of SO2 of 1.23 mmol/h*L with 100% conversion of SO2 to H2S. The reactor utilized 1.8 mmol H2 and 2.3 mmol CO per mole of SO2 reduced. The suspended solids concentration was measured to be 3.17 g/L, and the MPN was determined to be 2.3 x lOlZ/L when the reactor was determined to operate at steady- state. These numbers indicate that 22% of the suspended solids are viable SRB. While the mass transfer characteristics of the CSTR are certainly different than the serum bottles, the comparable biomass concentrations in both systems suggests that the CSTR could be mass transfer limited, i.e. the effective concentration of the liquid-phase substrates were approximately zero. A N2 sparge directly into the reactor was necessary to insure adequate stripping of H28 from the reactor vessel to prevent inhibition of culture growth and productivity. Reis et al. (1992) found a concentration of 547 ppm was capable of totally inhibiting grth in SRB and suggests that this toxicity is reversible. The pH was maintained at 6.6 to give good growth characteristics while keeping the produced H2S as hydrogen sulfide that was capable of being stripped. At higher pH, much of the H2S exists as S2" (Grethlein, 1992b) and cannot be stripped from the liquid. The equilibrium reaction under aqueous conditions is given in Equation 36 below (Grethlein, 1992). HZS¢:>H++HS' pKa1=7.04 (36) HS‘c>H++S2' pKa2=12.0 The mass transfer characteristics of the reactor with conventional sparging were determined. Once again, the inert methane tracer was used to adjust for changes in 120 pressure and dilution of the component gases by N2. The volumetric mass transfer coefficients for CO and H2 were calculated from the slopes in Figure 36 and Figure 37, respectively. The data are plotted as suggested by Equations 31 and 32. Under mass VLR T g transfer limited conditions, the slope should be equal to K La 4.5 .. 4 .. 3.5 .. 3: l >- O. . 2-5 -r y = 22.749x + 0.6764 8 R2 = 0.8896 3 2 .. 7) O 0 "9:15 .1. 1 .. 0.5 -- 0 1 : k 1 4' i 1 : 0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16 1N (min/mL) Figure 36: Mass transfer coefficient for CO 121 y = 47.211x - 0.2524 (PIHZIPOHZ'PiIIPOI) A 3 .. R2 = 0.9946 2 .. 1 .. 0 : : : i I. i : : 0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16 1N (min/mL) Figure 37: Mass transfer coefficient for H2 The subsequent calculated KLa values calculated from the slopes of Figures 36 and 37 with conventional sparging were 30 h’1 and 75 h'1 for CO and H2, respectively. These values are similar to the KLa value for CO measured in a mass transfer limited 1 L CSTR for a mixed triculture synthesis gas fermentation (28.1 h'l at 300 RPM (Klasson, 1992)). The KLa values for CO and H2 are related through their diffusion coefficients (Cussler, 1984) as shown by the defining equation for the mass transfer coefficient based on film theory. k,= D I (3 7) The ratio between diffusion coefficients calculated by the Wilke-Chang equation is 1.93 for H2 to CO. The ratio of the experimental mass transfer coefficients measured is 2.5. 122 8.3.3 Microbubble Sparging In the control run, a separate liquid reservoir bottle with surfactant was used to simulate the microbubble generator; the liquid was pumped from the reservoir bottle to the reactor at a flowrate that would be comparable to the rate liquid would be carried over in the microbubbles, i.e. from 3 mL/min to 7 mL/min. The dual filter system was capable of handling the increased liquid processing requirements. Fresh feed was added to the reactor at 0.2 mL/min and removed from the reservoir bottle at 0.2 mL/min. The system was brought to a maximum productivity of 1.25 mmol/h*L with conventional sparging and maintained for two days. The microbubble generator replaced the liquid recycle bottle, and the synthesis gas was added to the reactor as microbubbles. The synthesis gas flow rate to the reactor remained at a constant 10 mL/min (of gas phase) for both conventional and microbubble sparging. The flow rate of 802 was slowly increased over the period of 24 hours until a maximum productivity was obtained with microbubble sparging. The SO2 throughput in both phases of this experiment and the subsequent 8032' concentration are shown in Figure 38. 123 400 .. .. 2.5 350 .. .2---~______ _. _ _ _ _. .. 2 A 300 .. -a-Sulfite (mg/L) 1 g. —o— 802 Throughput (mmol/h.L) l a d 250 "' .. 1.5 a g: :: v 5 3 200 .. a 2.: .. .. 1 = 0:) 150 .. e .4: [-1 100 .. N . 0.5 8 50 .. Figure 38: Reactor productivity The maximum reactor productivity obtained using microbubble sparging is similar to the value of 1.27 mmol/h*L obtained by Selvaraj et al. (1997a) using soluable substrate; sewage digest medium. This result indicates that at the maximum reactor productivity using microbubble-sparging the reactor might no longer be in the mass transfer limited regime. These results also provide more evidence that the CSTR under conventional sparging was mass transfer limited. By keeping the flowrate of the gaseous substrate constant and only decreasing the average bubble size entering the reactor, a 2-fold improvement in the maximum SO2 throughput was obtained. To obtain the mass transfer coefficients using microbubble sparging, the same approach as conventional sparging was used. The derivations of Equations 31 and 32 did not account for the small amount of dissolved CO or H2 that is carried into the bioreactor 124 by the interstitial liquid of the microbubble foam. However, this amount of dissolved gases was calculated to be negligible compared to the amount of gas that enters with the microbubbles and is transferred to the medium. Due to the difficulty of maintaining long- term (more than 2 or 3 days), high-productivity, microbubble-sparged fermentations, only two data points have been obtained for the evaluation of KLa values. Using those two data points, an estimate of the KLa value for CO was 104 h’1 and for H2 was 190 h]. These numbers represent a 3.4-fold increase for CO and a 2.6-fold increase for H2 with a H2/CO ratio of 1.85. These numbers compare well with the mass transfer coefficient determined in Chapter 6 of 91 h'1 for CO. These numbers are comparable regardless of impeller speed, because the volumetric mass transfer coefficients from microbubbles have been shown to be independent of impeller rate (Chapter 6, Kaster 1990). The difficulty in maintaining long-term, high-productivity SRB fermentations is the potential instability due to the rapid build-up of a toxic compound, 8032', if the fermentation fails. Once the SO32" concentration begins to rise, the reaction rate slows down, causing the S032" concentration to rise faster. At the maximum productivity using microbubble sparging, toxic levels (>40 ppm) can build up within 20 minutes if a failure in the bioconversion occurs. 8.3.4 Carbon and Electron Balances The sulfur, carbon and electron balances were calculated for the data giving the maximum productivity for both conventional and microbubble sparging. For conventional sparging, the inlet sulfilr was recovered to 98% in the outlet sulfur compounds, with 99.6% of the outlet sulfur being H28 and 0.4% being 82’. There was no measurable SO32" present in the reactor. The available carbon was 95% recovered in the 125 outlet carbon dioxide. The remainder of the carbon may have been incorporated into various liquid phase products such as acetate, propionate, and cell mass. Of the available electrons, 88% were recovered in the H2S and 82'. For microbubble sparging, the inlet sulfur was recovered to 91.6% with 97.2% of the available sulfur recovered in H2S. The available carbon was recovered to 88.4% in outlet carbon dioxide. Similarly, 86.2% of the electrons were recovered in the outlet sulfur compounds. 8.4 Conclusions Experiments in both serum bottles and stirred-tank reactors provided evidence that synthesis-gas-fed sulfate reducing bacteria fermentations are typically mass transport limited. In the serum bottles experiments, increasing biomass concentrations did not increase the rate of synthesis gas uptake. In the stirred tank reactor, the maximum productivity increased significantly upon switching from conventional bubbles to microbubbles. The KLa values for the conventionally-sparged SRB fermentations in the stirred tanks were 31 h'1 for CO and 75 h'1 for H2. These values are comparable to literature values for synthesis gas fermentations in lab-scale stirred tanks. The maximum productivity was 1.25 mmol/h*L. Upon switching from conventional sparging to microbubble sparging, with no change in the flow rate of the synthesis gas, a 2-fold increase in reactor productivity was obtained within 24 hours. By increasing only the interfacial area available for mass transfer, a maximum reactor productivity of 2.5 mmol/h*L was obtained. Finally, preliminary mass transport coefficients obtained during microbubble sparging indicate a 126 3.4 fold increase in Km for CO and a 2.6 fold increase for H2 over the conventionally sparged-stirred tank. 8.5 Nomenclature 8.5.1 Symbols a = interfacial area per unit gas volume (ml) CO = flow of moles CO (mol/min) CO2 = flow of moles CO2 (mol/min) G = Total gas flow rate (mL/min) H2S = flow of moles H2S (mol/min) H = Henry’s Law coefficient (atm/mol‘L) H = stagnant layer film thickness (m) KL = mass transfer coefficient (m/s) n = number of moles (mol) R = gas constant S2' = flow of moles SZ‘ (mol/min) SO2 = flow of moles SO2 (mol/min) 8032' = flow of moles SO32' (mol/min) T = Temperature (K) t = Time (h) VL = Volume (L) 8.5.2 Greek Symbols y = reductance degree 8.5.3 Subscripts and Superscripts 127 in,i out,o C0 C02 SO2 H28 H20 so32' convert prod inlet outlet Carbon monoxide Carbon dioxide Sulfur dioxide Hydrogen Hydrogen sulfide Water Inert component Sulfide Sulfite converted produced 128 9. COALESCENCE 9.1 Introduction Coalescence in stirred vessels has been extensively studied for liquid—liquid dispersions by Coulaloglou and Tavlarides (1977), Tsouris and Tavlarides (1994), and Sovova and Prochazka (1981) among others. The above authors have successfully used the population balance equation (PBB) model to describe the relative rates of breakage and coalescence in dispersive systems. The pbe model has also been used to describe flocculation in colloidal systems (Tsouris, 1995b; Tsouris, 1995a) and for bubbles (Prince, 1990; Oolman, 1986b). This model is very useful in describing the history of a pOpulation in terms of properties such as size, age, and concentration, and interaction events with themselves and their environment. Prince and Blanch (1990) studied bubble coalescence in a bubble column under laminar, buoyant, and turbulent shear. They considered that the coalescence of two bubbles in turbulent flow occurs through three steps. The first step is the collision of two bubbles, which traps a thin film of liquid between them. The rate of collisions is a key kinetic parameter during this step. The second step is drainage of the liquid film between the two bubbles until a critical thickness is reached, and the third step is the rupture of the thin film, at which point the two bubbles become one. The coalescence efficiency is a second paramater that controls whether two bubbles will coalesce and is defined as the fractional probability that two colliding bubbles will undergo a coalescence event. For two bubbles to coalesce, the time the bubbles remain in contact must be greater than the time required for the drainage of the liquid film between the bubbles. The same mechanism for drop coalescence was proposed by Coulaloglou and Tavlarides (1977). 129 The effects of surfactants in both liquid-liquid and gas-liquid dispersions have also been studied. Koshy et al. (1988) studied the breakage of droplets in a water-styrene- Teepol (surfactant) system. They reported that the breakage of drops in stirred dispersions is affected by the surfactant not only through the reduction of the surface tension, but also through the generation of an interfacial surface tension gradient along the drop surface through drop deformation (Koshy, 1988). The surface tension gradient results in an extra stress (Maragoni stress) that adds to the turbulent stress, increases the breakage rate, and thereby reduces the average droplet size formed. This idea is also supported by work in gas-liquid dispersions by Zahradnik et al. (1995), which found that there is an overall smaller bubble size in an electrolyte-laden system over a ‘clean’ system. This observation indicates decreased bubble coalescence in the presence of the electrolyte due to surface immobilization. These observations support the earlier work of Prince and Blanch (1990), who also studied the effects of electrolytes in bubble columns. One of the assumptions made in the development of the Prince and Blanch model was that bubble coalescence does not occur in systems where the salt concentration is sufficient to immobilize the gas-liquid interface. Therefore, in their model, the coalescence efficiency of bubbles with immobile interfaces is taken to be zero. This assumption is based on film thinning times. The film thinning time for bubbles with immobile interfaces has been predicted by Prince and Blanch to be on the order of seconds, and even in the case that surfactant transport is incorporated to give a partially mobile gas-liquid interface, the coalescence times are still on the order of hundreds of milliseconds (Prince and Blanch, 1990). Comparatively, the contact time in turbulent 130 dispersions was calculated to be 10 ms, which is much smaller than the coalescence time that is needed for coalescence to occur. Because the proposed structure of microbubbles generated in this work includes a stagnant water shell, there should be minimal coalescence, according to the assumptions from Prince and Blanch (1990). However, microbubbles do coalesce, as shown by the dynamic measurements of bubble size performed in this study. Therefore, the assumption that coalescence is negligible in systems with immobile surfaces was rejected. However, the coalescence rate of microbubbles is low in comparison to conventional bubbles. It is the purpose of this chapter to describe the dynamic coalescence phenomena of microbubbles in a stirred tank. The effects of increasing the amount of surfactant present in the surrounding liquid was investigated through both experiments and modeling. 9.2 Mathematical Model 9.2.1 Population Balance Equation Model The population balance equation model is presented below with terms for both coalescence and breakage. 6n(v, t) at where n(v,t) is the number of bubbles of volume v at a given time t, and D, B represent +D—B=O o& the death and birth rates of bubbles from coalescence and breakage. These terms are given in the following equations: D = n(v,t) f1(1),v')h(v,v')n(v')dv'+g(v)n(v,t) (39) and B = VJMV — v' )h(v — v' )n(v — v’ )n(v' , t)dv'+oj[fl(v, v' )U(v' )g(v' )n(v' , t)dv' (40) 0 131 where u(v’) is the number of bubbles formed from the breakage of a drop of volume v’; ,B(v,v’) is the probability density of the daughter bubbles, which represents the probability of forming bubbles of size v from breakage of drops of size v’; g(v’) is the breakage frequency of bubbles of volume v’; h(v,v) is the collision frequency of bubbles of volumes v and v’; l(v,v’) is the coalescence efficiency once collision occurs between bubbles of volumes v and v’. Volume discretization of equation 38 gives a system of integro-differential equations that represent an initial value problem that can be solved numerically (Coulaloglou and Talvarides, 1977; Tsouris and Talvarides, 1994). This model describes the history of the bubble population in terms of bubble properties, such as size, age, and concentration, during the course of bubble interaction events with themselves and the surrounding environment. Since the microbubbles formed in this work are very small in comparison with typical bubbles, the rate of breakage may be negligibly small and hence neglected in the pbe model. The small size of the bubbles would require extremely high levels of shear in order for further breakage to occur. In order to evaluate whether breakage may be significant, the maximum stable bubble size, dcr, can be estimated using the following correlation (Shinnar, 1961), EN2 dc, = Cm We‘o'éD, We = LL— (39) 0' where We is the Weber number, defined as the ratio of surface tension forces to dynamic pressure forces, and Cm is a constant in literature (Sprow, 1967) as 0.125. For dilute suspensions, the critical diameter below which bubbles do not break, def, is the same as the maximum stable drop size, domax (Shinnar, 1961). To account for the holdup effect on 132 the maximum stable drop size for a non-dilute suspension, considering Kolmogorov’s 3 microscale (with the ratio K— ), Doulah (1975) showed the following, 5 8: = (”—1 (42) 8 V . . . . . . '3 . where 8C 15 the energy drssrpatron rate per unit mass for the pure contrnuous phase, 8 rs the energy dissipation rate per unit mass of the dispersion, and v. and vc are the kinematic viscosities for the dispersion and continuous phase, respectively. Taylor (1932) derived the following relationship for the apparent viscosity of a liquid dispersion. g‘ = )1, 1+ 2.54%] (43) lud + Iuc Using Equations 42 and 43, and the assumption that for dilute suspensions, pc = pi, the following expression can be used to describe the effect of holdup on the maximum stable bubble size (Tsouris and Tavlarides, 1994). 1.2 dc, = cm We‘0-6D, [1 + 25¢[WJ] (44) #d + #c For the experimental system used in this work, dcr was calculated to be 356 um, above which bubble breakup becomes significant enough that it can no longer be neglected. From the dynamic experimental data obtained over six minutes, the mean bubble size was found to range between 66 um and 130 pm with 98% of all bubbles measured to be below the calculated dcr. Thus breakage was neglected in the population balance equations over the time frame examined. The simplified birth and death terms are shown below. 133 D = n(v,t)?fl.(v,v‘)h(v, v')n(v')dv’ (45) v/2 B: JA(v—v',v')h(v—v',v')n(v—v',t)n(v',t)dv' (46) 0 9.2.2 Collision Frequency Energy dissipation in the system is how much mechanical energy is being supplied to the fluid in the vessel. In the experimental system used in this work, the energy dissipated is supplied by the impellers. The amount of energy dissipated depends on the impeller size, speed, and tank geometry as shown below. N31).5 T4; (47> 8(x,y,z) = kl (x, y, z) where T and H represent the vessel diameter and height respectively. Assuming that the constant, k1 is not a function of position in the vessel, the mean energy dissipation can be determined. Bertrand (1980) used the following relation if T=H=2Di. .1} = 0.81N3Df (48) where 1:: is the average energy dissipation in the system. The influence of inhomogeneity of the turbulence in stirred vessels on the coalescence and breakage has been studied (Coulaloglou and Talvarides, 1977; Tsouris and Talvarides, 1994). The inhomogeneity of the turbulence is an important factor in the determination of the breakage fiequency, which however, was neglected in the experimental system used here. The model in this study uses the mean energy dissipation as suggested by Bertrand (1980), but with the result doubled to account for the presence of dual impellers in the experimental system used in this study. Since the system has a low gas hold-up, the gassed power draw on the second impeller is the same as the power draw on the first impeller (Tatterson, 1991). 134 The assumption that the bubbles are in the inertial subrange is necessary to neglected short-range forces such as Van der Waals and that energy is transferred from eddies to bubbles. In order to verify this assumption, the wave number of the bubbles (the inverse of the bubble diameter) is compared with the limits of the eddy wave number. The lower limit on the size of the eddies in the vessel is given by the Kolmogorov microscale and it is given by: 3 H4 r" = 2(3) (49) For a mean energy dissipation of 24,000 cmz/S" which corresponds to 500 rpm agitation speed in our system, rn=0.0025 cm and kn=398 cm'l (upper limit of the eddy wave number). The largest possible eddy size in the vessel is on the order of the impeller radius. For a 5 cm diameter impeller, ke=1/re=0.4 cm]. The number-averaged bubble size in the vessel ranged from 60 pm to 165 pm during the study, which gives a kd from 60 cm’1 to 167 cm". Therefore, ke < 60 50' r I I I I I I 0 1 2 3 4 5 6 7 Time (min) Figure 46: Dynamic average bubble size Since the coalescence efficiency is dependent on the bubble size, smaller bubbles are predicted to have higher coalescence efficiencies and were observed to have such behavior with the efficiency model by Coulaloglou and Tavlarides (1977). However, the trend is not as pronounced in this model as the previous model; in the Coulaloglou and Tavlarides (Equation #55) model the coalescence time is proportional to the bubble volume to the fourth power, while in the Prince and Blanch (Equation #65) model the coalescence time is proportional to the bubble volume to the one-half power. The coalescence efficiency calculated by the Prince and Blanch model is given (as a function of both bubble sizes in binary coalescence) in Figure 47. 152 0.18% . 0.16% 1 ° 0140/ oD2=60.0 um . 0 -l g 02 = 75.6 um 0.12% . A D2 = 86.5 um > 0 D2 = 95.2 um 0 0.10% . x02: 102.6 um C .3 0 - 02 = 109.0 um 15 0.08% . Lu 0.06% . ° 0 0.04% . D 0 0 Cl 000 0.02% . A U 6 DUE] I 11",” ............. 000% I '96 . IIIEEMIWISUI'I-I11111111.......... . .. , ., 0 50 100 150 200 250 300 Bubble Diameter (um) Figure 47: Coalescence efficiency The efficiency had a maximum of 0.17%, indicating that out of every 1000 collisions, there will be 17 occasions of coalescence. The efficiency is a function of the starting minimum bubble size and drops off significantly for larger bubbles. The collision frequency increases with bubble size (larger bubbles are more likely to have a collision with another bubble) and with the number of bubbles present in the system. A higher rate of collision is most likely to occur with microbubbles since for a constant volume of gas, there will be more bubbles in the system due to their small size. Another factor that contributes to reduced coalescence rates is the reduction in the number of bubbles in the system available for coalescence. Of the 10 mL of microbubble dispersion initially injected, 68%, or approximately 6.8 mL was gas (Bredwell, 1995). Assuming an average diameter of between 65 um and 75 um (depending on the data set), 153 this volume corresponds to approximately 28,000 bubbles present. At the end of the experiment, there were significantly fewer bubbles in the system, due to bubble coalescence. Basing the starting conditions on the experiment (where 28,000 bubbles were determined at time zero based on the average initial bubble size and the amount of gas injected into the system) the number of bubbles as a function of time that are predicted by the model are given in Figure 48. 30000 1 8 - _ 25000 . o g A A A 0.0 mg/L Tween 20 in Bulk $ 0 A A A g 180 mg/L Tween 20 in Bulk 8 o D A A A A 300 mg/L Tween 20 in Bulk > 1:1 A A 2 .E O D U A A A A U) 0 D D D A A A A A A e o D D AAAAAAA 3 o o '3 D A A A A .6 15000. 0 DDDDDD : 00 000000 m 0 O o O D D D U C] C] H- O O O o o O 0 O o o o 0 0 O a; 10000. 000066 .o E :3 z 5000 . O I I I I I I I 0 1 2 3 4 5 6 7 Time (min) Figure 48: Number of Bubbles in Vessel 9.4.3.2 Physical Significance of Fitted Parameters The fitted parameters have physical significance in terms of film thickness and other surfactant properties, such as bulk hindrance due to steric effects. Also, the surface tension, which was measured in a static system, may not accurately describe the dynamic surface tension that occurs at the interface between the two bubbles. As mentioned 154 previously, a surface tension gradient (Koshy et al., 1988) can form when a bubble is deformed in the presence of surfactant. The main difference in the three cases studied was the surfactant concentration, which manifests itself both in the surface tension and in terms of steric hindrance and/or fihn thickening. In order to investigate the effects of surface tension in the value of k5 obtained for all three surfactant concentrations, the following group was calculated. k5 C1 = 401/2 (68) The form of C1 was determined by the form of the model for coalescence time (Equation #63). This normalization combines the effects of surfactant properties into one fittable parameter. Since it is difficult to obtain a true measure of the surface tension in a dynamic system, the dynamic effects are lumped into C]. The effect of surfactant concentration on C1 is shown in Figure 49. 155 36. 341 32. 30 . y =1.12x + 28.09 2 ._ 28 4 R — 0.9979 26.. C1 (cm/dyn)“2 24. 22. 20 0 1 2 3 4 5 6 Dimensionless Surfactant Concentration Figure 49: Effect of surfactant on C1 The increase in C1 with DSC can be related to the initial and final film thickness between h two bubbles. Prince and Blanch (1990) proposed that k5 is related to In —°— , where ho and f hf are the initial and final film thickness between two bubbles. It is not clear how to incorporate into the model the effects of the physical size and presence of surfactant in the liquid film, either absorbed at the gas-liquid interface, or present as micelles in the liquid layer. Therefore, the absolute value of k5 is not significant, but the general trends are significant. The increasing trend seen in Figure 48 suggests that the surfactant causes either increased film thickness in the liquid fihn surrounding a bubble, or the higher concentration of surfactant becomes a greater physical barrier to coalescence, or both. 156 The increasing trend continues beyond the critical micelle concentration, which would lend evidence to the presence of a greater concentration of micelles and surfactant in the surrounding liquid. This might cause an increased resistance to coalescence through a larger amount of bulk hindrance and electrostatic effects in the system. The predicted coalescence time calculated using the optimal k5 values for a pair of 75 um bubbles is 18.2 ms in the presence of no surfactant and 21.8 ms at 300 mg/L of surfactant with coalescence efficiencies on the order of 0.04%. In contrast, the coalescence time for a conventional bubble calculated using film thicknesses suggested by Prince and Blanch (1990) was 0.2 ms, giving a coalescence efficiency of 87.5%. The absolute value of C1 is not meaningful in this study because this parameter is thought to be dependent on other, poorly characterized influences. For instance, the model developed by Li (1996) could not explain data for systems containing high concentrations of surfactant. He suggested that the discrepancies could be due to electrostatic forces. In addition, steric hindrance and surface blocking could be involved. While these phenomena may not be important in turbulent systems due to high inertial energy, they could have a significant role in processes involving microbubbles. The C] parameter does not distinguish the various influences, but rather includes them all into one term. These data from the coalescence studies support the trends seen in Figure 22 in which increasing the surfactant concentration in the bulk liquid reduced the mass transfer coefficient, KL. This trend is most likely due to thickening of the liquid film or shell around the microbubbles and physical (steric) hindrance by the surfactant molecule in the shell. The mass transfer resistances around a microbubble can be considered using a 157 resistances in series model as proposed by Worden and Bredwell (1998) in the following Equation 69. —=—+— (69) where l/kL,S is the resistance due to the total shell of a microbubble (surfactant layers and liquid layer), and RL is the local resistance of the bulk liquid. Using a dynamic model of a single microbubble, Worden and Bredwell (1998) found that for a kl”, value of 0.0001 m/s or less, the shell resistance is expected to dominate. The mass transfer coefficients calculated in Chapter 6 and shown on Figure 22 are all less than 0.00013 m/s in the presence of surfactant. This indicates that shell resistance is most likely to dominate resistance in this system. Princen and Mason (1967) suggested that gas transfer through soluble surfactant monolayers occurs through simple Fickian diffusion, and Quintana et al. (1990) has reported that the presence of soluble surfactants does not significantly affect the diffusion of small molecules. Also, gas-side film resistances have been shown to be negligible for sparing soluble gases (Motarjemi, 1978). These results suggest that the primary resistance to mass transfer in the microbubble shell would be in the stagnant water layer around the microbubble. The thicker this shell of water, the lower the mass transfer coefficient. Therefore, the evidence from the coalescence data suggests that the liquid shell around a microbubble has been effectively thickened by increased concentrations of surfactant. However, other phenomena, such as surface blocking and electrostatic forces may also contribute significantly. 158 9.5 Conclusions The coalescence rate of microbubbles in batch stirred vessels has been measured through video microscopy and found to decrease with increasing bulk surfactant concentration. Bubble size versus time data were analyzed using a population balance approach with two different mathematical models: one giving the coalescence efficiency for deformable drops and one based on the dynamic drainage of the liquid film between two colliding bubbles. The deformable drop model grossly underpredicted the coalescence efficiency for larger bubbles. The liquid drainage model reasonably modeled the data, although it still underpredicted the coalescence of larger bubbles. Therefore, better models should be developed. The film drainage model suggested that the liquid film between two colliding bubbles is thicker in the presence of higher concentrations of surfactant. This result is consistent with the mass transfer results, where KL decreased with increasing surfactant concentration in the bulk liquid. The resistance to coalescence was increased through possible mechanisms such as steric hindrance and physical blocking in the liquid film. Modeling work done by other researchers indicates the majority of the mass transfer resistance from microbubbles occurs in the surrounding liquid shell, and the thickening of that shell would therefore decrease the mass transfer coefficient. 9.6 Nomenclature 9.6.1 Symbols A = Hamaker constant B = Birth term of bubbles Cm = Constant for maximum stable bubble size 159 D Di d°m.x.dcr d g(V) H h(d,d’) h(V,v’) hf kL,s 0°70 rearn Constant (cm/dyn)”2 Concentration of surfactant molecules (mol/L) Death term of bubbles Impeller diameter (m) Maximum stable bubble size Bubble diameter (rn) Breakage frequency (8") Liquid height in vessel (m) Collision frequency (s'l) Collision frequency (5") Final film thickness (m) Initial film thickness (m) Eddy wavenumber (m'l) Constant Overall mass transfer coefficient (m/s) Bulk liquid mass transfer coefficient (m/s) Shell mass transfer coefficient (m/s) Agitation speed (3") Number of bubbles Term used to simplify Equation 60 Radius of the liquid disk between bubbles (m) Gas constant (L*atm/K*mol) Length scale (m) 160 r, rb = Bubble radius (m) T = Vessel diameter (m) t = Time (s) i = Average contact time (s) u = Velocity (m/s) 1; = Average velocity (m/s) v,v’ = Bubble volume (m3) We = Weber number 9.6.1 Greek Symbols B(v,v’) = Probability density of breakage s = Energy dissipation (mZ/s) 2 = Average energy dissipation (mz/s) q; = Holdup 7L(d,d,) = Coalescence efficiency k(v,v’) = Coalescence efficiency u = Viscosity (Kg/m*s) n(v) = Number of bubble formed from breakage v = Kinematic viscosity (mZ/s) p = Density (kg/m3) o = Surface tension (N/m) 1 = Average coalescence time (s) 161 9.6.3 Subscripts and Superscripts c, l = Continuous phase (1 = Dispersed phase * = Dispersion (both phases) 162 10. SUMMARY AND CONCLUSIONS Microbubbles, or colloidal gas aphrons, have been characterized in terms of formation time, stability, gas void fraction, and bubble size. The minimum formation time was obtained at the critical micelle concentration for the surfactants tested and was unaffected by the addition of salts. Microbubble dispersions were most stable at the critical micelle concentration. The gas void fraction was found to be independent of surfactant concentration and approached the maximum packing limit for monosized spheres in a face-centered cubic cell, or 0.67. The average bubble size was measured to be approximately 60 um using dynamic light scattering techniques. Surfactants were tested for biocompatibility with Butyribacterium methylotrophicum. Tween, or polysorbate surfactants, did not significantly affect the grth or product formation patterns. Similar results were obtained for some Brij surfactants. An effect of chain length was noted on the toxicity of the surfactant, with shorter chain surfactants generally being more toxic than longer chains. Most ionic surfactants tested were toxic to the organism. The mass transfer properties of microbubbles were studied using both a steady- state method and an unsteady-state method in a large-scale fermenter. In the steady-state method, the axial dispersion for flow of microbubbles through a bubble column was much lower than for conventional bubbles. The steady-state KL values were similar in magnitude to literature values for small, rigid bubbles. Increasing concentrations of surfactant decreased the KL as much as 75%. This decrease was presumed to be due to shell thickening and steric hindrance. The resulting KLa values ranged from 200 h'] to 1100 h]. These values are an order of magnitude higher than those measured in synthesis 163 gas fermentations using conventional sparging. Dynamic mass transfer studies conducted in a 100 gallon fermenter showed that the majority of the available gas was transferred into the liquid phase. The corresponding KLa values were considerably higher than for conventional sparging at low power to volume inputs. The calculation of the power input required to generate microbubbles for synthesis gas fermentations gave a Power number of 0.036 for the microbubble generator and a corresponding P/V value of 0.01 kW/m3 of fermentation capacity. Microbubble-sparged synthesis gas fermentations were run using both B. methylotrohicum and sulfate reducing bacteria. The B. methylotrohz'cum fermentation yielded a six-fold improvement in the volumetric mass transfer coefficient over conventional sparging: 91 h'1 vs. 14 h], even through a lower gas flow rate was used for the microbubbles. A two-fold improvement in SRB reactor productivity was obtained upon switching from conventional bubbles to microbubbles. The maximum productivity obtained with the microbubble—sparged fermentation was similar to that obtained using a soluble substrate, in which there was no interphase mass transfer resistance. The coalescence of microbubbles was studied in a stirred vessel. The bubble size distribution was monitored over time using a high-speed video camera. A population balance equation approach was used to analyze the data. The coalescence efficiency was modeled using a film drainage model. Results from these studies indicated that decreased coalescence occurs with increased bulk liquid surfactant concentration. The decreased coalescence is thought to be due to shell thickening and steric and physical hindrances at higher concentrations of surfactants. 164 11. APPENDICES 165 11.1 Appendix A: Axial Dispersion The program used to evaluate the axial dispersion in the system is listed here with subroutine PATERN eliminated. A listing of the PATERN subroutine can be found in Thompson (1993). C C C C SUBROUTINE PROC FOR PROGRAM CALLED PATERN C C PURPOSE: TO ESTIMATE DIFFUSION COEFFICIENTS BY READING IN DATA C CONTAINING TIME -VS- CONCENTRATION VALUES. C C MAIN PROGRAM C C INTEGER NDATA,NP,NPASS,IO integer ip, i, nin, nres, nterm real pm(20), cdi(200), cwork(lO), toin, dtin, tores, dtres, S cdr(200), tau, p(10), step(10) common /mainl/cdi,toin,dtin common /main2/cdr,tores,dtres common /main5/nin,nres common /pbparm/pm common /main4/tau,nterm open(2,file='file2.csv',status='old') open(3,file='filel.csv',status='old') open(4,file=’file3.csv',status='old') OPEN (UNIT=30,FILE='FOR30.DAT',STATUS='NEW') ip=3 read(2,*) (pm(i), i=1riP) write(30,*) (pm(i), i=1rip) 300 format(8f10.0) read(3,*) toin,dtin write(30,*)toin,dtin 100 format(2f10.0) nin=0 110 format(10f8.0) I do 10 i = 1, 10 read(3,*) cwork(i) if (cwork(i) .lt. 0.0) goto 90 nin=nin+l if (nin .gt. 200) goto 90 cdi(nin)=cwork(i) write(*,*) cdi(nin) write(30,*) cdi(nin ) 10 continue goto I 90 nres=0 166 20 read(3,*) tores,dtres write(30,*) tores,dtres do 20 i=1,10 read(3,*) cwork(i) if (cwork(i) .lt. 0.0) goto 91 nres=nres+1 if (nres .gt. 200) goto 91 cdr(nres)=cwork(i) write(*,*) cdr(nres) write(30,*) cdr(nres continue goto 2 ) 91 read(4,*) tau,nterm 200 C C C C399 399 00000000000000 write(30,*) tau, nterm format(f10.0,i10) call norsig call fouexp SEARCH INITIALIZATION P(l)= 0.01 p(2)= 1.0e-3 STEP(1)=0.01 step(2)=1.0e—3 NP=2 NPASS=3 IO=3 START SEARCH CALL PATERN(NP,P,STEP,NPASS,IO,COST) SEARCH COMPLETE, PRINT RESULTS PRINT 399, P(l), P(Z), COST FORMAT(F10.3,10X,FIO.3) WRITE (30,399) P(l),P(2),COST FORMAT(EIO.3,5X,F6.3,5X,ElO.3) STOP END THIS FILE IS A PAIR OF SUBROUTINES WRITTEN TO BE COMPATIBLE WITH THE OPTIMIZATION SUBROUTINE PATERN. THEY SIMULATE A PROCESS USING DISCRETE DIFFERENCE EQUATIONS AND COMPARE THE SIMULATION OUTPUT WITH THE ACTUAL OUTPUT (READ IN THROUGH A DATA FILE), CALCULATING AN ERROR OR "COST" ASSOCIATED WITH THAT SIMULATION. PATERN USES THESE SUBROUTINES ITERATIVELY IN ORDER TO FIND THE OPTIMUM SET OF TRANSFER FUNCTION PARAMETERS TO FIT THE DATA. SUBROUTINE PROC(P,COST) dimension p(10) 167 common/pbparm/pm(20) common/mainB/error pm(2)=p(l) Pml3l=p(2) call precur cost=error RETURN END SUBROUTINE BOUNDS(P,IOUT) DIMENSION P(10), STEP(10) IOUT=O IF(P(1).LE.O) IOUT=1 if(p(2).le.0) iout=1 RETURN END subroutine norsig 10 20 30 integer znin,znres common /main1/ cdin,ztoin,zdtin common /main2/ cdres,ztores,zdtres common /main5/znin,znres common /signal/toin,dtin,nin,cin(200), tores,dtres,nres,cres(200) dimension cdin(200),cdres(200) nin=znin toin=ztoin dtin=zdtin nres=znres tores=ztores dtres=zdtres call normlz(dtin,nin,cdin,ain,cin) call normlz(dtres,nres,cdres,ares,cres) return end subroutine normlz(dt,n,cd,area,cn) dimension cd(200),cn(200) area = 0.0 do 10 i=1,n area=area+cd(i) continue area=area*dt if(area.eq.0.0) then write(*,600) do 20 i=1,n cn(i)=cd(i) continue return elseif(area.lt.0.0) then write(*,610) endif do 30 i=1,n cn(i)=cd(i)/area continue 168 return 600 format(lh0,'area of curve is zero(returned signal is not normal $ized).'/) 610 format(lh0,'area of curve is negative.'/) end subroutine fouexp common /main4/ztau,znterm common /signal/ toin,dtin,nin,cin(200), $ tores,dtres,nres,cres(200) common/fdata/ tau,nterm,aoin,ain(200),bin(200), $ aores,ares(ZOO),bres(200) tau=ztau nterm=znterm if(tau.lt.(tores+dtres*float(nres))/2.0) then tau=(tores+dtres*float(nres))/2.0*1.5 write(*,610) tau endif if(nterm.lt.1.or.nterm.gt.ifix(tau/dtres+0.5)) then nterm=ifix(tau/dtres+0.5) write(*,620) nterm endif if(nterm.gt.200) then nterm=200 write(*,620) nterm endif call foucoe(toin,dtin,nin,cin,tau,nterm,aoin,ain,bin) call foucoe(tores,dtres,nres,cres,tau,nterm,aores,ares,bres) return 610 format(lh0,'half period (tau) is replaced by',e15.7) 620 format(lh0,'number of terms (nterm) is replaced by',i5) end subroutine foucoe(to,dt,nt,cn,tau,nterm,ao,a,b) dimension cn(200),a(200),b(200) data pi/3.141593/ ao=0.0 sq=0.0 do 10 i=1,nt ao=ao+cn(i) sq=sq+cn(i)**2 10 continue ao=ao*dt/tau sq=sq*dt cotest=ao/2.0 sqtest=2.0*(ao/2.0)**2*tau do 20 n=l,nterm x=0.0 y=0.0 sv=float(n)*pi/tau do 30 i=1,nt z=sv*(to+float(i-1)*dt) x=x+cn(i)*cos(z) y=y+cn(i)*sin(z) 30 continue 169 20 4O 10 a(n)=x*dt/tau b(D)=y*dt/tau cotest=cotest+a(n) sqtest=sqtest+(a(n)**2+b(n)**2)*tau if(mod(n,10).eq.0) then rat=sqtest/sq endif continue return end subroutine precur common /main3/err common/signal/toin,dtin,nin,cin(200),tores,dtres,nres, cres(200) common/fdata/tau,nterm,aoin,ain(200),bin(200),aores, ares(200),bres(200) common/pbparm/ pm(20) dimension aclc(200),bclc(200) call clccoe(pm,tau,nterm,nt,aoin,ain,bin,aoclc,aclc,bclc) x=2.0*(aores/2.0-aocIc/2.0)**2 y=2.0*(aores/2.0)**2 do 40 i=1,nt x=x+(ares(i)-aclc(i))**2+(bres(i)-bclc(i))**2 y=y+ares(i)**2+bres(i)**2 continue err = sqrt(X/y) return end subroutine clccoe(pm,tau,nterm,nt,aoin,ain,bin,aoclc,aclc,bclc) dimension pm(20),ain(200),bin(200),ac1c(200),bclc(200) data pi/3.141593/ call transf(0,0.0,pm,an,bn) aoclc=aoin*an do 10 n=1,nterm w=float(n)*pi/tau call transf(l,w,pm,an,bn) aclc(n)=ain(n)*an+bin(n)*bn bclc(n)=bin(n)*an-ain(n)*bn if(an**2+bn**2.lt.1.0e-8) then nt=n return endif continue nt=nterm return end subroutine transf(ictl,w,pm,an,bn) dimension pm(20) real*8 d1,d2 save d1,d2 complex*16 cs,cphi,cex,coth,cf if(ictl.eq.0) then d1=Pm(3)/(pm(l)*pm(2)) d2=pm(1)/pm(2) 170 endif if(w.eq.0.0) then an=1.0 bn=0.0 return endif cs=cmplx(0.0,w) cphi=cdsqrt(l+4.0*d1*d2*cs) cex=1—cphi coth=1.0/(2.0*d1)*cex cf=cdexp(coth) an=cf bn=dimag(cf) return end 171 11.2 Appendix B: Bubble Coalescence The program used to evaluate the bubble coalescence in a stirred tank is listed below. The subroutine EDRIVB, the EPISODE B package, has been eliminated from the listing. A complete listing of EPISODE can be obtained from Byrne and Hindmarch (1976) C====================================================================== C= THIS PROGRAM CALCULATES THE BUBBLE COALESCENCE IN STIRRED TANK. = C: THE BUBBLE COLLISION FREQUENCY IS ESTIMATED FROM THE ISOTROPIC = C= TURBULENT FLOW. THE COLLISION EFFICIENCY IS ESTIMATED FROM = C: PRINCE AND BLANCH = C: WRITTEN BY COSTAS TSOURIS, MODIFIED BY MARSHALL BREDWELL - 1.16.98 = IMPLICIT DOUBLE PRECISION (A-H,O-Z) DIMENSION FO(300),YO(300),VY(300),SY(300),VD(300),IWORK(IOOO) DIMENSION D(300),CUMVY(300),CUMSY(300),R(300),WORK(4000) INTEGER*2 IHR,IHRO,IMIN,IMINO,ISEC,ISECO,IIOOTH COMMON/HISTRY/Y(I300) COMMON/MODS/TOL,YREL,HBIG,NITCON,LOUT COMMON/EPCOMl/T,H,HMIN,HMAX,EPSC,SS,UROUND,NC,MFC,KFLAG,JSTAR COMMON/EPCOMZ/YMAX(300) COMMON/EPCOM3/ERROR(300) COMMON/EPCOM4/SAVE1(300) COMMON/EPCOMS/SAVE2(300) COMMON/EPCOM6/PW(IOOOO) COMMON/EPCOM7/IPIV(300) COMMON/EPCOM8/EPSJ,NSQ COMMON/EPCOM9/HUSED,NQUSED,NSTEP,NFE,NJE COMMON/DATl/F(300,300),VOL(300),EFFRQ(300,300) DOUBLE PRECISION VC,PI,A1,A2,SIGMA,ED,RPM,DNSC,C3,C4,DIMP,HO,HOLD2 DOUBLE PRECISION D3,D2,CALL,DNSD,EFl,EF2,HOLDl,NUMBUB,DAVE2,TOT,D1 DOUBLE PRECISION EF3,EF4,COLFRQ COMMON/DAT3/EPSI,JMAX EXTERNAL FUNl PI=3.1415926 C C READ ALL REQUIRED PARAMETERS C OPEN(20,FILE='COLLEF.RES',STATUS='UNKNOWN') OPEN(21,FILE='COLLEF2.RES',STATUS='UNKNOWN') OPEN(22,FILE='POTENT.RES',STATUS='UNKNOWN') OPEN(23,FILE='PSTATS.RES',STATUS='UNKNOWN') OPEN(24,FILE='PBERES.RES',STATUS='UNKNOWN') OPEN(25,FILE='TESTOUT.RES',STATUS='UNKNOWN') OPEN(l,FILE='BUBBLE.DAT',STATUS='OLD') C CALL GETTIM(IHRO,IMINO,ISECO,IlOOTH) WRITE(*,'(20H PROGRAM STARTED AT ,2(I3,1H:),I3)')IHRO,IMINO,ISECO 172 READ(l,*)OUTT,NCL,ICL,TEND,DMIN,PINIT,VC,SIGMA,RPM,C3,C4,DIMP, + HO,DNSC,DNSD,EPSABS,EPSREL,LIMIT,(FO(K),K=1,ICL) CALCULATE ENERGY DISPERSION VIA BERTRAND ED=2*(O.81*((RPM/60.)**3)*((DIMP/lOO.)**2))*lOOOO TO=O.lE-5 HO=l.OD-4 MF=10 ML=NCL+l MU=NCL+1 VMIN=PI*(DMIN**3)/6. VMAX=VMIN*NCL*2 DMAX=(6.*VMAX/PI)**(1./3.) HS=VMIN DO 20 J=l,NCL VOL(J)=VMIN+(J-l)*HS D(J)=(VOL(J)*6./PI)**(l./3) R(J)=D(J)/2 CONTINUE GIVE INITIAL CONDITIONS VP=O. SDTD=O. DO 22 K=1,ICL Y0(K)=F0(K)*PINIT VP=VP+Y0(K)*VOL(K) SDTD=SDTD+Y0(K)*D(K) D3=D3+Y0(K)*(D(K)**3) D2=D2+Y0(K)*(D(K)**2) CONTINUE DAVEO=D3/D2 VSIM=VP*PDENS/PCONC WRITE(*,*)DAVEO DAVE=SDTD/PINIT PRATIO=1. TINIT=O. DO 24 K=ICL+1,NCL YO(K)=O. CONTINUE EPS=1.0 IERROR=2 TOUT=0.1D—5 INDEX=1 HBIG=O.1 YREL=EPS LOUT=5 NITCON=3 TOL=EPS DO 26 J=l,NCL SY(J)=YO(J)/SDT 173 VY(J)=VD(J) /VDT 26 CONTINUE CUMSY(1)=SY(1) CUMVY(1)=VY(1) DO 27 J=2,NCL CUMSY(J)=CUMSY(J-l)+SY(J) CUMVY(J)=CUMVY(J-1)+VY(J) 27 CONTINUE WRITE(23,210)TO,(D(K),VOL(K),SY(K),CUMSY(K),VY(K),CUMVY(K), + K=1,NCL) C C CALCULATE COALESCENCE FREQUENCY C DO 28 J=l,NCL DO 30 K=J,NCL Al=D(J) A2=D(K) HOLD1=(O.5*(l./R(J)+l./R(K)))**(-l.0) EFl=EXP(-C3*(HOLDI**3*DNSC/(16*SIGMA))**(l./2.)/ + (HOLD1**(2./3.)/(ED**(l./3.)))) HOLDZ=(SIGMA)*(Al**(2./3.)+A2**(2./3.)) EF=EF1 EF3=1.07*ED**(2./3.)*Al**(2./3.) EF4=l.O7*ED**(2./3.)*A2**(2./3.) COL=3.1415926/4.0*(Al+A2)**2.0*(EF3+EF4)**(l./2.) F(J,K)=COL*EF P(KIJ)=F(JIK) WRITE(20,212)A1,A2,COL,EF,F(J,K) 3O CONTINUE 28 CONTINUE 212 FORMAT(F20.14,3X,F20.l4,3X,F20.l4,3X,F20.l4,3X,F20.14,3X,F20.l4, + 3X,F20.l4) C C 10 CALL EDRIVB(NCL,TO,HO,YO,TOUT,EPS,IERROR,MF,INDEX,ML,MU) SDT=O. VDT=O. D3=0. D2=0. D1=0. RDT6=O. SDTD=0. tot=0. DO 68 K=1,NCL SDT= SDT+Y0 SDTD= SDTD+ D3= D3+Y0(K ( Y D ) ) D2= D2+YO(K ) ) ) ( ) ( K) (K) D( **3) D( **2) D1: D1+YO(K (K TOT= TOT+YO VD(K )= O