TRANSPORT OF A BACTERIAL AEROSOL IN TURBULENT MIXING REGIONS Thesis for ”W Degree of ph. D. MIEEISM STATE UNIVERSITY Dennis Ray I-EeIdman 1965 “mats (”9v .’ 'II A.“ (~— fi L I B R A R I‘v’iichignL 751 Univex‘sirj C ——<... ABSTRACT TRANSPORT OF A BACTERIAL AEROSOL IN TURBULENT MIXING REGIONS by Dennis Ray Heldman The knowledge of transport mechanisms of bacterial aerosols is, of primary importance when analyzing air- borne contamination control methods for food processing and packaging operations. This investigation was con- ducted to determine the transport characteristics of a bacterial aerosol under conditions simulating two adjacent rooms with different aerosol concentrations. The experi- mental investigation involved determination and description of transfer coefficients for transport of an aerosol through an opening in a partition between two 64 ft3 compartments with different uniform aerosol concentrations. Theoretical description and analysis of experimental data involved the use of a mixing region model, which described turbulent dispersion of aerosol through the opening and entrainment in the low concentration compartment. Air velocity fluctuations were detected by a hot wire anemometer and recorded for several locations and flow conditions. Information on turbulent energy, eddy diffusivi- ties and dispersion of the aerosol in the mixing region was Dennis Ray Heldman obtained by means of a statistical analysis of the -air velocity records. Experimental data were collected for equal air flow rates from 20 to 60 ft3/min through both compartments of the aerosol chamber. In addition, tests were conducted with air/flow rate gradients as high as 40 ft3/min in the same direction or opposite the concentration gradient. Transfer coefficients and turbulent energy increased sig- nificantly as air flow rate increased from 20 to A0 ft3/min, whereas the intensity of turbulence was relatively constant over the entire range investigated. For aerosol flow rates above 30 ft3/min, transfer coefficients were maximum when air flow rates through both compartments were equal. The area of the opening between the two compartments was varied by changing the vertical height to 3, 6, and 9 in. Decreasing transfer coefficients with increasing vertical height of opening were related to the shape of the mixing region profiles. The partition width at the initial point of mixing influenced the transfer coefficients slightly. An increase in transfer coefficient occurred as the width was increased from 0.0625 to 0.3125 in., but the coefficient decreased as the width was increased to 0.5625 in. Slight increases in turbulence due to increased partition width and a reduction in mixing region width were factors involved in the explanation of this relationship. Dennis Ray Heldman The influence of temperature gradients on transport characteristics was determined by heating the air in one compartment of the aerosol chamber. Transfer coefficients increased consistently as the temperature gradient was increased from -lA° to +12.5°F A dimensionless relationship was derived, based on the turbulent mixing region model, and used to present all data obtained in the investigation. Approved png M Major Professor and Department Chairman 2L.“ 39,/465 Date TRANSPORT OF A BACTERIAL AEROSOL IN TURBULENT MIXING REGIONS By Dennis Ray Heldman A THESIS Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Agricultural Engineering 1965 ACKNOWLEDGMENTS The guidance and leadership of Dr. C. w. Hall, chairman of the Agricultural Engineering Department, is gratefully acknowledged. The inspiration provided through- out this portion of my graduate program and during the course of this investigation has made it a pleasing and rewarding experience. Sincere appreciation is extended to Dr. T. I. Hedrick (Food Science) for supplying facilities required for conducting experiments and for providing advice and suggestions whenever requested. Additional acknowledgment is offered to Dr. K. L. Schulze (Sanitary Engineering) and Dr. F. H. Buelow (Agricultural Engineering) for serving as guidance committee members and providing advice whenever needed. The unfailing support and encouragement provided by my wife, Joyce, has supplied inspiration needed throughout my undergraduate and graduate education. ii To: Joyce, Cynthia and Candace Mr. and Mrs. M. L. Heldman Mr. and Mrs. H. S. Anspach iii TABLE OF CONTENTS Page ACKNOWLEDGMENTS . . . . . . . . . . . . . ii LIST OF TABLES. . . . . . . . . . . . . . vi LIST OF FIGURES . . . . . . . . . . . . . viii LIST OF APPENDICES . . . . . . . . . . . . xi NOMENCLATURE . . . . . . . . . . . . . . xii Chapter 1. INTRODUCTION 1 2. LITERATURE REVIEW u 2.1. Characteristics of Air-Borne Bacteria. A 2.1a. Existence. . u 2.1b. Sedimentation and Deposit 5 2.10. Viability. . . . 7 2.1d. Sampling 9 2.2. Production and Study of Uniform Bacterial Aerosols . . . . . . . . . 11 2.2a. Production of the Aerosol . . . 11 2.2b. The Aerosol Chamber . . . . . 13 2.3. Transport Processes. . . . . . . . IA 2.3a. Turbulence . . . . . . 14 2.3b. Turbulent Diffusion . . . . 16 2.3c. Transport in Mixing Regions. . . 25 3. THEORETICAL CONSIDERATIONS . . . . . . . 29 3.1. Basic Diffusion Equations. . . . . . 32 3. 2. Turbulent Diffusion. . . . . . 37 3. 3. Turbulent Transfer Coefficient . . . . 39 3.4 Eddy Diffusivity . . . . . 41 3.5 Transport in the Mixing Region . . . . ug iv 3.5a. Isovel- Isothermal Flow. . 3.5b. Turbulent Mixing with Unequal Velocities 3.50. Influence of Partition Width. 3.5d. Turbulent Mixing with a Temperature Gradient. 4. EXPERIMENTAL PROCEDURES AND EQUIPMENT 4.1. Equipment 4.1a. Aerosol Chamber 4.1b. Aerosol Generation 4.1c. Air Sampling . . 4.1d. Air Velocity and Turbulence 4.2. Bacteriological Methods. 4.2a. Handling Techniques. 4.2b. Plating and Counting 4.3. Scope of Tests. 4.3a. Turbulent Transfer Coefficient 4.3b. Eddy Diffusivities . . 4.3c. General Procedures 4.4. Analysis of Turbulence Data 5. RESULTS AND DISCUSSION 5.1. Characteristics of the Aerosol Chamber. 5 Turbulent Diffusion . . . . 5.2a. Turbulent Energy. . 5.2b. Autocorrelation Coefficients. 5.3. Concentration Distributions 5.4. Transport of the Bacterial Aerosol 5.4a. Experimental Transfer Coefficients. 5.4b. Dimensionless Relationships 6. CONCLUSIONS. 7. RECOMMENDATIONS FOR FUTURE WORK APPENDIX REFERENCES Page 43 46 46 48 48 48 E J 58 58 60 60 61 61 63 63 64 67 67 73 73 85 86 86 92 102 105 106 . 126 LIST OF TABLES Influence of relative humidity on death rate constants of Serratia marcescens. . Summary of results Pressure drop, air flow rate and mean air velocity. Influence of air flow rate on transport charactgristics of air-borne bacteria with 1.5 ft opening Influence of air flow rate gradient on trans- port characteristics of air-borne bacteria with 1.5 t 2 opening and aerosol flow rate of 30 ft /min. . . Influence of air flow rate gradient on trans- port charactsristics of air-borne bacterial with 1.5 t opening and aerosol flow rate of 40 ft /min. . Influence of air flow rate gradient on trans- port charactSristics of air-borne bacteria with 1.5 t opening and aerosol flow rate of 50 ft /min. . . Influence of width at initial point of mixing on transport characteristics of air-borne bacteria" at 50 ft 3/min Influence of width at the initial point of mixing on transport characteristics of air— borne bacteria at 40 ft /min. . Influence of width at initial point of mixing on transport characteristics of air-borne bacteria at 30 ft 3/min Influence of air flow rate gradient on trans- port characteristics of air-borne bacteria with 0.3125 in. partition vi Page 80 108 112 113 114 115 116 . 117 . 118 . 119 Table Page A.10. Influence of air flow rate gradient on trans— port characteristics of air—borne bacteria with 0.5625 in. partition. . . . . . . 120 A.11. Influence of air flow rate gradient on trans- port charac eristics of air-borne bacteria with 1.0 ft opening . . . . . . . . 121 A.12. Influence of air flow rate gradient on trans- port charac eristics of air—borne bacteria with 0.5 ft opening . . . . . . . . 122 A.13. Influence of temperature gradient on trans— port characteristics of air-borne bacteria at 61 ft3/min. . . . . . . . . . . 123 A.14. Turbulence data at various locations and flow conditions . . . . . . . . . . . 12“ A.15. Dispersion data at various locations and flow conditions . . . . . . . . . . . 125 vii Figure 4.1. 4.3. 4.4. 4.5. 4.6. 4.9. 4.10. 4.11. 4.12. 5.1. 5.3. 5.4. LIST OF FIGURES viii Page Temperature and relative humidity recording instruments . . . . . . . . . 49 Schematic diagram of aerosol chamber and related instrumentation 50 Partition and opening between compartments of aerosol chamber 51 Air diffuser in ceiling of one compartment of aerosol chamber 51 Ultra—high efficiency air filter 53 Venturi tube and micromanometer used for air flow measurement 53 Casella slit air sampler. 55 Prechamber and related parts of aerosol gen- eration system 55 Aerosol atomizer and culture reservoir 57 Air sampling station and point of aerosol injection into duct 57 Hot wire anemometer and auxilary equipment. 59 X-Y recorder. 59 Dynamic emptying characteristics of aerosol chamber 68 Variation of turbulent energy and intensity of turbulence with air flow rate . . . 71 Influence of transverse location on intensity of turbulence. 71 Variation of turbulent energy in transverse direction . . 74 Figure Page 5.5. Agreement of experimental data with equation 3.16. . . . . . . . . . . . . . 74 5.6. Turbulent energy profiles with 70 ft3/min on high air flow rate side . . . . . . 76 5.7. Turbulent energy profiles with 60 ft3/min on high air flow rate side . . . . . . . 76 5.8. Turbulent energy profiles with 50 ft3/min on high air flow rate side . . . . . . 76 5.9. Autocorrelation coefficients as influenced by time lag . . . . . . . . . . . . 79 5.10. Variations in apparent mixing region width with air flow rate and partition height. . . . 79 5.11. Concentration distribution in mixing region . 87 5.12. Variation of transfer coefficient with air flow rate . . . . . . . . . . . . 89 5.13. Influence of air flow rate gradient on trans- fer coefficient . . . . . . . . . . 89 5.14. Influence of temperature gradient on trans- fer coefficient . . . . . . . . . . 91 5.15. Influence of temperature and relative humidity on death rate of Serratia marcescens . . . 91 5.16. Influence of aerosol viability on the relation- ship between temperature gradient and trans- fer coefficient . . . . . . . . . . 91 5.17. Dimensionless transport as influenced by turbulence, dispersion time and mean air velocity . . . . . . . . . . . . 93 5.18. Variation in dimensionless transport with air flow rate gradient for aerosol flow rate of 50 ft3/min. . . . . . . . . . . . 96 5.19. Variation in dimensionless transport with air flow rate gradient for aerosol flow rate of 40 ft3/min. . . . . . . . . . . . 96 ix Figure Page 5.20. Variation in dimensionless transport with air flow ate gradient for aerosol flow rate of 30 ft /min. . . . . . . . . . . . 96 5.21. Influence of partition width on dimensionless transport . . . . . . . . . . . . 98 5.22. Influence of opening height on dimensionless transport . . . . . . . . . . . . 98 5.23. Influence of the dimensionless temperature ratio on dimensionless transport . . . . 98 5.24. Experimental correlation of dimensionless groups involved in bacterial aerosol trans— port. . . . . . . . . . . . . . 101 A.l. Mean air velocity profile . . . . . . . 110 Appendix A.1. LIST OF APPENDICES Air Flow Measurement Solution to Diffusion Equation Mean Air Velocity Profile. Experimental Turbulent Transfer Coefficients xi Page 107 108 111 111 NOMENCLATURE apparent width of aerosol mixing region defined by equation (3.36), in. partition width at initial point of mixing, in. bacterial aerosol concentration, No./ft.3 bacterial aergsol concentration at time equal to zero, No./ft. apparent Eddy diffusivity presented in equation (3.3), in /min. specific egdy diffusivity defined by equation (3.32), in /min. air flow rate, ft3/min. air flow rate with equal flow in both compart— ments of aerosol chamber, ft /min.. integral of complementary error function defined in equation (3.47), in. vertical height of Opening between compartments of aerosol chamber, in. vertical height of aerosol chamber. length scale of turbulence defined in equation (3.34), in- death rate constant of bacterial aerosol, l/min. experimental constant in equation (3.42). turbulegt transfer coefficient, No./ft2,min (No./ft ). mass flux of bacterial aerosol, No./min. autocorrelation coefficient defined in equation (3.21) xii ‘14 0‘t aerosol dispersion time, min. time required for R(§) to approach zero, sec air velocity parallel to x—axis, ft/min. turbulent energy factor for longitudinal direction, ft2/min2 air velocity parallel to y—axis, ft/min. turbulent energy factor for transverse direction, ftZ/min2 volume of one compartment in aerosol chamber, ft3 rectangular coordinate parallel to vertical direction in aerosol chamber rectangular coordinate transverse to opening between compartments of aerosol chamber statistical dispersion length defined by equation (3.25). 1m? turbulent thermal diffusivity, in2/min. coefficient defined in equation (3.40). -temperature, °F. absolute temperature, °R. kinematic viscosity, in2/min turbulent viscosity, ine/min time lag in instantaneous velocity correlations, sec Bar above symbol indicates mean component Prime above symbol indicates instantaneous component Subscripts: H refers to compartment with highest magnitude L refers to compartment with low magnitude xiii 1. INTRODUCTION A small percentage of the 1,500,000 particles in an average cubic foot of air are viable microorganisms. How- ever, the existence of these viable bacteria in the air of dairy and food processing plants is of greater importance than the small percentage indicates. Reports on the populations of air—borne microorganisms are varied both in approach and results. Olson and Hammer (1934) and Cerna (1961) reported counts as high as 12 and 55, respectively, settling on a standard size petri dish per minute in various areas of dairy plants. Labots (1961) and Heldman §t_a1. (1964) reported mean counts of 5 to 85 per ft3 when using vacuum slit samplers. The fact that populations of air-borne microorganisms exist in food processing areas can be attributed to the many sources present in these areas. Isolation of floor drains as a source of air— borne microorganisms during flooding is just one example (Heldman 92:11., 1965). The existence of an air-borne microorganism popula— tion of any magnitude provides a chance for air-borne contamination of exposed products. In many cases, this contamination may occur after processing and will result in significant reductions in product shelf-life. The importance of this type of contamination is increasing significantly due to the prOSpects of packaging sterile products such as milk and other milk products. Contamin— ation of the sterile product with a single microorganism will result in an unacceptable storage life for these products. The deveIOpment of ultra—high efficiency or ABSOLUTE filters for air may solve at least part of the air-borne contamination problem. These filters, which are designed to remove 99.97% of all 0.3 micron particles, will provide air which is practically free of air-borne microorganisms. However, the more difficult and unsolved problem is that of secondary contamination or contamination of the filtered air from the many sources of air-borne microorganisms in the processing plant. A partial solution to the latter problem is localized control by use of "laminar air flow" (Whitfield, 1963) or jets of filtered air to protect selected spaces. These mthods are limited, however, to small spaces and much is unknown about the effectiveness of laminar air flow in the mixing regions. Before complete control of air—borne contamination can be attained, basic information on the transport of air-borne microorganisms from the source to the product must be obtained. Within a room, the movement and flow patterns of the air is of major concern. However, when considering transport of air-borne microorganisms from one room to an adjoining room, factors such as the mixing of two air streams at openings between rooms is of importance. Unless there is considerable momentum transport or air movement through the Openings, the mixing of the air streams must provide the major portion of the transport. In addition, the mixing of a high concentration air stream with a low concentration air stream may be influenced by other vari— ables such as flow conditions of the air streams, geometry at the initial point of mixing and differences in tempera— ture and relative humidity between the air streams. The purpose of this investigation is to determine the transport characteristics of air—borne microorganisms in a mixing region which would simulate that encountered between two rooms with different air-borne concentrations. The results obtained should not be limited to food and dairy processing plants because of the wide-spread interest in the same subject in hospitals and "dust free" rooms. Mixing regions are also encountered in many contamination control devices, and results obtained in this investigation may lead to improved control methods. 2. LITERATURE REVIEW 2.1 Characteristics of Air-Borne Bacteria The develOpment of procedures and techniques for the study of air—borne bacteria has occurred, to the greatest extent, in the last 15 to 20 years. The stimulation of these developments has been related mostly to: (a) increased frequency of air—borne infection by antibiotic resistant strains of bacteria in hospitals and (b) the possibilities of biological warfare. 2.1a Existence Air—borne bacteria exist as aerosols, which are defined as liquid or solid particles in air. According to Wolf, g§_gl. (1959), the biological particles may exist in any of the following forms: (a) single unattached cells, (b) clumps composed of a number of microorganisms, (c) cells adhering to a dust particle, or (d) a free floating microorganism surrounded by a film of dried organic or inorganic material. In addition, the microorganism involved may be a vegetative cell or a spore. The relative importance of vegetative cells as compared to spores depends on the air space involved. Wolf, et_al, (1959) indicates that vegetative cells are of greater importance when con— cerned with communicable diseases. However, air—borne 4 contamination of a processed food by spores is of equal importance, since the conditions are usually ideal for germination. 2.1b Sedimentation and Deposit An air—borne microorganism is subjected to the influence of gravity and the motion of the surrounding fluid. Most evidence (Decker, et_al:, 1962) indicates that air-borne microorganisms, other than viruses, are 0.3 micron in diameter or larger. According to Wells (1955), particles of this size will settle from the air in a manner described by Stokes's law. Tanner (1963) and DallaVallo (1948) express this law as: (9 —ng) s dzc vg -. P 91817 (2.1) where: vg = terminal velocity of aerosol particles pp = density of aerosol particles pg = density of air, g = gravitational constant d = diameter of aerosol particles u = viscosity of air The Cunningham "slip" correction factor (0) is prOportional to the mean free path of the gas molecules and becomes increasingly important for particles less than 20 microns in diameter. Tanner (1963) discusses the difference between quiescent and turbulent aerosols and indicates that Stokes equation (2.1) will apply in both conditions. In general, a quiescent aerosol will "fall-out” at the constant rate of particle fall described by Stoke's equation (2.1). A turbulent aerosol possesses a constant rate of fallout of particles, which is prOportional to the rate of particle fall (vg) and the aerosol concentration. Tanner (1963) assumes that deposit on walls and ceiling of a chamber is negligible, and develOped the following equation for evanescence of a quiescent aerosol: C v t . (5) =1- a.— M O Q Both Wells (1955) and Tanner (1963) described the sedimen- tation of a turbulent aerosol as: (C) ( V—E—t) ( > — = 1 - exp - 2.3 where: C = concentration of aerosol at any time, t CO = concentration of aerosol at t = O. vg = terminal velocity of aerosol particles. h = height of aerosol chamber The fact that aerosol particles will deposit when subjected to certain conditions is demonstrated by Porter, et a1. (1963) while studying the decay of an aerosol moving through a duct. Experimental results indicated that this decay or deposit was a function of particle size, velocity and duct size. 2.10 Viabiligy Although some air—borne spores may have nearly unlimited Viability, the viability of a vegetative cell will be limited depending on the conditions to which it is exposed. The death of vegetative cells is expressed by Wells (1955) in the following manner: Ln <%—>= —Kt (2.4) .0 where: K = death rate constant t = time Here the death rate constant (K) is dependent on many factors such as bacterial species, air temperature and relative humidity. The factors affecting the viability of air-borne bacteria have been studied in detail by Webb (1959). When aerosols consisted of bacterial cells from distilled water, the death of the cells appeared to occur in two stages; rapid loss in viability during the first second followed by a slower death rate which obeyed first order kinetics at low relative humidities. The results suggested that death of the cells resulted from movement of water molecules in and out of the cell, in an equilibrium system, resulting in a collapse of the natural structure of cellular protein. Wells (1955) indicated that, in general, the initial death rate is higher in dry air, but longevity of survivors appears to be greater in dry air than in moist air. Experimental determination of death rate constants was conducted by Kethley, et a1. (1957) for bacterial aerosols of Serratia marcescens. The determinations were made in an aerosol chamber and by use of equation (2.4). The results obtained are presented in Table 2.1 TABLE 2.1.--Influence of relative humidity on death rate constants of Serratia marcescens. Washed Cells Cells Dispersed from Dispersed from Water 0.3% Beef Extract Broth Relative Humidity Ave. K S. E. Ave. K S. E. % (l/min) (l/min) 16 0.021 0.0008 20 0.032 0.0020 0.020 0.0050 25 0.040 0.0010 0.020 40 0.060 0.0030 0.025 52 0.025 0.0003 60 0.044 0.0020 0.032 0.0020 80 0.008 0.0003 90 0.036 0.0020 95 0.021 0.0010 The results (Table 2.1) illustrate the influence of relative humidity on death rate constants for Serratia marcescens dispersed from different types cd' aqueous'nmdia. The types of media were selected to represent the conditions KO surrounding an air-borne bacteria. The results indicate that the maximum death rate for both plain bacterial cells and cells surrounded by proteinaceous material occurs between 40 and 60% relative humidity. However, it is evident thatcmfljjsdispersed from beef extract broth had lower death rates thancxiKHSdispersed from distilled water, indicating a protective influence of the material surrounding the cell. 2-1d flew The methods available for sampling air—borne micro- organisms are very similar to those used for other air- borne particles. According to Wolf, et_al, (1959), the methods can be grouped as follows: (a) impingement in liquids, (b) impaction on solid surfaces, (c) filtration, (d) sedimentation, (e) centrifugation, (3) electrostatic precipitation and (g) thermal precipitation. All methods used for air—borne particles must be modified to allow for recovery of living biological particles. The methods for sampling and evaluation of air— borne biological particles are discussed by Wolf, gtual. (1959). In general, methods employing the impingement on liquid principle have very high efficiencies for collecting and enumerating the bacterial cell suspended in air. However, high impingement velocities may result in losses of viability of vegetative cells. Methods 10 which sample air by use of filtration will provide collection efficiencies as high as the efficiency of the filter media. The methods tend to be somewhat more suitable for spores and other resistant microbial forms since vegetative cells may not resist desiccation associated with filter collection. Sedimentation is probably the most inexpensive and simple method for determining the microbiological quality of air. However, the method has the serious disadvantage of being selective for larger air—borne particles and not the entire particle—size distribution. In addition, the influence of air movement prevents an accurate correction of this factor. For the results obtained by the sedimenta- tion technique to be of quantitative value, the aerosol must be allowed to settle quiescently onto a collecting surface in a closed container. Due to the long periods of time required for settling, vegetative cells lose viability before reaching the surface. Centrifugation was one of the first successful methods develOped for quantitatively evaluating air-borne bacterial populations (Wells, 1933). Although the effic- iency may be very high, it is dependent on Operating conditions and particle size. Two of the more complex sampling methods described by Wolf, gt_al. (1959) are the electrostatic and thermal precipitation samplers. The electrostatic unit provides high collection efficiency at relatively high sampling rates, but the instrument is 11 complex and requires careful attention to maintain accuracy. Thermal precipitation samplers are complex also and sampling rates are very low. Wolf, et;al. (1959) describes impaction samplers as best adapted to determining the concentration of particles which contain bacterial cells. One of the impaction samplers commonly used is the slit sampler which allows collection of the viable particles on an agar surface. Bourdillon, g£_al. (1941) found the slit samplers to be highly efficient for the smallest bacteria-carrying particles, under the proper conditions (air flow, slit width, and distance of slit from surface). 2.2 Production and Study of Uniform Bacterial Aerosols To study the characteristics of a bacterial aerosol, two factors are desirable: (a) production of bacterial particles which are similar to those normally present in airvborne populations and (b) control of aerosol concen— tration and distribution by using an aerosol chamber. 2.2a Production of the Aerosol According to Greene (1965), it is impo ible to (.11 ()1 experimentally produce an aerosol which simulates normal conditions because of the manner in which air—borne micro- organisms normally exist. Decker, et a1. (1962) reviews methods used to produce bacterial aerosols such as small 12 glass or plastic atomizers. With such devices, it is possible to produce aerosols containing a high percentage of particles approximately 1 micron in diameter. Wells (1955) points out that production of an aerosol involves two stages: (a) atomization involving the formation of liquid droplets containing bacterial cells and/or associate particles and (b) evaporation of the liquid as described by Raoult's Law. Although Greene (1965) indicated that more experimental work is being conducted using lyOphilized culture powders, Kethley, et_al, (1956) and Porter, et_al, (1963) have had reasonable success by atomizing liquid cultures into a prechamber. The prechamber provided con- ditions for evaporation of liquid portion of the particles and sedimentation of large particles. Using this technique, with 0.3% beef extract culture media, Kethley, et a1. (19 6) \J I were able to produce aerosols with average particle sizes ranging from 1.8 microns at 25% relative humidity to 2.3 microns at 80% relative humidity. The aerosol contained particles which had no more than two bacterial cells with 90% of the particles containing only one cell. Porter, 394a1. (1963) reported experiments with aerosols which con- tained between 1 and 8 microns when using similar techniques. In general, Kethley, e£_al, (1957) concluded that aerosol particle sizes could be predicted on the basis of the atomized drOplet volume, the concentration of solids or low vapor pressure liquids in the dispersed media and the response of the components to relative humidity. 13 2.2b The Aerosol Chamber The more prevalent and well-known uses of aerosol chambers include the study of air-borne infection (Druett and May, 1952; Henderson, 1952; Laurell, g£_al., (1949; Leif and Krueger, 1950; Robertson, 91:31., 1946; Rosebury, 1947; Urban, 1954; Weiss and Stegeler, 1952) and the effec- tiveness of various germicidal agents (DeOme et al., 1944); Kaye, 1949; Mackay, 1952; Rentschler, 1942; Twort, gE;al., 1940) However, chambers designed for the mentioned pur- poses are not well adapted for studying the nature and composition of bacterial aerosols during long time trials. A chamber which is well suited to the latter purpose was designed by Kethley, et_al. (1956). Experimental results indicated that the chamber designed would allow an increase in aerosol concentration according to the standard ventilation equation (Silver, 1946): C = CO [1 — exp (- 2%)] (2.5) and would produce disappearance of the aerosol by a similar equation: C = C0 [exp (— $91)] (2.6) Kethley, et a1. (1957) Proved that the aerosol concentra- tion in the chamber was uniformly distributed by deter- mining concentrations at points throughout the chamber. In addition, the ability of the chamber to maintain a l4 consistent concentration for long periods of time was tested by sampling at intervals up to 130 minutes. 2.3 Transport Processes Crank (1956) states that diffusion is the process by which matter is transported from one part of a system to another as a result of random molecular motions. Hinze (1959) shows that transport of a transferable quantity by random fluid motion is diffusive in nature. Since turbu— lent fluid motion is a random fluid motion, the transport of matter in a turbulent fluid must involve both molecular and turbulent diffusion (Frenkiel, 1953). Crank (1956) states Fick's law of diffusion as: F = _ D 3? (2.7) where F represents mass flux and, in the case of diffusion due to turbulent motion, D represents the sum of the molecular diffusion coefficient and the eddy diffusivity. 2.3a Turbulence Hinze (1959) presents the following definition: "Turbulent fluid motion is an irregular condition of flow in which the various quantities show a random variation with time and space coordinates, so that statistically distinct average values can be discerned." As many authors (Hinze, 1959; Schlichting, 1960) indicate, there are distinct differences between turbulence generated by friction forces 15 at fixed walls (wall turbulence) and that generated by the flow of layers of fluid with different velocities past or over one another (free turbulence). Most frequent theoretical approaches to the study of turbulence involves the assumption of isotropic turbulence (Hinze, 1959; Townsend, 1956; Frenkiel, 1953). Although this type of flow is ideal and does not exist except under local conditions, its value is that it may provide a fundamental basis for the study of the types of turbulence which actually exist. According to Frenkiel (1953) the term isotropic implies that the statistical characteristics of the flow will be invarient under rotation or reflection of the axes. This basic approach to the study of turbulence was begun by Taylor (1921) and has resulted in a definite trend toward the study of the statistical properties of turbulence rather than the use of phenomenological theories, which describe the influence of mean flow only. The term homogenious turbulence is frequently used to describe the turbulent field in which the statistical characteristics are not changed by translation of the axes (Frenkiel, 1953). This type of turbulence exists in real situations, but will form a part of a nonisotropic or aniso— trOpic turbulence. As revealed by Hinze (1959), the contri— butions to the theory of nonisotropic are small due to the extreme complexity of the problem. 16 According to Townsend (1956), nonhomogenious turbulent flow exists primarily in the following: (a) in the boundary regions between a field of homogenious turbulence and an adjacent undisturbed field, (b) when turbulent intensities and other quantities are symmetrical about a plane and (c) when turbulent intensities and other quantities are axisym- metric. As would be expected, theoretical development of inhomogenious turbulence is at an early stage. 2.3b Turbulent Diffusion The basic transport mechanisms of momentum, heat and mass in turbulent flow are very similar. According to Pasquill (1962), the theoretical treatment of the turbulent diffusion of these quantities has proceded according to two approaches: (a) transfer theory in which the transport rate is proportional to a concentration gradient with a prOportionality factor or constant and (b) statistical description in which an analytical technique for representing the history of the fluid elements in terms of the statistical prOperties of the turbulent motion is used. Schlichting (1960) points out that the first approach (transfer theory) involves calculations based on empirical hypothesis which endeavors to establish a relationship between the Reynold's stresses produced by mixing motions, and the mean values of the velocity components together with suitable hypothesis concerning heat and mass transfer. 17 According to Hinze (1959), the more complete and correct solution to turbulent flow problems can be obtained by expressing the turbulent-transport rate of the transferable quantity completely in terms of statistical functions of the turbulent velocity field and of boundary or initial conditions. The basic concepts of transfer theory in turbulent flow were introduced by Boussinesq (1877) who described Reynold's stress in turbulent flow by: _ au- Tt — AT -—-— (2.8) :<; where AT is called a mixing coefficient which is dependent on the mean velocity. A similar relationship for transfer of mass in turbulent flow: a; F = - 3 K —§ (2-9) was prOposed by Pasquill (1962), with K equal to the pro- portionality constant and 3% equal to the concentration gradient. In order to use the preceding equations, it is necessary to have knowledge of the manner with which the coefficients vary with the mean velocity or other measure— able quantity. One of the most useful concepts in the description of turbulent-transport processes is the "mixing length" theory introduced by Prandtl (1925). Prandtl‘s mixing length hypothesis is: du = 2 __ — 18 where T is shear stress and A is a mixing length described as the distance in the transverse direction which must be covered byznlagglomeration of fluid particles traveling with its original mean velocity in order to make the difference between its velocity and the velocity in the new lamina equal to the mean transverse fluctuation in tur- bulent flow. As indicated by Hinze (1959), the "mixing length" does not describe turbulent flow entirely correctly but does provide a useful tool for calculation purposes. A second and similar transfer theory was introduced by Taylor (1915). This second theory differs from Prandtl's momentum transport theory in that it describes the diffusion of vorticity rather than momentum. Schlichting (1960) indicates that this theory has particular application in free turbulent flow. Taylor (1932) has shown that the momentum transport and vorticity transport theories agree when turbulent motion is two—dimensional and confined to the plane perpendicular to the mean motion. However, when turbulent motion and mean motion are confined to two-dimen— sions, the results of the two theories differ significantly. The differences between the two theories are evident again when comparing velocity and temperature or concentration distributions in wakes or free turbulent mixing regions. The momentum transport theory would predict the distribu- tions to be identical. Taylor's vorticity transport theory l9 predicts different distributions which have been confirmed experimentally by Fage and Falkner (1932). If the turbulent flow is assumed homogenious and isotrOpic, solutions to the parabolic equation of diffusion become: very useful. This equation, as presented by Pasquill (1962), 19:2. LC _3_ 9.9 .‘L .22 at 3x (xx 3x + By <§y 8y) + 32 (32 82) (2°11) with Kx’ Ky, Kz equal to eddy diffusivities in various coordinates, can be adapted and solved for a given set of conditions. Several such solutions for point or line sources of concentration are presented by Frenkiel (1953) and Hinze (1959). Experimental investigations involving turbulence and turbulent diffusion have been concerned mainly with the description of eddy diffusivities in terms of mean flow. Towle and Sherwood (1939) determined eddy diffusivities for turbulent flow in ducts and found that values increased prOportionately to Reynolds' number. In addition, the results indicated that the scale of turbulence in free duct flow is significantly larger than that produced by a grid. Results presented by Sherwood and Woertz (1939) show that the eddy diffusivity is essentially constant over the central portion of a turbulent gas stream in a duct. The product of eddy diffusivity and gas density was found to be about 1.6 times larger than the eddy 20 viscosity. A semi-theoretical relationship was established relating the eddy diffusivity to mean velocity, duct radius and the friction factor. The importance of having knowledge of the scale of turbulence is revealed in investigations by Kalinske and Pien (1944). Results revealed that the eddy diffusivity is directly related to the scale of turbulence, which must be measured to provide an accurate prediction of the turbulent diffusion. A statistical description of the turbulent field is required in order to accurately solve a turbulent transport problem. The basic concepts involved in statis— tical theory of turbulence were introduced by Taylor (1921) and develOped by Taylor (1935). In general, the statistical Concept involves the correlation between the velocity of a particle at one time and that of the same particle at a later time, or the correlation between simultaneous velocities at two fixed points. Statistical aspects of turbulent flow are reviewed by Frenkiel (1953), Hinze (1959), and Pasquill (1962). Basically, turbulent flow cannot be described by a mean velocity, only,since the velocity fluctuationsaround the mean velocity are an indication of the intensity of turbu- lence. Usually, the instantaneous velocity is represented by: u = E + u' (2'12) 21 where u' represents the turbulent velocity or instantaneous difference from the time mean velocity J and such that ET = 0. However, since the turbulent velocity changes continuously with time, Dryden and Kuethe (1930) suggested the use of the root mean square value /:?§_ . The intensity of turbulence, degree of turbulence or turbulence level is (13—?! >16 then defined as for the longitudinal direction and _ u L W”)2 ————— would be the transverse intensity of turbulence. Turbulent transport processes may be statistically described in two ways according to Hinze (1959). The Eulerian description involves the variation of some pro- perty with respect to a fixed coordinate system. The variation of the property connected with a given fluid particle or fluid lump while moving through the flow field is the Lagrangian description. Frenkiel (1953) defines the Eulerian longintudinal correlation coefficient for velocity as: RX(x) = i Q (2.13) we ““th The value of this correlation coefficient will range from one when the points (P and Q) coincide to zero when the points are far apart. A similar correlation coefficient can be calculated in the transverse direction. The Eulerian time correlation coefficient would be defined as: 22 u'(t) u'(t+h) R (h) = p . P ‘3 /u"p"'" (It)? /u£) (t+h) 2 (2.14) where h represents the time lag between instantaneous velocities at the same point. Frenkiel (1953) defines the Lagrangian correlation coefficient in the longitudinal direction as: u uA(t7 uA(t+h) RtL(h) = (2-15) /uA(t)2 /uA(t+h)2 This correlation coefficient corresponds to a Lagrangian time—scale of turbulence: u w u : LtL = f RtL (a) d a (2.16) O A similar Lagrangian length-scale of turbulence can be calculated based on the distance traveled during time (h). According to Taylor (1935) this length is analogous to Prandtl‘s mixing length. The corresponding Eulerian length-scale is an indication of the average eddy size. Additional description of turbulent flow is obtained by determination of the spectrum which measures the relative contribution of various frequencies of velocity fluctuations to turbulent energy. Frenkiel (1953) defines the longitudinal Spectrum of turbulence as: Fx(k') = elm 577' /{ Rx(s) cos (k's) ds (2.17) o 23 where the functiOn [Fx(k')] represents the contribution to 577 at the wave number k'. The dispersion of a fluid element or particle in turbulent flow is usually iescribed in terms of the variance of the coordinate system components (f7, if; or 2?). Assuming homogenious isotrOpic turbulence, Frenkiel (1953) derived the fundamental equation of turbulent diffusion: ‘7' ‘2' t y = 2 v of (t-a) Rh (a) d a (2.18) (a). where a is equal to time lag used for calculating Rh When dispersion time is large compared to the Lagrangian time-scale of turbulence (equation 2.16) the correlation coefficient becomes very small and equation 2.18 becomes: 37’2" ~ 2 \77 Lht (2.19) If dispersion time is small compared to the Lagrangian time-scale (Lh)’ Frenkiel (1953) has derived the expression: ~ 2 yr: [1-1.2: :‘Vztz (2.20) h or when dispersion time is small compared to the microscale of turbulence Ah: 577 = \72' t2 (2.21) For the case when dispersion time neither large nor small compared the Lagrangian time-scale, Frenkiel (1953) introduces a dispersion factor: 24 _ 1 -— Further description of the turbulent flow involves representing the correlation curve by known functions. If vt = v7'Lh and dispersion time is large, equation 2.19 becomes: §7 = 2vtt (2.23) and the eddy diffusion (Dt) becomes: D = v + v (2.24) when the molecular diffusivity coefficient is assumed equal to the kinematic viscosity. Frenkiel (1953) defines a factor of turbulent diffusion: _T n = 2 d(dt) = ‘77 Uj/t Rh (a) d a (2.25) By replacing the constant eddy diffusivity in the diffusion equation with the above factor (n), the equation becomes valid for a large number of dispersion times (Pasquill, 1962). Since Fick's diffusion equation is not valid even in homogenious isotropic turbulence, experimental eddy diffusivities obtained in this manner can only represent apparent coefficients for conditions studied. Frenkiel (1953) explains that the ratio of the apparent coefficient to the real eddy diffusivity is related to the statistical properties of the turbulence: 25 U 51) Inf ,4 N - ° ‘ (2.26) U C N ._3 The ratio is then a function of relative dispersion time (T = t/Lh) and shape of the Lagrangian correlation curve. 2.3c Transport in Mixing Regions Typical examples of free turbulent flow and the corresponding mixing which occurs are reviewed by Schlichting (1960). A free jet boundary occurs between two streams which are moving at different speeds in the same direction. A free jet occurs when a fluid is discharged from a nozzle or orifice. The turbulent region behind a solid body moving through a fluid or a solid body in a stream of air is called a wake. These types of free turbulent flow can be described by boundary layer equations which have been solved for various sets of conditions by Schlichting (1960) and Pai (1954). A basic requirement for the description and solution of turbulent mixing problems is knowledge of the velocity distribution. Goldstein (1930) provided a detailed solution to the boundary layer equations for two-dimensional steady motion. Howarth (1934) and Tomotika (1938) used the vorticity transfer theory to describe the velocity distri— bution in plane and axially symmetrical jets. Results revealed an identical distribution when compared to momentum transfer theory, but the authors concluded that experimental temperature distributions would be required to test both 26 theories. Kuethe (1935) assumed that Prandtl's mixing length is proportional to the breadth of the turbulent mixing region and obtained solutions for velocity fields for mixing of two parallel streams of different velocities and for an axially symmetrical jet. Albertson, et_al. (1948) used three assumptions to solve for the flow pattern in a submerged jet. The assumptions were: (a) the pressure is hydrostatically distributed throughout the flow; (b) the diffusion process is dynamically similar under all conditions; and (c) the longitudinal component of velocity within the diffusion region varies according to the normal probability function at each cross section. Experimental results indicated validity of the assumptions. Pai (1949) solved the equation of motion for the mixing region of a two—dimensional jet and obtained a solution containing the Gaussian error function. Lock (1951) obtained solutions for the velocity distribution in the laminar boundary layer between two parallel streams which differ in density and viscosity. Results indicated that the solutions depend on the ratio of the velocities of the two streams and the product of the viscosity and density ratios. Torda, et_al. (1953) used the von Karman integral concept to analyze the turbulent incompressible symmetric mixing of two parallel streams. The velocity distribution in the mixing region and the thickness of the region was evaluated while accounting for the influence of the upstream boundary layers. 27 Pai (1955) solved the equations for two—dimensional and axially symmetrical turbulent jet mixing of two gases at constant temperature assuming constant exchange coefficients in the mixing region. A second essential requirement for study and descrip— tion of turbulent transport of quantities other than momentum is the concentration distribution. The concentra- tion distribution for a circular jet with annular coaxial stream was measured by Forstall and Shapiro (1950). The results indicated that concentration diffuses more rapidly than momentum. Pai (1954) compared theoretical equations with experimental results and indicated that for mixing regions far downstream, the concentration profiles can be represented by error functions. Pai (1956) presented solutions to laminar jet mixing problems for velocity, temperature, and concentration distributions. In general, all solutions contain some form of the error function. Batchelor (1957) discusses the statistical characteristics of diffusion in jets, wakes, and mixing layers. The author's hypothesis is that the velocity of a fluid particle in free turbulent shear flow exhibits a corresponding Lagrangian similarity and can be transformed to a stationary random function” Csanady (1963) solved a differential equation derived from the energy balance of the mixing layer and obtained solutions for turbulent intensity profiles which agreed with experiments. 28 Several investigations have dealt specifically with velocity and concentration distributions in mixing regions created by wakes. Goldstein (1933) presented the calcula- tions of the velocity distribution in the wake behind an infinitely thin plate parallel to a fluid stream. Holling- dale (1940) and Townsend (1949) have presented theoretical and experimental results which describe velocity distribu— tions and transport in the wake mixing regions. The validity of the mixing length theory for turbulent shear flow has been questioned. Experimental results, reviewed by Batchelor (1950) presents general mechanisms for transfer of momentum, turbulent energy and heat. However, analytical theory corresponding to the experimental results has not been formulated. Cheng and Kovitz (1958) present solutions to the initial value problem involving mixing and chemical reaction in a laminar wake of a flat plate. Coles (1956) proposes the use of universal flows to describe the mean velocity profile of two—dimensional incompressible turbulent boundary layer flows. The wake model for free-streamline flow is used by Wu (1962) to treat the two-dimension flow past an obstacle with wake or cavity formation. 3. THEORETICAL CONSIDERATIONS The transport of air—borne bacteria through an Opening between two rooms with different concentrations probably involves many mechanisms acting individually or simultaneously. For purposes of this investigation, a theoretical mixing region model will be proposed. The model consists of two different uniform concentrations of air—borne bacteria separated by a partition. The two-dimensional mixing region is located at an Opening in the partition through which transport of the aerosol occurs. The opening is considered infinitely long in the direction perpendicular to air flow past both sides. The mechanism of transport considered is the turbulent dispersion of the high concentration aerosol into the low concentration air and the subsequent entrainment on the low concentration side of the model. The analysis of the model involves consideration of several factors: a. Equal air flow on both sides of the mode1—-this analysis will involve determination of disper- sion in highly turbulent air moving at low flow rates typical of ventilation systems. Depending on the intensity of turbulence in the free stream, the influence of turbulence created by a wake of the partition may need to be considered. 29 30 b. Unequal air flow across the mixing region-~a complete analysis of this case involves several considerations. Due to unequal flow, the possibilities of momentum transport through the mixing region and variation of turbulence across the mixing region must be taken into account. c. Variation in Opening size-—the height of opening in the same direction as the air flow will have a direct influence on the extent of dispersion. In addition, any influence of opening size on turbulence characteristics should be established. d. Width of partition-—the width of the partition at the initial point of mixing will require at least two considerations. One is the possibility of increased turbulence due to the wake. The other is a reduction in dispersion length due to the increased width at the initial point of mixing. e. Temperature gradient-—the primary considerations involving temperature gradient are the influence on viability of the aerosol and the possibility of increased convective currents. In order to allow an analysis which lends itself to mathematical ease and clarity, several assumptions are required: 31 a. The particle size of the bacterial aerosol is uniform, ige., the influence of transport mecha-' nisms and gravitational forces is the same for all particles. b. The die—away characteristics of all particles in the aerosol are uniform. c. The influence of a mean pressure gradient existing at the opening is negligible. d. Transport due to molecular effects is very small compared to turbulent transport, i.e., Brownian motion is negligible. e. Turbulence in the mixing region is isotropic, i.e., G77'= VT? = W77. f. The differences in temperature encountered are sufficiently small to allow the use of constant fluid properties. The first two assumptions are based on work conducted by Kethley, g§_al. (1956) with experimentally generated Serratia marcescens. These assumptions would rarely apply under actual conditions, but they are necessary simplifying assumptions which can be met experimentally. The third, fourth and fifth assumptions depend primarily on the inten— sity of turbulence which exists in the mixing region. The model specifies sufficient turbulence to maintain uniform aerosol concentrations on both sides of the partition; therefore, the indicated assumptions would appear to be 32 good. Since a 20°F. temperature gradient will produce less 'than fivecper cent change in the prOperties of air, the last assumption can be used without major concern. 3.1 Basic Diffusion Equations The two-dimensional mixing region described in the model is not unlike the laminar or turbulent jet boundary described frequently in fluid mechanics literature. In most cases, the mixing region formed by the two parallel streams is treated as a boundary layer and the same simpli— fying assumptions to the basic equations are adopted (Schlichting, 1960). Using this approach the transport in the mixing region model can be described by the following steady—flow equations: a. Equation of motion: —-afi —- 35 ._ 323 u 3; + v _§ - V By: (3.1) b. Equation of turbulent energy: — aui2 -— au'2 _ au'z uax +VW‘VIW (33) c. Equation of diffusion: — 3C — ac _ 326 u 3X +V ~53,- -Da§—)7Y (3-3) d. Equation of energy: — ae — 58 _ 329 u-a—X' + V W - at??? (30“) e. Equation of continuity: 33 av _ .32. +53; -0 (3-5) 33 These equations should completely describe the turbulent transport of aerosol particles in the model proposed. The solution to the equation of motion will describe changes in mean velocity across the mixing region when unequal air flows are analyzed. The very low magnitudes of the mean air velocity and small differences in mean air velocity between the two sides should allow treatment of the mixing region as a laminar jet boundary and the use of known fluid properties for air. The prOposed model requires sufficient turbulence to maintain homogeneous concentrations on both sides of the mixing region. This turbulence must be superimposed on the low mean velocities which exist, and therefore moves slowly past the mixing region. The proposed equation of turbulent energy describes the distribution of the statistical turbulent energy factor (J77). This distribu- tion becomes particularly important in cases where a gradient of the turbulent energy exists across the mixing region model. The equations of diffusion and energy are prOposed in order to describe distributions of concentration and temperature in the model mixing region. Because the pro- posed model considers only transport due to turbulent effects, the constants, Da and a will be referred to as t apparent eddy diffusivity and turbulent thermal diffusivity, 34 respectively. The continuity equation specifies that con— servation of mass exists in the mixing region. The purpose of the proposed mixing region model is to provide some theoretical basis for the transport of a bacterial aerosol through an opening in a partition separating two different concentrations. To provide mathe— matical ease in the description of this transport, it would be desirable to Obtain closed solutions to the equa- tions which represent the distribution of the various parameters in the mixing region. One approach to attaining this objective is to make additional assumptions with respect to flow conditions in the mixing region. The assumption that V = 0 should be valid based on specifi— cations of the model. As indicated, the overall movement of air past the Opening is in the downward (+x) direction. All other components of air movement are turbulent fluctua— tions. Therefore, the only non-negative mean velocity is u (x—component). The assumption that v = 0 simplifies the previous equations considerably and thus they may be stated in the following manner: a. Equation of motion: —aJ _ ‘325 1.1—; - \z W (3.6) b. Equation of turbulent energy: "311,2 5213—17- ax t ayz (3.7) 35 0. Equation of diffusion: - 19 = 320 u 8x Da 5;? (3.8) d. Equation of energy: —-39 _ 329 u 53? ‘ “t W (39) The equations of interest are now reduced to more simple forms similar to the heat conduction equation. Such equations have been used frequently by Pai (1949, 1955, 1956) to obtain solutions which describe velocity distri— butions in laminar and turbulent jet mixing regions. More firm support for the use of the preceding equations in regions of free turbulence was provided by Reichardt (1941, 1944). Reichardt's inductive theory of turbulence is derived almost completely from experimental evidence, which indicates that velocity profiles in free turbulent flows can be approximated very successfully by the Gaussian error function. Reichardt's fundamental equation for describing the velocity distribution in free turbulent flow is: au“z 32.32 ( . 3X _ A(X) ayz .3-10) where A is a momentum transfer length. The similarity of Reichardt's equation and equations preposed for use in this investigation is apparent. Solutions to equations 3.6, 3.7, 3.8, and 3.9 can be obtained by use of boundary conditions which describe ,1 3b the model mixing region. These boundary conditions can be stated as follows: At x = 0, y > 0: — = —— ‘77 = ‘T? c = c = u uL, u uL , L e 6L (3.11) x > 0, y —+ 0o: — = __ —77 = —TT = = u uL, u uL , C CL, 6 6L (3.12) x > 0, y —+ —w: — = —— _TT = —TT = . = u uH, u uH , C QH’ e 6H (3.13) “'— 12 '2 u —u u -u —=— H L 72:72" H L C= u u. + 2 , u uL + 2 , %, C -C 6 -6 + H L , e: 8L+ H2L (3.114) These boundary conditions lead to the following solutions (See Appendix A.2) to the proposed equations: a. Equation of motion: = % 1 - erf % uL (3.15) VX $1 cl L—1 J U_ H b. Equation of turbulent energy: uv2_u12 -' fif- ~ _ 1 .y / L H L c. Equation of diffusion: C - C L I ———————— = 1 — erf L (3.17) CH ’ CL 2 Dax [UH-J 37 d. Equation of energy: a -e /u— L y L 317;“ ‘ l ’ “TC at x (3‘18) I\)|I—‘ H L The above solutions contain the error function which is typical of solutions to the heat conduction equation. Several workers (Reichardt, 1944; Albertson, 1948, Liepmann and Laufer, 1947; Corrsin, 1943; Hinze, §t_al,, 1948; Schlichting, 1960 and Forstall and Shapiro, 1950) have presented experimental results which confirm that velocity, temperature, and concentration distributions in jet boun— daries and wakes can be represented by solutions containing error functions. 3.2 Turbulent Diffusion A complete description of the transport in the model mixing region depends on an accurate determination of the air flow conditions. Since turbulence is the primary mechanism of concern, the statistical description introduced by Taylor (1921) offers the most accurate approach. The following derivation of the fundamental equation of turbulent diffusion should apply to the model mixing region proposed in this investigation. For transport of the bacterial aerosol to occur in the model mixing region, the particles must move in a direction perpendicular to the opening (y-direction). The y-location of a given particle after a dispersion time of t would be: 38 t y = / V'(E) d a (3.19) o where 5 represents the time lag between instantaneous velocity measurements. By computing the variance of the y-component of particle location: t 2 t I7= / V'(€) d 5:] = [ V'(t) V‘(€) d5 (3.20) O O ' and consideration of the auto correlation coefficient: _ V'(tI V'Ig) Rug) - v (3.21) then: t t / v'(t7 v'n) d5 =57? / Rd.) at (3.22) O 0 By integration of the left side of equation 3.22 ET??? // NTTEI d6 = v' t) y = VT? /[ R(£) 95 o o (3.23) By restating: ‘T“a l d [‘E] - '77" /[t R( ) d ( 24) V y 2 a? y ' V O E.» E 3. then: T t 'y‘T= 2 W / / Rm dadt (3.25) o o which is the fundamental equation of turbulent diffusion first derived by Taylor (1921). This expression provides an accurate and direct method of determining the statistical 39 dispersion of the bacterial aerosol through the Opening into the low concentration air. This information is attainable by determination of a turbulent energy factor (577) and auto— correlation coefficients, both of which can be obtained by measurement of instantaneous velocity fluctuations. 3.3 Turbulent Transport Coefficient The transport of the bacterial aerosol through the experimental opening will be presented as a turbulent transfer coefficient. This coefficient is based on Fick‘s law of diffusion (Treybal, 1955): A a O (3.26) I II I U I 3> O) y which describes the movement of some component due to a concentration gradient. The mass flux (NA) represents the rate of movement and D is the proportionality constant or diffusion coefficient. When describing turbulent diffusion, such as in this investigation, the equation is written as: A :_D 8—C. A t 3y (3-27) where Dt is the eddy diffusivity. 40 The mixing region model and the corresponding trans- port of aerosol through an opening perpendicular to the concentration gradient, the variables of equation 3.27 can be separated in the following manner: NA b CL A‘ /f dy = —Dt ,/r dC (3-28) 0 c H where b is equal to the width of the mixing region. By integration, equation 3.28 becomes: NA D 77 = E} (cH - CL) (3.29) with CH and CL being equal to the high and low concentra- tions, respectively, the ratio (Dt/b) can then be replaced by the transfer coefficient (kc) in the following manner: NA = kc A(CH — CL) (3.30) This equation describes the turbulent transport coefficient to be used. One important factor must be recognized. Because of the entrainment mechanism involved in the model, this coefficient (kc) will be proportional to the mixing region width and will not decrease as suggested in equation 3.29. However, the relationship of the coefficient to the mass flux (NA)’ area (A) and concentration gradient (CH — CL) remains. Therefore, the turbulent transfer coefficient (kc) is excellent parameter to be measured experimentally in this investigation. 41 3.4 Eddy Diffusivity Transport in the model mixing region occurs only due to turbulence as specified by basic assumptions in the description of the model. As indicated by equation 3.27, turbulent diffusion is described in terms of an eddy diffusivity Dt which is analogous to the diffusion coef— ficient in molecular diffusion. However, a basic differ- ence does exist in that the molecular diffusivity is characteristic of the fluid in which diffusion occurs, whereas the eddy diffusivity is related more closely to the flow cOnditions which exist. Because of this differ- ence, it will be necessary to determine an eddy diffusivity which corresponds to each set of flow conditions investi— gated. The most desirable method for determining eddy diffusivities in the mixing region model would be by measurement of concentrations at points in the mixing region and use of equation 3.17. However, this approach, is time consuming and would not provide the accuracy desired because of difficulties encountered in making point concentration measurements of bacterial aerosols. The statistical approach proposed by Taylor (1935) offers an alternative method. From the definition of the auto- correlation coefficient in equation 3.21, it is evident that the coefficient is unity at E = 0 and zero at E —+ w Because of this relationship, it is possible to define some time (T) beyond which R(g) = 0. Then, it is possible to state equation 3.23 as: ___ t yV' = V'2 // R (5) d5 (3.31) o where yv' is constant for t > T even though y? is increasing continuously. For these conditions, the constant (yv') is defined as the eddy diffusivity: t Dt = 677' /f R<£> dé (3.32) o which is constant for a given turbulent energy factor (577) and integral of the autocorrelation coefficient. In addition, Taylor (1935) defines a ”length scale of turbulence" as: 1 't ‘3«(({7_—'_7)/2 = V77. // 3(5) d6 (3.33) O 01" 1’ t - 2, = (777)? I R(g) d5; (33“) O This length scale is assumed to have the same relation to turbulent diffusion as the mean free path does in molecular diffusion. This value may provide considerable information on the mechanism of transfer for the mixing region model of this investigation. 3.5 Transport in the Mixing Region For purposes of this investigation, transport of the bacterial aerosol from the high concentration region 43 to the low concentration region will occur due to disper- sion in the mixing region and entrainment on the low concentration side. Equation 3.25 describes the extent of dispersion in the direction perpendicular to the Opening and (577)15 should then represent the width of the mixing region on the low concentration side. The entrain- ment of the aerosol would occur due to the mean flow of air along the low concentration side. If the dispersion time (t) is large compared to the time (T) required for R(£)—+ 0, then equation 3.25 can be stated as: T t ,/ I?" = 2V” of 13(5) d5 / dt = 2(FTV 2.1: (3.35) O For turbulence sufficient to maintain uniform aerosol concentrations, such as specified for in the mixing region model, dispersion time (t) should be large compared to the time (T). For each set of flow conditions, the apparent mixing region width can be defined as: t L t b = (375V = 2HmmmeO amumcmte mmeMm cowmwwwum wwmwwmw “WNW sou anm nHa ovum loam ha< .muHamma no unmassm11.na.m mqm.o moo.m mmmm.o m m.mm KNOH om OOH.O OO.O OO.H OOm.O OOO.H Oem.H ONOO.O O e.Om m.Hm Om aOm.O HN.OH OO.O OH:.O mmO.m OON.H ONOO.O O ON om Om OHN.O me.O OO.O Omm.O OOa.m NOO.O ONOO.O O OH O: OO Hem.O OO.ON OO.H OOO.H OOe.m OOO.H ONOO.O O OH- OO Om mmm.O Ha.OH OO.O HOm.O HmO.m OOO.H OOOO.O O O om Om OOO.O OH.OH OO.O OOO.O OOH.a mmm.H ONOO.O O O O: OO MOO.O Oa.am OO.O HOm.H Oem.a OHO.H OmOm.O O O OO OO OO.O OO.O OO.H mem.O OOO.O mmm.m OmHm.O O m.Om e.OH Om mmO O0.0 OO.H OOOO 23m HOOH OmHmO O TON m.Hm Om. Os.O Hm.OH OO.H OHa.O msH.m OOO.H OmHm.O O O OH OO OO.O me.O OO.H Omm.O OAO.: mOH.H OmHm.O O OH on .O 2O.O OO.OO OO.H OOO.H an.a OHO.H OmHm.O O OH- OO Om O0.0 HOOH OO.H HOMO O0.0; mOmH mmHmO O O om Om HO.O OH.mH OO.H OOO.O OO0.0 mm .H OmHm.O O O on O: em.H OO.:N OO.H HOO.H OOO.O OOO.O OmHm.O O O OO Om HH-OHxv AeHe\O.eHO .emwmwmmcwmwa H.eHO H.eHO H.eHO HeHe\maOO HeHe\mauc HeHe\maOO OO~.: x~.: -mmi so seeHOHOOOOO O.OOOH3 on : .OOHO .OOOO .OOOO. ..... s. was a...” nH< oumm 30am hfi< .muHsmoa no >pm8€smti.oa.m mqm¢a 83 OO.O NO.O NOH.H NO O.eO O.NH O0.0 OO.N N:H.H NO 3O NH OH.O :0.0 mNH.H OO OO OH ON.O mm.O MOO.H Os mm m :m.: NN.O O.H me me O HO.m OH.: OOO.O mm OH O - OO.N OO.N OO0.0 0.00 0.00 OH- Om.N NH.N Osm.O HO Om HH- HO.H Om.N ONO.O NO m.OO NH- :s.H OO.N OO0.0 m.OO O.aO aH- OHNOH\.Ozv CHEINUM\.OZQ A.h0v A.mov A.&ov _ O OH Om pupa .aceHem Oquom . . m . m OOOHOOMO o mmooo O manpmpo Ems magpmam EOE magpmpm Ema D x pomwcmhe oaumm Cowumpucoocoo coapmppcoOcoo 20H eme .mpHsnet so semesOm-.OH.O mHmae 84 be somewhat high. However, it is assumed that the area produced by lack of damping is small and, in addition, would be relatively consistent in all calculations. The use of equations 3.25: T t 377 = 2 F? / / 11(5) d5 dt (3.25) O O b — (Ff/2 - [2(Wfii-tlg (3°36) with VT? replaced by ET? (assumption e. in section 3), allows calculation of the statistical dispersion and an apparent mixing region width (b). As is evident from equation 3.36, the mixing region width is directly depen- dent on the dispersion time (t). The selection of a proper dispersion time for the aerosol moving from the high to the low concentration compartment of the aerosol chamber was not easy. The most likely selection appeared to be the equilibrium time discussed earlier in section 5.1. Since the dispersion did not start until the aerosol and filtered air came in contact at the initial point of mixing, aerosol moving from one side to the other was not exposed for the entire equilibration time. Calculations are based on dispersion times which are a fraction of the equilibration time. The fraction is based on the location of the initial point of mixing with respect to the overall height of the aerosol chamber. Using dispersion times (t) and equation 85 3.36, the apparent mixing region widths in Table 5.1 were calculated. The influence of air flow rate on the apparent mixing region width is illustrated in Figure 5.10. The results indicate that the width (b) increases with increas- ing flow rate. This relationship exists even though the dispersion time (t) increases significantly with decreasing flow rate. In addition to statistical dispersion and apparent mixing region width, the autocorrelation coefficient is required in the calculation of the specific eddy diffusi— vity (Dt) according to equation 3.32: t D1: = 1770/ R(€) d: (3.32) These values have been calculated for the required air flow situations and are presented in Table 5.1. 5.3 Concentration Distributions In order to verify equation 3.17, which indicates that the concentration distribution in the mixing region should be described by some form of the error function, point concentration measurements were conducted in a limited number of situations. Measurements Obtained permitted calculation of a single eddy diffusivity value at each point in the concentration profile. The values varied considerably; however, in order to satisfy equation 3.17, the apparent eddy diffusivity (Da) must be constant. A mean of the values obtained in the concentration profile 86 was used to present the data in Figure 5.11. Data obtained with flow conditions of 50 ft3/min in the high concentra- tion compartment and 50 ft3/min and 30 ft3/min in the low concentration compartment give good agreement with the proposed function. The obvious weakness to the above method for deter- mining eddy diffusivity values is lack of experimental accuracy for obtaining point concentration data. In addition, the function requires that the eddy diffusivity be constant in the mixing region which may or may not be true: Use of equation 3.22 allows calculation of specific eddy diffusivities (Dt) at points in the profile. There— fore, the values obtained by point concentration measure- 'ments will be referred to as apparent eddy diffusivities ma). 5.4 Transport of the Bacterial Aerosol The complete description of the transport of a bacterial aerosol will combine two phases of study: (a) experimental determination of turbulent transfer coeffi— cients (kc) defined by equation 3.30 and (b) use of the dimensionless transport relationship presented in equation 3.42, 5.4a Experimental Transfer Coefficients The influence of air flow rate on the experimental turbulent transfer coefficients (kc) is illustrated in Elia ‘1 87 .Cowmoa mCHxHE CH soapsofihpmfip COHpmapcoocOOIl.HH.m masmflm E .H .8555 3858:. O. O O O N O N- T O- O- O H.O a. __._____._ o o - a [NO 85:52.8 0 £2: 13 V .. IOO 1% 8.8-6 3 tel 15.. OO-OO-o a Q a o 552.8 .2... .2 a c 01103 uoiiaiiuaouog 88 Figure 5.12. The experimental data produced a coefficient of variation of i0.l76 for kC from the curve selected according to procedures presented in section 4.4. As is evident, the coefficient (kc) increases significantly between 20 and 40 ft3/min, but becomes relatively constant above 40 ft3/min. Results (Figure 5.13) reveal the influence of air flow rate gradients on the turbulent transfer coefficient. For aerosol flow rates of 40 and 50 ft3/min, the coeffic— ient was maximum at equal flow rates. If the aerosol flow rate was 30 ft3/min, the maximum transfer coefficient occurred between air flow gradients of -15 and —20 ft3/min. Figure 5.14 presents the relationship between a temperature gradient across the mixing region and the transfer coefficient (kc). The correlation appears to be linear for the range of gradients investigated with a standard error of estimate (SEy) of :(0.651 for experimental coefficients. The best fit curve for the experimental points can be described by the following equation: kc = 0.2195 119 + 5.09 (5.2) Since the influence of a temperature gradient on transport of the bacterial aerosol in the mixing region could be related to several factors, an attempt was made to isolate a portion of the influence. Several investigators (Kethley, gt_al,, 1956; DeOme, gt_§l., 1944; Webb, 1959; Hayakaw and Poon, 1965) have determined and 1.1 89 l Turbulent Transfer Coefficient, kc b l , 1 1 1 1 1 I0 20 so , 40 so Air Flow Rate, f., f'/rnin Figure 5.12.-~Variation of transfer coefficient with air flow rate. \ Turbulent Transfer Coefficient, ltc '1’ Aerosol Fla-l rate ft‘/min. 30 -—-—— 4O —— \ P so ----- 0 J l I l l l l l J -40 -20 o , 20 40 Air Flori Rate Difference, Af, n/min Figure 5.13.--Influence of air flow rate gradient on transfer coefficient. 60 i- L—o . 9O discussed the influence of temperature and relative humidity on viability of bacterial aerosols. Since the temperature gradients maintained in the aerosol chamber were developed by heating the air, both factors could contribute to the relationship illustrated in Figure 5.14 Death rate constants for aerosols of Serratia marcescens were determined by Kethley, et a1. (1956) for various temperatures and relative humidities. These data indicate the drastic influence of relative humidity along with the influence of temperature. However, experi— mental data were available only up to 80°F. while data of this investigation were obtained at temperatures as high as 98.5°F., therefore, extrapolation was required. In order to extrapolate the available death rate informa- tion as accurately as possible, the data were presented on semi-logrithmic coordinates as shown in Figure 5.15. Use of the loglccversus 1/6a relationship implies that the death of air-borne bacteria obeys a first-order kinetics reaction. This type of relationship has been used successfully by Webb (1959) and Hayakaw and Poon (1965). Results in Figure 5.15 indicate two stages of death, at least at lower relative humidities. The influence of temperature on death rate is small from 40° to about 70°F., but becomes very significant as the temperature increases above this level. The death rate Turbulent Transfer Cull-dent, t. Figure 5.14.—-Inf1uence IIYVII] I /.. “I I M! Role Constant, 91 1,- aizsmov 5.0910615 Steward Error of [oi-note t 20015 for I, l L .14 -m -s -2 o 2 6 to lo Ten-perature Grad-ant. be." of temperature gradient on transfer coefficient. - Relative Nullllly,°/9 0 - 00 A - 70 D - 60 V -50 1 -4o .1 l 1 l 1 l l l 19’) 19 I65 18 inverse Aowiule Temperature, I/9.°R Figure 5.15.--Influence of temperature and relative humidity on death rate of Serratia marcescens» Turbulent Transfer Coefficient, I, [I'fliOIBfol -—— Account-M lav «Iii-tr chance I l l 1 I 1 1 1 l 1 44 -IO ~‘ '2 3 2 annual-in Gradient, 119. °F 1 l L 41__j ‘ lo I. Figure 5.16.——Influence of aerosol viability on the relation- ship between temperature gradient and transfer coefficient. .4“ constants used to account for viability changes due to the temperature gradient were obtained from Figure 5.15. By adjusting the experimental turbulent transport coefficients to account for changes in viability during transport, the data illustrated in Figure 5.16 were obtained. The results reveal an increase in the coeffic— ients (kc) at negative temperature gradients and a decrease at positive temperature gradients. However, it is evident that viability does not account for the entire influence of the temperature gradient and other factors must be contributing to the apparent transport. 5.4b Dimensionless Relationships In order to describe the transport of bacterial aerosols due to turbulent diffusion, dimensionless trans— port groups derived in section 3.5 will be used. The influence of each of the five basic parameters will be presented and-discussed separately and then all data except that obtained for a temperature gradient will be used to establish an eXpression to describe the influence of all parameters. The influence of dispersion time (t) and mean air ke Dt / f(B) velocity (u) on dimensionless transport u'zb bo O is illustrated in Figure 5.17. The results indicate an increase in transport with an increase in the dimension- less value (£,/Et) and reveals that the influence of the 93 .zpfiooao> saw some cam oEHp coampoamfie Noocoasnpzp mp econo3Hmcfl mm pLOOmcmLp mmOHCOHmcoEHQII.OH.m oaswflm E Oo_ x 4%- OO ON OO OO OO OH ON 2 OO _ H _ _ _ H _ _ \ 1 _ O \\ \\ 58.23252 3 as... 2225 1 . ONOOO e \\ NO ONsO .N. \ m \ ONOO.O - o \ \ \ :5 52; see: \ IHO \ \ \ lie T5588 - .5 we .N «D 01 13 1 NO \ . \ 1 OO \ \ \ \ ssaiuoisuauiig lJOdSUDll °q all 1 10 9)‘ J. / °q (€11 94 length scale of turbulence (2,) dispersion time (t) over- shadows that of the mean velocity (3). The relationship is approximated very well by the equation: kCDt = 0.00796 fl 5451 — 0.0139 ETTbO at bo for the three partition widths at the initial point of mixing. The standard error of estimate (SEy) is t0.0868 for the dimensionless transport group. The influence of gradients in air flow between the two compartments of the aerosol chamber on the dimenSion— k D less transport group —9——3 is presented in Figures 5.18, ET7bO 5.19 and 5.20. For all three situations (50, 40 and 30 ft3/min through the high concentration compartment), the transport decreases with an increasing positive air flow gradient. However, the influence of a negative gradient is not consistent. For f0 = 50 ft3/min (Figure 5.18), the transport decreases with increasing negative air flow gradient in about the same manner as indicated for, a positive gradient. The transport also decreases for small negative air flow gradients with f0 = 40 ft3/min (Figure 5.19), but becomes relatively constant at a higher level than the positive gradient. When fO was decreased to 30 ft3/min, the transport increased with increasing negative air flow gradient before decreasing to a level nearly equal to that at equal air flow rates. 95 The transport characteristics illustrated in Figure 5.18, 5.19 and 5.20 are probably more closely related to turbu- lence in the mixing region than to the mangitude of the air flow gradient. This is clearly evident from Figure 5.20 where the dimensionless transport was low with equal flow rates of 30 ft3/min. By increasing the flow in the low concentration compartment, the turbulence and transport ‘both increased. This corresponding increase in both para- meters continued until momentum transport from the low to the high concentration compartment was sufficient to over- come the increased transport due to turbulence. This same sequence of events did not occur at high air flow rates since turbulence levels were sufficiently high to produce nearly maximum transfer at equal flow rates. An increase in negative gradient did not increase turbulence enough to overcome the influence of increased momentum transfer. The increase in positive air flow gradient does not reflect an apparent influence of momentum trans- fer, indicating that the decreases resulted in signifi— cant decreases in turbulence. Results (Figure 5.21) reveal the influence of parti- tion width and apparent mixing region width on the dimen- sionless transport group. Several factors are illustrated by the results, however, most evident is the increase in transport with increasing partition width followed by a decrease to a level nearly equal to the transport with 96 071— ad - \ Au no. Rm = sori’j-u-r- Pas-tive Gradient 0.5 - \ — — — Newt-n Grad-ant g 0.4 1— g .z .2 0.3 '— 5 0.2 - 0|— C) 0 0.! 0.2 0,3 0 4 0.5 06 0.7 0.8 111 la Figure 5.18.-—Variation in dimensionless transport with air flow rate gradient for aerosol flow rate of 50 ft3/min. as s 1,: 101- /---i-- _— Positive Gradient — —- Negative Gradient 0 5 5|; .1 I! . l l l l l l l J 0 0| 02 0.3 0,4 0 5 0 6 0.7 0.8 .NIE Figure 5.19.—-Variation in dimensionless transport with air flow rate gradient for aerosol flow rate of 40 ft3/min. 7 r—- P 61— ,. Air ria- Rate of 3011’In-n liar-um Grad-em ‘5 3 5._ ———-Neqot-: Gradient :15.- - 1- /’—‘\\ '5 a / \ e , \ r- / \ v- / \ E 3_ / . \ .§ // \ C c 7 \. 6 I Z '- \ I h— . 1 L 1 l 1 l 1 l 1 l 1 1 JJ "0 02 or as as 10 12 14 09/“ Figure 5.20.—~Variation in dimensionless transport with air flow rate gradient for aerosol flow rate of 30 ft3/min. 97 the narrow partition. The increase detected must be due to increased turbulence in the mixing region even though it was not detected by measurements. It seems possible that small increases in turbulence could occur in the, mixing region without detection and cause increased transport. The decreased transport with an even wider partition would be due to a significant decrease in the apparent mixing region width while turbulence may have increased only slightly. The relationship between the dimensionless transport 3:2: group £77 b and dimensionless opening height (H/H') o is presented in Figure 5.22. The decrease in transport per unit area with increasing opening height is apparent in all flow situations, but is most significant at equal flow rates. An explanation of this relationship is probably related to results presented in Figure 5.10 where the apparent mixing region width was shown to be a function of the square root of the distance from the initial point of mixing. Therefore, transport per unit area based on the dispersion-entrainment mechanism would larger for small Openings. The relationship between the dimensionless transport kc Dt group ETYb and dimensionless temperature (eHC/GLC) based O on the temperature gradient is presented in Figure 5.23 98 08r- 1W) ..0 o I / k0- D-n-ens-onless Transport, “5.5. O O a U! T I o 1.,- l o N l Oli- 1-1 $11- 11%.- 0 -50 A -40 -30 Figure 5.21.-—Inf1uence of partition width on dimensionless transport. IZ— D S IO— Air Flo- Rate D-ffcuncollt'lmn) :lA 1- 0 ‘0 \ A -lO 0- 3 .p— U ‘20 .218 v -2er _ . X -393 5 o. 1‘ ts— ' ..: § " 41- 1;: C E 1- 5 2_ \ A 1 1 1 1 4m " 005 01 one 0.2 H lH' Figure 5.22.-—Inf1uence of Opening height on dimensionless ‘cD' 16gb. D-rnensionleu Transport transport. 6!— 5r— 4»— 3.— 11, 0. 2L- 0 Standard Error of Est-note = 201501 for =——— 11"!)° l— ,__,1 l 1 4 1 1 l 1 1 1 1 I 1 L 1 1 1 1 1 1 1 J o ' 03 09 10 11 1.2 0 “/01: Figure 5.23.——Inf1uence of dimensionless temperature ratio on dimensionless transport. 99 This relationship is very similar to the linear relationship presented earlier in Figure 5.14 kc Dt u' b = 16.29 (eHC/eLC ) O — 12.1 (5.4) where 6H0 and 6LC represent temperatures in the high and low concentration compartments respectively. An adequate explanation of the influence of temperature gradient other than the influence on viability is not apparent. A possible explanation is that convective heat transfer and a vapor pressure gradient across the mixing region contri- bute or decrease the transport of bacterial aerosol. The influence of these two factors would be in addition to the turbulent transport normally present. The dimensionless transport equation derived in section 3.5 was: k D u v D c t = K‘ f(B) L t _E (3.“2) u'zbO b0 u'2 H vt where K' represents a constant to be determined from experimental data. Using all data obtained without a temperature gradient, the slope of the straight line approximated by the least squares method was 0.01276. Therefore, K' = 0.01276 and equation 3.42 becomes: kc D 0 D U. t = 0.01276 féB) —£——45 33 u'zbO o u'2 H t + 0.772 (5.5) 100 Additional investigation indicated that data obtained with a temperature gradient could be described reasonably well by the same relationship by using the temperature ratio (GHC/BLC) as an exponent to the dimensionless trans— uL Vt port group ET? H . This provides the following dimen- sionless relationship for data presented in Figure 5.24: )Dt + 0.7u1 k D G' v (9 /0 c t = 0.0136“ f(8) L t HO LC u'zbO b0 u‘2 H Vt (5.6) This relationship describes the transport of bacterial aerosols for all parameters studied in this investigation. The standard error of estimate (SEy)_Of'il.OOU for dimen—~ sionless transport values applied for a range of dimension— less turbulence values from 0 to 0.7. lOl .pmoomcmhp Homopwm Hafihopown CH vm>ao>cfl munchw mmoHGOHmomEHo no COHumHmspoo Hmpcoefipooxmrl.:m.m opswflm o. x F Azfiv on n .0 Aud¢\u=¢v «$4: Amvw com co» com com ooe can oo~ oo- 00 d e _ _ _ _ _ — \o 23:35 22235:. x \ 523 35:0“. D ‘N 222. 9.226 Q B \ \ Essa 2.: .5: .2 a \\a 1 22 .5: .2 0 €\ \ GD 11 ¢ G w_fl. Im M... mu la \\ \\\\ 1 . . I I: on . on «m... _E.o+ $ A thlsv g vmfioo u .o as lo. 182$ - LN. 6. SUMMARY AND CONCLUSIONS 1. By increasing the air flow rate gradient between two turbulent air streams, an increase in momentum transfer was indicated. A significant shift of the turbulent energy profiles to the lower flow rate side was revealed for rates of 30 ft3/min or higher. At flow rates of less than 30 ft3/ min on the low flow rate side, an apparent decrease in momentum transfer resulted due to decreased turbulence in the mixing region. 2. Concentration distributions in the mixing region gave good agreement with the expression: C — C‘ u L l y L 1" = — l - erf CH - 6L 2 2 Da x where Da is an apparent eddy diffusivity which is constant in the mixing region. 3. Transport of bacterial aerosol in a turbulent mixing region increased with increasing equal flow rates. The following relationship obtained from experimental data: kc Dt = 0.00796 :l— f(8) - 0.0139 u'zbO ut bo indicates that transport is a significant function of turbulence and dispersion time and is only slightly influenced by Changes in mean air velocity. 102 103 3/min, transport of A. At air flow rates above 30 ft bacterial aerosol was maximum at equal flow rates through both compartments of the aerosol chamber. At these flow rates, transport decreased with increasing positive or negative air flow rate gradients. 5. The width of partition at the initial point of mixing had only a slight influence on transport at air flow rates of 30 ft3/min or higher. Transport increased at a partition width of 0.3125 in., but decreased as the width was increased to 0.5625 in. The decrease was attributed to a decrease in mixing region width corres— ponding to the increase in partition width. 6. Transport of bacterial aerosol decreased with increasing Opening height. This was related directly to the apparent mixing region width, which was a function of the square root of the distance from the initial point of mixing. 7. A temperature gradient across the mixing region influenced transport of the bacterial aerosol as indicated by the expression: kc D t u'zb o = 16.29 (eHC/eLC) — 12.1 which applies to a range of temperature gradients from —lAO to +12.5°F. 104 8. The tranSport of a bacterial aerosol was described by the following dimenSionless relationship, which considers all parameters studied: * k D , p u- v (6 /6 -) D _E__£ = 0.01364 féB) _£__E_ HC LC —£- + 0.741 u'zbo o W}: “t This relationship applies for dimensionless turbulence valués between 0 and 0.7. 7. RECOMMENDATIONS FOR FUTURE WORK The results of this investigation indicate the need for additional work in the following areas: 1. Investigation of turbulence levels and momentum transfer in mixing regions between laminar and turbulent air streams. Additional investigation of transport mechanisms present in mixing regions formed by an air flow rate gradient. Further investigation of mechanisms related to transport of bacterial aerosol in mixing regions as influenced by partition width at the initial point of mixing. Investigation of bacterial aerosol transport with a temperature gradient across the mixing region. 105 APPENDIX 106 APPENDIX A.l Air Flow Measurement From Eckman (1950), the following equation was used to calculate air flows through both compartments of the aerosol chamber: 2 2 n CVT B D ¢V q = b /M'(\)m — v? 1 ————2gh II /————¢1 _ 8 Mb V. / where: q = air flow rate, ft3/sec CVT = venturi discharge coefficient 8 = diameter ratio = d/D D = pipe diameter, ft. 9 = rational expansion ratio vb = specific volume of gas at base conditions, ft3/lb. Mb = moisture factor at base conditions M, = moisture factor at upstream conditions Vm = weight density of manometer fluid, lb./ft3 0 = density of fluid over manometer fluid, lb./ft3 f g = acceleration due to gravity, ft/sec.2 h = manometer differential, ft. 107 108 Table Arlshows the corresponding air flow rates and pressure drOp values used in this investigation. TABLE Aal——Pressure drop, air flow rate and mean air velocity Mean air _ Pressure Drop Air flow rate velocity (u) (in. H20) (ftj/min) ‘(ft/min) 0.018 10.7 0.667 0.074 21.3 1.333 0.260 40 2.500 0.410 50 3.120 0.550 58 3.625 0.600 60 3.750 0.610 61 3.813 0.800 70 4.375 and A.2 Solution to Diffusior Equation According to equation (3.8): _ . D 32 u 3? — a a 2 J O k 9:: boundary conditions (3.11), (3.12) and (3.14): H C) At x n 0, y > 0: C L X>O,y=oo: C=CT “ C —C H L x > O, y = 0: C = CL + 2 -— (A. (.1 (A l) .2) .3) .4) 109 The solution to equation (A.l) will be of the form: O: f (11‘, Da, By use of linear relationships between variables, it is evident that: ‘ _ E C-AF y/Dax) (A. By substitution of equation (A.6) into equation (A.l): F'Kn) = - % F'(n) (A. where: _. Lil; n - y D x (A. ['8' By solving the differential equation (A.7): n PM) = c = A / epr-(n/2)1dn + B (A. 0 Since: n 2 [ expE—nz] an . (A. H? 0 is equal to the Gaussian error function, equation (A.9) can be expressed as: = 1 /J C A erf <; D x + B (A. a Using the stated boundary conditions: C —C C -C - H L . _ H L A — - 2 , 13- CL + 2 (A So: _ C C _l - L 1 y E = — l — erf ——— (A. CH-CL 2 2 Dax L _. X. y) (A. 5) 6) 7) 8) 9) 10) 11) .12) 13) 110 .mHHmosQ mpfiooflo> Lam smozll.a.< opsmfim E ;.. 13835 $13,?ch w m N o N- s- m- m- . 4. a _ _ _ 4 _ a M . M _ _ a _ o mu 1L. MW U WV. 1 M w EE\:_.m .11“ . n... .52?2. $8 m/H. :_E\n:m._~ u 11m 111 A.3 Mean Air Velocity Profile Using the solution to the equation of motion as presented in equation (3.15): 5| l c l NlI—J lverf 521 VX a predicted mean velocity profile can be obtained. The profile for the air flow conditions of 50 ft3/min in one 3 compartment and 21.3 ft /min in the other compartment of the aerosol chamber is present in Figure A.1. A.4 Experimental Turbulent Transfer Coefficients The data presented in Tables A.2 through A.13 represent individual trials conducted to determine the turbulent transfer coefficients (kc) for various situations. From equation 3.30, it is evident that: NA kc = A(CHVCL (A.14) In the following tables, the steady—state transfer represents counts in air samples collected at the low con— centration compartment outlet. The mass flux (NA) was calculated from the mean of the steady—state values for flow rate in the low concentration compartment. The turbu— lent transfer coefficient was calculated from equation A.14. 112 TABLE A-2.-—Inf1uence of air flow rate on transport charac- teristics of air-borne bacteria with 1.5 ft2 opening. Air Flow Steady-State Mass Concentration Transfer Rate Transfer Flux Gradient Coefficient [No./min-ft2 (ft3/min.) (No./ft3) (No./min.) (No./ft3) (No./ft3)] 61 22-21-25 1382.66 189 4.8771 22-27-24 1484.33 170 5.8209 24-22-25 1443.66 154 6.2496 54-50—44 3009.33 405 4.9536 58 25-24—16 1256.67 157 5.3362 42—34—27 1991.33 250 5.3102 27—19-25 1372.66 221 4.1408 35-31-33 1914.00 269 4.7435 50 54-47-50 2516.66 375 4.4741 70-51-68 3150.00 387 5.4264 36—40-38 1900.00 216 5.8642 54-29-36 1983.00 238 5.5556 40 75-77—62 2853.33 358 5.3135 87-65-67 2920.00 342 5.6920 69-80-65 2853.33 415 4.5837 77-63-79 2920.00 396 4.9158 30 28-38-31 969.99 283 2.2850 37-37-26 999.99 242 2.7548 61—61—58 1800.00 375 3.2000 29—66-70 1650.00 382 2.8796 43-38—42 1230.00 332 2.4699 65-46-57 1680.00 373 3.0027 21.3 21—22—23 468.60 436 0.7165 17-34—21 511.20 441 0.7728 24-24—26 525.40 407 0.8606 21-28-29 553.80 400 0.9230 113 TABLE A-3.—-Inf1uence of air flow rate gradient ogutransport characteristics of air—borne bacteria3with 1.5 ft opening and aerosol flow rate of 30 ft /min. Air Flow Rate Steady—State Mass Concentration Transfer Difference Transfer Flux Gradient Coefficient 3 3 3 [No./min-ft2 (ft /min) (No./ft ) (No./min) (No./ft ) (No./ft3)] 19.3 52-39-41 470.80 326 0.9628 37-40-55 470.80 382 0.8216 8.7 36-39-43 837.80 261 2.1399 40-40—43 873.30 308 1.8903 28-20—20 482.80 187 1.7212 —10 49-54-58 2146.67 284 5.0391 45-44—48 1826.67 220 5.5354 ~20 57—46—62 2750.00 445 4.1199 72-75-46 3216.67 482 4.4491 32-27-34 1550.00 198 5.2189 25—28-51 1900.00 205 6.1789 —30 42-39-30 2120.00 295 4.9717 33-31—31 1900.00 277 4.5728 73-78—80 4620.00 484 6.3636 73-54-61 3759-99 432 5.8025 —40 40—38-26 2426.67 400 4.0444 28-29-35 2146.67 369 3.8783 114 TABLE A-4.-—Influence of air flow rate gradient 02 transport characteristics of air-borne bacteria with 1.5 ft opening and aerosol flow rate of 10 ft3/min. Air Flow Rate Steady-State Mass Concentration Transfer Difference Transfer Flux Gradient Coefficient 3 3 3 [No./min-ft2 (ft .min.) (No./ft ) (No./min.) (No./ft ) (No./ft3)] 29.3 48—53 540.35 323 1.1153 63-64-65 684.80 307 1.4871 18.7 52-69-93 1519.40 357 2.837 60—67-92 1554.90 356 2.911 10 89-91—88 2680.00 352 5.0758 98—97-93 2880.00 431 4.4548 —10 38-37-35 1833.33 291 4.2001 33-25-34 1533-33 263 3-8868 24-30—27 1350.00 170 5.2941 23-48-24 1416.67 168 5.6217 -20 39—30-34 2060.00 306 4.4880 26—28—32 1720.00 262 4.3765 -30 40—39—37 2706.67 486 3.7129 35—36—32 2403.33 446 3.5924 115 TABLE A—5.-—Inf1uence of air flow rate gradient o5 transport characteristics of air—borne bactegia with 1.5 ft opening aerosol flow rate of 50 ft /min. Air Flow Rate Steady-State Mass Concentration Transfer Difference Transfer Flux Gradient Coefficient 3 3 3 [No./min. ft2 (ft /m1n.) (No./ft ) (No./min.) (No./ft ) (No./ft3)] 39.3 35—35—42 399.47 398 0.6691 33-28-35 342.40 342 0.6675 26—27-29 292.47 186 1.0483 17-11—17 160.50 154 0.6948 28.7 38-39-43 852.00 255 2.2275 38—44—37 844.89 207 2.7211 92-60—53 1455.50 389 2.4944 42—34-45 859.10 300 1.9091 20 56-68-63 1869.90 425 2.9333 71-49-54 1740.00 403 2.8784 31—34—39 1040.00 225 3.0815 10 31—23—27 1080.00 187 3.8503 28—21—12 813.33 174 4.6743 15-38-33 1013.30 191 3.5369 -10 50—58-48 3120.00 366 5.6831 36-31—34 2020.00 303 4.4444 30-32-28 1800.00 272 4.4118 -20 44-33—32 2543 33 447 3.7932 29—29-34 2146.67 418 3.4237 116 TABLE A—6.——Influence of width at initial point of mixing on transport characteristic§ of air-borne bacteria at / 50 ft min. Partition Steady-State Mass Concentration Transfer Width Transfer Flux Gradient Coefficient 3 3 [No./min§ft2 (in.) (No./ft ) (No./min.) (No./ft ) (No./ft )1 0.3125 44—46-51 2350.00 272 5.7598 47-35-41 2050.00 264 5.1768 78-55—72 3415.67 321 7.0959 65—60—34 2650.00 291 6.0710 0.5625 24-27—33 1400.00 253 3.6891 33—26—38 1616.67 246 4.3812 25—32—32 1483.33 209 4.7315 40-27-36 1716.67 243 4.7097 29-23-42 1566.67 308 3.3911 43—52—47 2366.67 330 4.7811 4.5 53-59—51 2716.67 428 4.2316 40-35-42 1950.00 350 3.7143 0.0625 54—47-50 2516.67 375 4.4741 70-51-68 3150.00 387 5.4264 36-40-38 1900.00 216 5.8642 54—29—36 1983.00 238 5.5556 TABLE A-7.--Influence of width at initial point of mixing 117 on transport characteristics of air-borne bacteria at 40 ft3/min. Partition Steady-State Mass Concentration Transfer Width Transfer Flux Gradient Coefficient 3 3 [No./min-ft2 (in.) (No./ft ) (No./min) (No./ft ) (No./ft3)] 0.0625 75-77-62 2853.33 358 5.3135 87-65-67 2920.00 342 5.6920 69-80-65 2853.33 415 4.5837 77-63-79 2920.00 396 4.9158 0.3125 57-40-42 1866.67 230 5.4106 71-60-57 2506.67 254 6.5792 0.5625 58-44-51 2040.00 297 4.5791 52-48—49 1986.67 284 4.6635 53-30-47 1733-33 310 3-7275 34-47-41 1626.67 305 3.5556 4.5 56-46-50 2026.67 297 4.5492 70-57-55 2426.67 322 5.0242 37-28-29 1253.33 267 3.1294 32-41—43 1546(67 249 4.1410 118 TABLE A-8.--Influence of width at initial point of mixing on transport characteristicg of air—borne bacteria at 30 ft /min. Partition Steady-State Mass Concentration Transfer Width Transfer Flux Gradient Coefficient 3 3 [No./min—ft2 (in.) (No./ft ) (No./min.) (No./ft ) (No./ft3)] 0.0625 28-38—31 970.00 283 2.2850 37-37-26 1000.00 242 2.7548 61-61—58 1800.00 375 3.2000 29-66-70 1650.00 382 2.8796 43-38-42 1230.00 332 2.4699 65-46—57 1680.00 373 3.0027 0.3125 50-59-66 1750.00 293 3.9818 57-67-69 1930.00 310 4.1505 0.5625 46-44-38 1280.00 348 2.4521 61-58-67 1860.00 437 2.8375 50-39-36 1250.00 315 2.6455 27-36-49 1120.00 293 2.5484 4.5‘ 16-28-13 579.00 237 1.6287 19-17-44 800.00 218 2.4465 31-35-39 1050.00 323 2.1672 27—44-40 1110.00 363 2.0386 119 TABLE A-9.——Influence of air flow rate gradient on transport characteristics of air-borne bacteria with 0.3125 in. partition. Air Flow Rate Steady—State Mass Concentration Transfer Difference Transfer Flux Gradient Coefficient 3 3 3 [No./min—ft2 (ft /min.) (No./ft ) (No./min.) (No./ft ) (No./ft3)1 “39.3 15-12-19 164.07 158 0.6923 11-9-10 107.00 134 0.5323 75-67-85 809.63 669 0.8068 56-48-42 520.73 551 0.6301 28.7 46-42—42 923.00 302 2.0375 38—38—37 802.30 229 2.3357 54—41—58 1086.30 415 1.7451 79-98-54 1640.01 425 2.5726 20 26-27-15 680.00 144 3.1482 25-38—33 960.00 204 3.1373 10 35-41—35 1560.00 197 5.2792 96-78-92 680.00 544 4.3464 101-117-116 3546.67 673 4.4114 0 44-46—51 2350.00 272 5.7598 47-35-41 2050.00 264 5.1768 78-55-72 3416-67 321 7-0959 65-60—34 2650.00 291 6.0710 —10 10-13-11 680.00 97 4.6735 120 TABLE A-10.--Inf1uence of air flow rate gradient on trans— port characteristics of air—borne bacteria with 0.5625 in. partition. Air Flow Rate Steady—State Mass Concentration Transfer Difference Transfer Flux Gradient Coefficient 3 3 3 [No./min-—ft2 (ft /min.) (No./ft ) (No./min.) (No./ft ) (No./ft3)] 39.3 12—15—11 135.53 131 0.6897 10—13-8-16 125.73 155 0.5408 18—7-13 135.53 108 0.8366 13-12-16 146.23 119 0.8192 28.7 25—18-21 454.40 178 1.7019 21-26—20 475.70 187 1.6959 20 21-17-13 510.00 110 3.0909 16—15—8 390.00 88 2.9545 10 37—28—28 1240.00 216 3.8272 23-14-15 693-33 137 3-3739 23-16-16 733 33 153 3.1954 0 24—27-33 1400.00 253 3.6891 33-26—38 1616.67 246 4.3812 25—32—32 1483.33 209 4.7315 40-27—36 1716.67 243 4.7097 29-23-42 1566.67 308 3.3911 43—52-47 2366.67 330 4.7811 -10 31-18-19 1360.00 214 4.2368 16—17~14 940.00 171 3.6647 13-9-12 680.00 130 3.4872 121 TABLE A—11.—-Influence of air flow rate gradient on trans— port characteristics of air—borne bacteria with 1.0 ft opening. Air Flow Rate Steady—State Mass Concentration Transfer Difference Transfer Flux Gradient Coefficient 3 3 3 [No./min-ft2 (ft /min.) (No./ft ) (No./min.) (No/ft ) (No./ft3)] 39.3 83—68-82 831.03 305 2.7247 37-32—21 321.00 198 1.6212 33-39-38 392-33 329 1.1925 57-56-61 620.60 368 1.6864 73-58-59 677.67 321 2.1111 54-38-61 545.70 293 1.8625 50-68—51 602.77 315 1.9135 28.7 25-28-29 653.20 271 2.4103 28-35-25 624.80 257 2.4311 23-28—29 568.00 216 2.6296 41-39-34 809.40 311 2.6026 20 57-43-39 1390.00 356 3.9045 39—41—41 1210.00 327 3.7003 60—53-54 1670.00 375 4.4533 67-70—76 2130.00 385 5.5325 10 58—62—50 2266.67 402 5.6385 33—27-32 1226.67 265 4.6289 38-37-47 1626.67 267 6.0924 38—23—29 1200.00 243 4.9383 0 29-26—30 1416.67 291 4.8683 41—43-53 2283.33 363 6.2903 44—48—40 2200.00 456 4.8246 48-59—61 2800.00 479 5.8455 —10 12—17—30 1179.99 168 4.6825 43-40—37 2400.00 380 6.3158 16—32—25 1279.99 162 5.2750 122 TABLE A—12.—-Inf1uence of air flow rate gradient on trans— port characteristics of air—borne bacteria with 0.5 ft opening. Air Flow Rate Steady-State Mass Concentration Transfer Difference Transfer Flux Gradient Coefficient 3 3 3 [No./min-ft2 (ft /min.) (No./ft ) (No./min.) (No./ft ) (No./ft3)] 39.3 40—42—45 452.97 333 2.7205 59—48-60 595.63 410 2.9055 43-51—62 556.40 440 2.5291 28.7 19-32—20 540.10 197 5.1178 17—40—28 603.50 268 4.5037 56-38—50 1022.40 428 4.7776 47—56—43 1036.60 447 4.6380 20 23-26—21 700.00 235 5.9575 30-19—13 620.00 240 5.1667 36-30-31 970.00 335 5.7910 39-24-30 930.00 341 5.4546 10 34-37-43 1520.00 373 8.1501 49-39-46 1786.67 376 9.5036 19-20-32 946.67 207 9.1465 25-30-27 1093-33 241 9.0733 0 42-45-39 2100.00 474 8.8607 46-45-30 2016.60 433 9.3148 52-29—17 1533-33 382 8.5515 18—22—27 1116.67 242 9.2287 —10 21-21-19 1020.00 298 6.8456 12-13—8 660.00 229 5.7642 123 TABLE A—13.—-Influence of temperature gradient on the transport characteristics of air-borne bacteria at 61 ft3/min. Temperature Steady—State Mass Concentration Transfer Gradient Transfer Flux Gradient Coefficient 3 3 [No./ft2—min (OF.) (No./ft ) (No./min) (No./ft ) (No./ft3)] -14 16-13-17 935.33 272 2.2925 20-13—14 955.67 252 2.4985 —12 24—21-20 1321.67 515 1.7109 18—24—26 1382.67 433 2.1288 -11 28-27—33 1789.33 376 3.1726 21-19—25 1321.67 350 2.5175 18—17—20 1100.00 248 2.9569 16-19-20 1100.00 257 2.8534 -10 41-34—43 2399-33 504 3.1737 38-37—38 2297.67 473 3.2384 —6 48—39-56 2907.67 453 4.2791 48-36—58 2887.33 404 4.7646 —5 40-43-41 2521.33 335 5.0176 34-34-38 2155.33 341 4.2138 0 22—21-25 1382.67 189 4.8771 22—27—24 1484.33 170 5.8209 24-22-25 1443.67 154 6.2496 54—50-44 3009.33 405 4.9536 5 29-27-29 1728.33 231 4.9880 29—22—25 1545.33 215 4.7917 19—25—15 1199.67 160 4.9986 30-23-35 1789-33 162 7-3635 10 16—29—18 1281.00 149 5.7315 21—29-19 1403.00 144 6.4954 26-26—21 1484.33 173 6.5970 25—20-24 1403.00 137 6.8273 12 28—30-19 1565.67 118 8.8455 23-20—24 1362.33 117 7.7626 12.5 28-28-21 1423.33 121 7.8421 17—25—27 1403.00 118 7.9266 124 TABLE A—14.—-Turbu1ence data at various locations and flow conditions. 72 /_ u f Location UT? 11'2 ' E (ft3/min) (in.) (ftZ/minZ) (ft/min) (%) 61-61 0 2.053 1.432 37.60 58-58 0 1.68 1.295 35.75 50-50 0 1.675 1.285 42.20 40-40 0 1.355 1.162 46.50 30-30 0 0.7503 0.866 46.15 20-20 0 0.419 0.646 48.60 50-20 8 0.025 0.158 11.89 50-20 6 0.0995 0.315 23.70 50-20 4 0.513 0.716 53.75 50-20 2 1.106 1.052 73.80 50-20 1 0.5069 0.711 41.70 50—20 0 0.7858 0.885 40 35 50-20 -1 1.343 1.159 43.10 50-20 -2 0.650 0.8055 27.20 50-20 -4 1.981 1.408 46.00 50-20 -6 1.277 1.128 36 85 20-20 -6 0.0041 0.064 4.8 20-20 -4 0.0096 0.0979 7.36 20-20 -2 0.112 0.3343 25.15 20-20 -1 0.164 0.4045 30.40 30-30 -4 0.503 0.710 37.75 30-20 8 0.2538 0.5035 26.82 40-40 8 0.4186 0.646 25.85 50-50 8 1.161 1.078 35.20 58—58 8 0.437 0.661 18.22 61-61 8 1.9994 1.412 37.03 20-60 8 1.415 1.189 31.68 50-70 0 1.479 1.215 32.65 50-60 0 2.538 1.59 46.65 40—70 0 1.647 1.282 37.30 30-70 0 2.8329 1.68 53.80 40-60 0 1.2095 1.099 35.18 30-60 0 1.942 1.393 49.60 40-50 0 0.675 0.821 29 20 20-50 0 1.4634 1.209 5.10 30-40 0 0.475 0.689 31.40 20-40 0 0.4334 0.6585 35.10 20-60 0 1.2057 1.098 43.90 30-50 0 1.0579 1.027 41.10 10-50 0 0.7823 0.884 47.15 10-40 -2 0.0587 0.2332 30.20 10-40 -6 0 0 0 10-40 0 0.6474 0.8035 50.75 125 000 0 00.0 00.00 000.0 000.0 000.0 000.0 0000.0 0.00-00 000.0 00.0 00.0 000.0 000.0 000.0 000.0 0000.0 0 00-00 000.0 00.0 00.0 000.0 000.0 000.0 000.0 0000.0 00-00 000.0 00.0 00.0 000.0 000.0 000 0 000.0 0000.0 00-00 000.0 00.0 00.0 000.0 000.0 000.0 000 0 0000.0 00-00 000.0 00.0 00.0 000 0 000.0 000.0 000.0 0000.0 00-00 000.0 00.0 00.00 000 0 000.0 000.0 000.0 0000.0 0.00-00 000.0 00.0 00.0 000.0 000.0 000.0 000.0 0000.0 0.00-00 000.0 00.0 00.0 000.0 000.0 000.0 000.0 0000.0 00-00 000.0 00.0 00.0 000.0 000 0 000.0 000.0 0000.0 00-00 000.0 00.0 00.0 000.0 000.0 000.0 000.0 0000.0 00-00 000.0 00.0 00.0 000.0 000.0 000.0 000.0 0000.0 00-00 000.0 00.0 00.00 000.0 000.0 000.0 000.0 0000.0 0.00-00 000.0 00.0 00.0 000.0 000.0 000.0 000.0 000.0 0.00-00 000.0 00.0 00.0 000.0 000.0 000 0 000.0 0000.0 00-00 000.0 00.0 00.0 000.0 000.0 000.0 000.0 0000.0 00-00 000.0 00.0 00.0 000.0 000.0 000 0 000.0 0000.0 00-00 000.0 00.0 00.0 000.0 000.0 000.0 000.0 000.0 00-00 000.0 00.0 00.0 000 0 000.0 000 0 000.0 0000.0 0.00-0.00 000.0 00.0 00.0 000.0 000.0 000.0 000.0 000.0 00-00 000.0 00.0 00.0 000.0 000.0 000 0 000 0 000 0 00-00 000.0 00.0 00.0 000 0 000.0 000.0 000.0 000.0 00-00 000.0 00.0 000.0 000.0 000.0 000.0 000.0 000.0 00-00 000.0 00.0 000.0 000.0 000.0 000.0 000.0 000.0 00-00 0.000 0000 00050 0:00 0005\000 0005\0000 00-00 x 0050 00:0E\0000 0:0ex0000 0 00 0 .0 0.0V 00 00000000 0.0 0 x .mcoHuHUCOO 3O0m 0c0 020000000 mzowp0> p0 0000 :o0mpmamHQ-I.m01< mqm<9 REFERENCES 126 REFERENCES Albertson, M. L., Dai, Y. B., Jensen, R. A. and Rouse, H. 1948 Diffusion of submerged Jets. Proc. Am. Soc. Civil Engrs. 14:1571. Batchelor, G. K. 1950 Note on free turbulent flows, with special reference to the two-dimensional wake. J. Aero. Sci. 11:441. Batchelor, G. K. 1957 Diffusion in free turbulent shear flows. J. Fluid. Mech. 3:67. Bourdillon, R. B., Lidwell, 0. M. and Thomas, J. C. 1941 A slit sampler for collecting and counting air— borne bacteria. J. Hyg. 41:197. Boussinesq, J. 1877 Essai sur la theorie des caux courantes. Memories presentes par divers savants a l'Academie des Sciences 23. Cerna, M. 1961 Study of microbiological purity of air in dairies. Promysl. Protravin. 12:374. Cheng, S. I. and Kovitz, A. A. 1958 Mixing and chemical reaction in the laminar wake of a flat plate. J. Fluid. Mech. 4:64. Coles, D. 1956 The law of the wake in the turbulent boundary layer. J. Fluid. Mech. 1:191. Corrsin, S. 1943 Investigation of flow in an axially symmetrical heated Jet of air. NACA Wartime Report W-94. Crank, J. 1956 The Mathematics of Diffusion. Clarendon Press, Oxford. Csanady, G. T. 1963 On the energy balance of a turbulent mixing layer. J. Fluid. Mech. 15:545. 127 128 DallaValle, J. M. 1948 Micromeritics, the Technology of Fine Particles. Pitman Publishing Company. Decker, H. M., Buchanan, L. M., Hall, L. B. and Goddard, K.R. 1962 Air Filtration of Microbial Particles. Public Health Service Publication No. 953. U. S. Govern- ment Printing Office, Washington, D. C. DeOme, K. B. et al. 1944 The effect of temperature,humidity, and glycol vapor on the viability of air-borne bacteria. Am. J. Hyg. 42:239. Druett, H. A. and May, K. R. 1952 A wind tunnel for study of air—borne infection. J. Hyg. 52:69. Dryden, H. L. and Kuethe, A. M. 1930 Effect of turbulence in wind tunnel measurements. NACA Tech. Report No. 342. Eckman, D. P. 1950 Industrial Instrumentation. John Wiley and Sons, Inc. New York. Fage, A. and Falkner, V. M. 1932 Note on experiments on the temperature and velocity in the wake of a heated cylindrical obstacle. Proc. Roy. Soc. A135z678. Flow Corporation. 1958 Model HWB2 hot wire anemometer theory and instruction, Flow Corp. Bulletin No. 37B. Forstall, W., Jr. and Shapiro, A. H. 1950 Momentum and Mass Transfer in coaxial gas jets. J. Appl. Mech. 72:399. Frenkiel, F. N. 1953 Turbulent Diffusion: Mean concentration distribu- tion in a flow field of homogenious turbulence. Adv. Appl. Mech. 3:61. Goldstein, S. 1930 Concerning some solutions of the boundary layer equation in hydrodynamics. Proc. Camb. Phil. Soc. 22:1. Goldstein, S. 1933 On the two-dimensional steady-flow of a viscous fluid behind a solid body—I. Proc. Roy. Soc. London. A142z545. 129 Greene, V. W.- 1965 Personal communication. School of Public Health. University of Minnesota. Aug. 10. Hayakaw, I. and Poon, C. P. 1965 Short storage studies on the effect of tempera- ture and relative humidity on the viability of air—borne bacteria. Ind. Hyg. J. 26:150.. Heldman, D. R., Hedrick, T. 1., and Hall, C. W. 1964 Air-borne microorganism populations in food packaging areas. J. Milk & Food Tech. 21:245. Heldman, D. R., Hedrick, T. I. and Hall, C. W. 1965' Sources of air—borne microorganisms in food processing areas-drains. J. Milk & Good Tech. 28:41. Henderson, D. W. 1952 An apparatus for the study of air—borne infection. J. Hyg. 32:53. Hinze, J. O., et a1. 1948 Transfer of heat and matter in the turbulent mixing zone of an axially symmetrical jet. Appl. Sci. Res. A-1:435. Hinze, J. O. 1959 Turbulence, An Introduction to its Mechanism and Theory. McGraw-Hill Co., Inc. New York. Hollingdale, S. H. 1940 Stability and configuration of the wakes produced by solid bodies moving through fluids. Phil. Mag. 23 (7):209. Howarth, L. 1934 Concerning the velocity and temperature distribu— tions in plane and axilly symmetrical jets. Camb. Phil. Soc. Proc. 34:185. Kalinske, A. A. and Pien, C. L. 1944 Eddy diffusion. Ind. Eng. Chem. 32:220. Kaye, S. 1949 The sterilizing action of gaseous ethylene oxide. Am. J. Hyg. 32:290. Kethley, T. W., Cown, W. B. and Fincher, E. L. 1956 A system for the evaluation of aerial disinfec— tants. Appl. Microbiol. 4:237. 130 Kethley, T. W., Cown, W. B. and Fincher, E. L. 1957 The nature and composition of experimental bacterial aerosols. Appl. Microbiol. 3:1. Kuethe, A. M. 1935 Investigation of the turbulent mixing region formed by jets. J. Appl. Mech. Trans. ASME. 31:A87. Labots, H. 1961 The estimation of the bacteria count of the atmosphere in dairy factories. Off. Org. K. Ned. Zuivelb. 33:772. Laurell, G., Lofstrom, G., Magnusson, J.H. and Ouchterlony,O. 1949 Air-borne infection. 2. A report on methods. Acta Med. Scand. 134:189. Leif, W. R. and Krueger, A. P. 1950 Studies on the experimental epidemiology of respiratory infections. J. Infectious Diseases. 21:103. Leipmann, H. W. and Laufer, J. 1947 Investigation of free turbulent mixing. NACA TN 1257. Lock, R. C. 1951 The velocity distribution in the laminar boundary layer between parallel streams. Quart. J. Mech. and Appl. Math. 4:42. Mackay, I. 1952 Hexylresorcinol as an aerial disinfectant. J. Hyg. 32:82. Olson, H. C. and Hammer, B. W. 1934 Numbers of microorganisms falling from the air in dairy plants. J. Dairy Sci. 11:613. Pai S I. , . 1949 Two-dimensional jet mixing of a compressible fluid. J. Aero. Sci. 12:463. Pai, S. I. 1954 Fluid Dynamics of Jets. D. Van Nostrand Company, New York. Pai, s. I. 1955 On turbulent jet mixing of two gases at constant temperature. J. Appl. Mech. 22:41. 131 Pai, s. I. 1956 Laminar jet mixing of two compressible fluids with heat release. J. Aero. Sci. 23:1012. Pasquill, F. 1962 Atmospheric Diffusion. D. Van Nostrand Company, Ltd. Princeton. Porter, F. E., et a1. 1963 The dynamic behavior of aerosols. Annals of the New York Academy of Sci. 105:45. Prandtl, L. 1925 Berichte uber Untersuchungen zur ausgebildetan Turbulenz. ZAMM. 21:257. Reichardt, H. 1941. Uber eine neue Theorie der freier Turbulenz. ZAMM. 21:257. Reichardt, H. 1944 Impuls-und Warmeaustausch in freier Turbulenz. ZAMM. 24:268. Rentschler, H. C. 1942 Bactericidal action of ultraviolet radiation on air-borne organisms. J. Bacteriol. 44:85. Robertson, 0. H., Puck, T. T. and Wise, H. 1946 The construction and operation of experimental rooms for the study of air-borne infection. J. Exptl. Med. 24:559. Rosebury, T. 1947 Experimental Air—borne Infection. The Williams & Wilkins Company, Baltimore. Schlichting, H. 1960 Boundary Lgyer Theory. McGraw-Hill Book Company, New York. Sherwood, T. K. and Woertz, B. B. 1939 The role of eddy diffusion in mass transfer between phases. Trans. AI Ch. E. 35:5r7. I 1 Silver, S. D. 1946 Constant flow gassing chambers: Principles in- fluencing design and operation. J. Lab. Clin. Med. 31:1153. 132 Tanner, H. G. 1963 Evanescence of cloud chamber aerosols. Annals of the New York Academy of Sci. 105:27. Taylor, G. I. 1915 Eddy motion in the atmosphere. Phil. Trans. Roy. Soc. A215zl. Taylor, G. I. 1921 Diffusion by continuous movements. Proc. London Math. Soc. 22:(2):196. Taylor, G. I 1932 The transport of vorticity and heat through fluids in turbulent motion. Proc. Roy. Soc. A135z685. Taylor, G. I. 1935 Statistical theory of turbulence. Part 1. Proc. Roy. Soc. A151:444. Tomotika, S. 1938 Application of the modified vorticity transport theory to the turbulent spreading of a jet of air. Proc. Roy. Soc. A165z65-72. Torda, T. P., Ackermann, W. O. and Burnett, H. R. 1953 Symmetric turbulent mixing of two parallel streams. J. Appl. Mech. 22:63. Towle, W. L. and Sherwood, T. K. 1939 Eddy diffusion: Mass transfer in the central portion of a turbulent air stream. Ind. Eng. Chem. 31:457. Townsend, A. A. 1949 Momentum and energy diffusion in the turbulent wake of a cylinder. Proc. Roy. Soc. A197zl24-140. Townsend, A. A. 1956 The Structure of Turbulent Shear Flow. Cambridge University Press, New York. Treybal, R. E. 1955 Mass Transfer Operations. McGraw—Hill Book Co., Inc., New York. Twort, C. C., Baker, A. H., Finn, S. R. and Powell, E. O. 1940 The disinfection of closed atmospheres with germicidal aerosols. J. Hyg. 42:253. 133 Urban, E. C. J. 1954 Two chambers for use in exposing laboratOry animals to inhalation of aerosols. Arch. Ind. Hyg. and Occupational Med. 2:62. Webb, S. J. 1959 Factors affecting the viability of air-borne bacteria. I. Bacteria aerosolized from distilled water. Can. J. Microbiol. 3:649. Weiss, E. and Stegeler, J. C. 1952 A cloud chamber for the uniform air-borne inocu- lation of mice. J. Infectious Diseases. 22:13. Wells, W. F. 1933 Apparatus for study of the bacterial behavior of air. Am. J. Pub. Health. 23:58. Wells, w. F. 1955 Air-borne Contagion and Air Hygiene. Harvard University Press, Cambridge. Whitfield, W. J. 1963 Design of a dust controlled clean bench and hood utilizing laminar air flow. Presented at 2nd. Annual Convention of AACC. Boston, Mass. Apr. 30. Wolf, H. W., Skaliy, P., Hall, L. B, Harris, M. M., Decker, H. M., Buchanan, L. M. and Dahlgren, C. M. 1959 Samplinngicrobiological Aerosols. Pub. Health Rept. 74. UT S. Government Printing Office. Washington, D. C. Wu, T. Y. T. 1962 A wake model for free—streamline flow theory— Part I. J. Fluid Mech. 13:161. "1111711111Iii-1711111111111“ 3 293 030