increase of frictional resistance in closed conduit systems

124
Increase of frictional resistance in closed conduit systems fouled with biofilms by Mark Douglas Groenenboom A thesis submitted in partial fulfillment of the requirements for the degree of Master of Science in Mechanical Engineering Montana State University © Copyright by Mark Douglas Groenenboom (2000) Abstract: Bacterial biofilms form slimy deposits in closed conduits and are responsible for significant pressure loss in many water and power systems. Bacteria bind to conduit surfaces via viscous-elastic polymers secreted by the microorganisms; the cells and the polymer matrix form a biofilm. As a biofilm covers the interior of the pipe wall, the solid interface is replaced with the pliable and complex topography of the biofilm. As this occurs the traditional methods used to predict losses in non-fouled systems become obsolete. In order to effectively deal with this problem, a full understanding of the mechanism of loss needs to be determined. The research presented in this thesis, both empirical and analytical, provides a further understanding of the problem of biofouling of closed conduit systems.

Upload: khangminh22

Post on 10-Jan-2023

0 views

Category:

Documents


0 download

TRANSCRIPT

Increase of frictional resistance in closed conduit systems fouled with biofilmsby Mark Douglas Groenenboom

A thesis submitted in partial fulfillment of the requirements for the degree of Master of Science inMechanical EngineeringMontana State University© Copyright by Mark Douglas Groenenboom (2000)

Abstract:Bacterial biofilms form slimy deposits in closed conduits and are responsible for significant pressureloss in many water and power systems. Bacteria bind to conduit surfaces via viscous-elastic polymerssecreted by the microorganisms; the cells and the polymer matrix form a biofilm. As a biofilm coversthe interior of the pipe wall, the solid interface is replaced with the pliable and complex topography ofthe biofilm. As this occurs the traditional methods used to predict losses in non-fouled systems becomeobsolete. In order to effectively deal with this problem, a full understanding of the mechanism of lossneeds to be determined. The research presented in this thesis, both empirical and analytical, provides afurther understanding of the problem of biofouling of closed conduit systems. 

INCREASE OF FRICTIONAL RESISTANCE IN CLOSED

CONDUIT SYSTEMS FOULED WITH BIOFILMS

by

Mark Douglas Groenenboom

A thesis submitted in partial fulfillment of the requirements for

the degree of

Master of Science in

Mechanical Engineering

MONTANA STATE UNIVERSITY-BOZEMAN Bozeman, Montana

April 2000

/

f i y n

G,n>-4APPROVAL

of a thesis submitted by

Mark Douglas Groenenboom

This thesis has been read by each member of the thesis committee and has been found to be satisfactory regarding content, English usage, format, citations, bibliographic style, and consistency, and is ready for submission to the College of Graduate Studies.

Approved for the Department of Mechanical Engineering

Dr. Doug Cairns __ t(Signature)

Approved for the College of Graduate Studies

Dr. Bruce McLeod(Signature) Date

STATEMENT OF PERMISSION TO USE

In presenting this thesis in partial fulfillment of the requirements for a master’s

degree at Montana State University - Bozeman, I agree that the library shall make it available to borrowers under the rules of the Library.

IfI dictated my intention to copyright this thesis by including a copyright notice

page, copying is allowable only for scholarly purposes, consistent with “fair use “ as

prescribed in the U.S. copyright Law. Request for permission for extended quotation

from or reproduction of this thesis in whole on in parts may be granted only by the

copyright holder. . -

Signature

Date

IV

To my family, friends, and especially to Monica, who has always been there for me.

V

ACKNOWLEDGEMENTS

This project was possible thanks to National Science Foundation.

I would like to thank all the members of my committee and the structure function

group at the Center for Biofilm Engineering. Particularly, Dr. Halulc Beyenal for his

sharing of his wealth of knowledge regarding biofilms. In addition, I would like to thank

John Neuman for his technical assistance. The help was greatly appreciated.

Vl

TABLE OF CONTENTS

LIST OF TABLES............................................................................................................. ix

LIST OF FIGURES............................................................................................................ x

LIST OF VARIABLES....................... xiii

ABSTRACT............... xiv

1. INTRODUCTION........................................... I

Introduction............................................................ IBackground...................................................................................................................2 'Literature Review...........................................................!..............................................8

2. APPROACHES TO REDICTING VELOCITY PROFILES INHETEROGENEOUS BIOFILMS........................................................................... 20

Introduction................................................................................................................. 20Velocity Profile of an Actual Biofilm.............. ,...21Comparison of Flow Velocity within a Biofilm to AtmosphericFlow in a Vegetative Canopy.................................. 23Comparison to Flow within an Artificial Canopy....:................ 28Evaluation of Flow through a Model BiofilmComposed of Cylindrical Elements.................. 33Conclusion.................................................................................................................. 37

3. MATERIALS AND METHODS................................. ...38

Introduction................................................. 38System Layout.......................................................................... 38Growth Medium and Sterilization...... ......................... ....... ,...................................... 41Method of Biofouling....... ............................................................... 42Measurement of Pressure Loss and Calculation of Friction Factors.......... ................43

4. CLOSED CONDUIT SYSTEM HEADLOSS AS AFUNCTION OFBlOFILMSTRUCTURE............................................................... 46

Introduction.........................................................................................................■....... 45Porosity Study............................................................................................................. ,47

a) Porosity Study Materials and Method.............................................................. 47b) Biofilm Imaging and Image Analysis............................................................... 48c) Results of the Porosity Study........................................................................ 48

Experiment to Quantify a Relationship betweenBiofilm Structure and Energy Loss .................................................. 51a) Materials and Methods.......................................................................................51b) Results.................................................................................... 53c) Dimensional Analysis........................................................................................54

Discussion and Conclusion........ .................................................................. ;............57

5. THE PHEOMENON OF INCREASING FRICTION FACTORWITH INCREASING REYNOLDS NUMBERIN BIOFOULED SYSTEMS....................................................................................61

Introduction...................................................................................................;............ 61Experiment Specific Details and Methods......................................................... 64Results........................................................................ 65Experimental Observations..........................................................................................68Discussion....,............................................................... 70Conclusions..................................................................................................................75

6. SIMULATING A FILAMENTOUS BIOFILM SURFACE.... ..................................76

Introduction.,.............................................................................................. 76Materials and Methods.................................................................................................76Results....................................................................................................... 79Discussion................................................................................................................ 80

7. A RELATIONSHIP BETWEEN MEAN FLOW VELOCITY AND FRICTIONFACTORS EXHIBITED BY BIOFOULED SYSTEMS......................................... 82

Introduction..... .................................................................................. 82System Layout and Reactors.......................................................... 82Constrictional Effects...................................................................................................83Results..........................................................................................................................84Discussion............................................................................................................. 86

8. CONCLUSION AND FUTURE WORK................................................ 91

vii

V lll .

REFERENCES CITED.......... :.......................................... ...............................................96

APPENDICES............................................................................................................... 100

APPENDIX A ......................................................................................... • ........... io iSpreadsheet Data / Measurements of Streamer..................................................102Lengths for the Experiment of Chapter 4 ............................................ ........... ...102Sample of Images Taken for the Experiment of Chapter 4 ..'.................. ............103

APPENDIX B ........ ........................ .............:........................................................... 104Spreadsheet Data for Chapter 5 Experiment....................................... ................105

IX

LIST OF TABLES

Table Page

Table 2.1 Attenuation Coefficients for Various Canopies............... .................................26

Table 5.1 Friction Factor Values and Comparison Blasius Solution................................67

Table 7.1 Experimental Parameters for the 12 Experiments............................................83

Table 7.2 Reynolds number, Velocity and MaximumFriction Factor of the 12 Experiments............... .............................................85

LIST OF FIGURES

Figure , Page

Figure I . I Diagram of Forces Acting on a Slug of FluidTraveling through a Pipe........................................ 2

Figure 1.2 The Moody Diagram.................................................. 7

Figure 2.1 Image of 3-Specie Biofilm Showing the Locations atwhich Velocity Profiles Were Measured............................................. 22

Figure 2.2 Plot of Flow Velocity as a Function of Vertical Positionfor the Five Locations Shown in Figure 2.1 ....................,......................... 22

Figure 2.3 Flow Profile Generated by Equation 2.1 with anAttenuation Coefficient of 2 ..................................................................... 24

Figure 2.4 Nondimensional Wind Profiles for Various Canopies...................................25

Figure 2.5 Superimposing of Biofilm Flow Profile on Figure 2.4................. 26

Figure 2.6 Spherical Canopy Illustration........................................................................28

Figure 2.7 Flow Profile Generated using the Canopy Model Composed of Spherical Elements and AssociatedVariables from Stoltzenbach (1989).................................. 30

Figure 2.8 Predicted Flow Profile for Sample Biofilm...................................................32

Figure 2.9 Proposed Cylinder Model......................................................... 33

Figure 2.10 Drag Coefficient as a Function of Reynolds Number for a Cylinder............34

Figure 2.11 Plot of Biofilm Flow Velocity Profile with Plots ofEquation 2.9 for Various Values of B............;..........................................36

Figure 3.1 The General System Layout used for the Experiments ofChapters 4 through 7 ...................................:................. ...........................39

Figure 3.2 Photo of an Actual ClosedConduit Reactor System............................................... '........................... 40

Figure 3.3 An Impeller Pump also used to provide Flow throughthe Recycle Loop and Closed Conduit Reactor........................................ 41

Figure 3.4 Cole Parmer Flowmeter................................................ ............................ ..43

Figure 3.5 GPI Flowmeter.................. ......................... ................ ;.................. ............44

Figure 3.6 Plots of Pressure Loss vs. Time for Two Experiments involving an 8 mm Square Glass Tubeand the Three Species Bacteria................................................................. 44

Figure 3.7 A Cole Parmer Pressure Transducer ........... ......................................... .......45

Figure 3.10 The Emco 24 Volt Power Supply used toProvide Voltage to the Pressure Transducers........................................ ...45

Figure 4.1 Porosity Profile Graph Showing Low Porosity at the Biofilm Base........... 49

Figure 4.2 Close-Up Image of Substratum and Biofilmwith Filaments Protruding Toward the Bulk............................................ 50

Figure 4.3 Typical Sample of %” I.D. Vinyl Tubingfrom Second Experiment.................... 52

Figure 4.4 Image of Sample Showing the Traces used toMeasure the Filament Lengths and Base Thickness................................. 52

Figure 4.5 Non-Submerged Sample with Filaments Adhered to the Base Film........... 53

Figure 4.6 Plot of Friction Factor vs. Dimensionless Factor ofStreamer Length / Conduit Length........................................................... 53

Figure 4.7 Plot of Biofilm Thickness vs. Friction Factor.............................................. 53

Figure 4.8 Plot of Friction Factor vs. Dimensionless Factor of StreamerLength / Conduit Length........................................................................... 57

Figure 5.1 Graph Comparing the Friction Factor Generated by theNon-Fouled Vinyl Reactor to the Friction Factor PredictedUsing Tradition Methods for a Hydraulically Smooth Pipe..................... 66

xi

X ll

Figure 5.2 Plot of Friction Factor vs. Reynolds Number forVarious Days of Growth Riofilm..................................... .........................68

Figure 5.3 Images of the Cross-Sections Removed from Experiment..........................69

Figure 5.4 Side Views of Cross-Sections Taken from the Experimenton day 7 and 32..................................... ................. ..................................70

Figure 5.5 Illustration of Flow Profile without the Presence ofFilamentous Biofilm and Proposed Velocity Profile forConduit Fouled with Filamentous Riofilm..................... ..........................72

Figure 5.6 Graph Showing the Dramatic Difference in the Friction Factors for the Clean Reactor and the Reactor fouled with 22 days of Biofilm Growth...............................................................73

Figure 6.1 Images of Simulated and Naturally Biofouled conduit................................ 77

Figure 6.2 Image of a Cross Section of the Tube Lined with the AcrylicFur with Measurement Showing the Hydraulic Diameter............. ...........78

Figure 6.3 Plot of Friction Factor vs. Reynolds Number for the ConduitLined with the Artificial with the Acrylic Fur ...........................................79

Figure 6.4 Plot of the Friction Factor vs. Reynolds Number forBiofouled and Non Fouled Conduit................................................... 80

Figure 7.1 Graph Illustrating the Relationship between ReynoldsNumber and Maximum Friction Factors Exhibited by the Fouled Closed Conduit Reactors......................................................... 85

Figure 7.2. Plot Comparing Mean Flow Velocities toMaximum Friction Factors....................................................................... 86

Figure 7.3 Biofilm Structure for Bulk Fluid Velocity of 1.33 m/s................................ 89

Figure 7.4 Biofilm Structure for Bulk Fluid velocity of 2.85....................................... 89

xiii

LIST OF VARIABLES

j[ = Cross Section Area of Cylinder a = Attenuation Coefficient aK = Kouwen Equation Coefficient d = Maximum Sphere Radius a = Minimum Sphere Radius A = Kouwen equation coefficient Cd = Coefficient of Drag C = Skin Friction CoefficientD = Hydraulic Diameter n = Effective Diameterl j Off

EI = Bending Stiffness f = Friction Factor g = Gravitational Acceleration h = Roughness Heights k = Kouwen Roughness Height

= Canopy Height = Head Loss

I = Conduit Length L = Length M =MassH = Number of Cylinders Per Area

= Exit Pressure

t = Timeu = Fluid Velocity within Canopyu = Wall Friction Velocityjj, = Open Channel flow VelocityUh = Velocity within CanopyUh =V elocity at Canopy Top y = Mean Conduit Fluid Velocity Volejf = Effective Volume z — Height within Canopyr = Conduit Wall Shear Stressa = Sphere Volume Fraction# = Ratio of Average Sphere Radius to

. Canopy Height§ = Absolute Boundary Layer Thickness§* = Displacement Layer ThicknessEpl = Pressure Loss per Conduit Length£ = Average Roughness Heightjl = Kinematic Viscosityp = Fluid DensityQ = Momentum Layer ThiclcnessH = Dimensionless Pi Group

p. = Inlet PressureRe = Reynolds NumberSlam = Average Streamer LengthSI = Streamer Length Per Conduit Length

XlV

ABSTRACT

Bacterial biofilms form slimy deposits in closed conduits and are responsible for significant pressure loss in many water and power systems. Bacteria bind to conduit surfaces via viscous-elastic polymers secreted by the microorganisms; the cells and the polymer matrix form a biofilm. As a biofilm covers the interior of the pipe wall, the solid interface is replaced with the pliable and complex topography of the biofilm. As this occurs the traditional methods used to predict losses in non-fouled systems become obsolete. In order to effectively deal with this problem, a full understanding of the mechanism of loss needs to be determined. The research presented in this thesis, both empirical and analytical, provides a further understanding of the problem of biofouling of closed conduit systems.

I

INTRODUCTION

Introduction

Bacterial biofilms form slimy deposits in closed conduits and are responsible for

significant pressure loss in water distribution systems and hydraulic lines used in power

generation. Bacteria bind to conduit surfaces via viscous-elastic polymers secreted by the

microorganisms; the cells and the polymer matrix form a biofilm. As a biofilm covers

the interior of the pipe wall, the solid interface is replaced with the pliable biofilm. As

this occurs, the traditional methods used to predict losses in non-fouled systems become

unreliable. In order to effectively deal with this problem, a better understanding of the

. mechanism of energy loss is needed. The research presented in this thesis provides a

better understanding of the problem of biofouling in closed conduit systems.

This chapter explains the traditional approach to the problem of frictional losses in

closed conduit system's. Also presented is a review of the literature relevant to the

specific problem of increased frictional losses related to biofouling. Chapter 2 evaluates

three approaches to modeling flow in highly porous heterogeneous biofilms. Chapter 3

provides the materials and methods used in the laboratory to generate the results given in

Chapters 4 through 7. Chapter 4 quantitatively relates increased frictional losses to

biofilm structure. Chapter 5 investigates an interesting relationship between frictional

2

losses and Reynolds number in systems fouled with biofilm, and Chapter 6 shows the

results of an attempt to artificially simulate the fluid biofilm interface in a closed conduit.

Chapter 7 shows how losses in biofouled systems are better related to fluid velocity than

to Reynolds numbers based on pipe diameters.

Background

As a slug of fluid flows through a straight closed conduit, it is subject to three

forces. Pressures acting on the inlet and exit areas account for two of the forces, the

remaining force is that of shear or friction which is created at the fluid-conduit interface.

Figure 1.1 Diagram of Forces acting on a Slug of Fluid Traveling through a Pipe.

Figure 1.1 illustrates these forces for a slug of fluid with velocity, V , traveling

through a circular conduit of length, I, and radius, r . With acceleration assumed to be

zero, the resulting force balance is given by

3

p ,nr2 - P eW2 - r w2wl = 0' 1.1

where p; is the inlet pressure, r is the pipe radius, pe is the exit pressure, Tw is the shear

stress caused by the conduit walls, and I is the pipe length. The change or loss in pressure

experienced by the slug of fluid can then be expressed as

1.2r

Equation 1.2 illustrates that the shear force exerted on the fluid by the pipe wall

causes the drop in pressure experienced across the slug of fluid traveling down the pipe.

Equation 1.2 also shows the direct relationship between the length of the pipe and the

pressure loss.

In many engineering applications, it is necessary to predict pressure losses in closed

conduits. Hence, the shear force and resulting pressure loss formed the focus of a

substantial amount of research during the first half of the 20th century. For laminar flow,

predicting losses is relatively simple. By applying the no-slip boundary condition at the

fluid-wall interface, along with the boundary condition of finite velocity at the pipe

center, one can analytically solve for the flow profile [I]. The resulting flow profile for

conduits of circular cross section is parabolic. From this velocity profile, shear is

determined. The results show that for laminar flow, shear stress is linearly related to the

Reynolds number, given by equation 1.3 [2],

Re 1.3

4

where p is the fluid density, V is the fluid velocity, D is the hydraulic diameter, and fi

is the dynamic fluid viscosity. For turbulent flow, no analytical solution is available and

prediction of shear stress and pressure loss is much more complex.

Munson states [2],

“Turbulent flow can be a very complex difficult topic - one that has yet defied a rigorous theoretical treatment. Thus, most turbulent flow analyses are based on experimental data and semi-empirical formulas, even if the flow is fully developed. These results are given in dimensionless form and cover a very wide range of flow parameters, including arbitrary fluids, pipes and flow rates.”

This has indeed been the approach used to predict pressure losses in closed conduit

systems.

In order to form a dimensionless equation applicable to predicting fluid losses in

closed conduits, Darcy and Weissbrook [2] applied the technique of dimensional analysis

to the relevant variables involved with the problem. The relevant variables were

determined to be fluid velocity, V ; fluid viscosity, ju; fluid density, p ; conduit length,

I ; conduit diameter, D, and the roughness of the conduit surface, s . Written

functionally.

A? - F (V ,D ,l,s ,p ,p ) 1.4

The result of the dimensional analysis performed on these variables is the Darcy -

Weisbach equation [2],

D 2 g 1.5

5

where the variables are: H l , the head loss; I ,the pipe length; V ,the average fluid

velocity; g , the gravitation constant, and D is the pipe diameter. The one undefined

variable is f, the friction factor.

For turbulent flow, the friction factor is functionally dependent on a complex

relationship between two other dimensionless quantities, the Reynolds number (equation

1.3) and the relative roughness. Relative roughness is the ratio of the statistically

averaged heights of the roughness elements to the pipe diameter ( s / D ).

A vast amount of empirical and theoretical research was directed toward

quantifying the friction factor during the first half of this century. In 1934, Pigott

compiled the results of over 10,000 experiments from various sources and formed the

Pigott chart [3]; the chart related relative roughness and Reynolds number to the friction

factor. Unfortunately, the data were compiled in such manner that practicing engineers

could not easily extract results. Pigotfs work was followed by Nikuradse who artificially

roughened pipes with sand grains [4]. Because of the substantial difference between

sand-roughened surface and surfaces encountered in practice, Nikuradse’s results varied

substantially from those of Pigott Based on Nikuradse’s results, von Karman and

Prandtl developed theoretical analyses of pipe flow that resulted in numerical constants

for the case of hydraulically smooth surfaces in which the roughness elements are small

compared to the boundary layer thickness [5]. In 1938, Colebrook continued to fit

formulas to the empirical data [5], Unfortunately, Colebrook's equation, like Pigoffs

chart, was cumbersome and difficult for the practicing engineer to use. Also in 1938,

6

Blasius [2] offered a solution that was much more tractable but only applicable to

hydraulically smooth surfaces.

In 1944, Lewis F. Moody [5] compiled the results of previous researchers in order

to “furnish the engineer with a simple means of estimating the friction factors to be used

in computing the loss of head in clean new pipes”. Moody’s relatively simple technique

of predicting frictional losses has become a standard method used by the practicing

engineer to predict pressure losses in closed conduits.

Moody provided a chart similar to Pigotfs that correlates the two dimensionless

quantities of Reynolds number, Re, and relative roughness, s / D , with the friction factor,

f. In addition, Moody provided the engineer with charts for determining the exact value

of relative roughness and Reynolds number for a closed conduit system. Moody’s chart

(Figure 1.2) is then used to correlate these two dimensionless quantities with the friction

factor. The corresponding friction factor is then inserted into the Darcy equation (1.5)

together with the fluid velocity, hydraulic diameter and length. With these parameters,

the Darcy equation can predict losses in clean, new systems to within 5%.

However, Moody's diagram is applicable only to “clean new pipes” [5], Many old

piping systems (and those prone to microbial fouling) exhibit substantial losses that

cannot be predicted using this traditional diagram.

Microbial fouling (biofouling) is a technical term referring to the adverse effects

caused by the attachment of microorganisms to liquid-solid interfaces. These

microorganisms, i.e. bacteria, bind to a surface and each other via viscous-elastic

extracellular polymers made mainly of polysaccharides produced by the bacterial cells.

7

Together, the cells and the polymer matrix are called biofilms. It was originally thought

that pressure losses caused by biofouling could be predicted using the approach proposed

by Moody [6], However, this approach failed mainly because of the fact that, as

biofouling occurs, the liquid-solid interface is changed from a rigid surface to a visco­

elastic surface with complex topography, and the resulting losses may be significantly

greater than those predicted using the Moody diagram in conjunction with the Darcy

equation.

(VO*) FOR WATER AT 60*F (VELOCITY IN-Sfc 4 6 e Ip ep 40 eo ap

X OlAMETOt M HOMES)VALUES OF

T O iB IL E

.004

.0008 5

.0006 W

.0004

0 0 0 2

.0001

.000,05

2(10* 3 4 5 6 8 |q 72(10*) 3 4 5

REYNOLDS NUMBER R = X L ( v IN^L , D IN FT, V INggg)

Figure 1.2 The Moody Diagram [5],

8

Literature Review

McEntee did some of the earliest study into the energy losses caused by biofouling'

in 1915 [7], McEntee submerged a series of steel plates in water, allowed a biofilm to '

form on the plates, and then tested them for frictional resistance to fluid flow. He

concluded that the biofouling resulted in a 2% increase in frictional resistance each day

the biofilms were allowed to grow. In a discussion of McEntee’s.paper, Sir Archibald

Denny, a member of the Society of Naval Architects and Marine Engineers,

acknowledged a similar increase in frictional resistance. He stated, “for each day a vessel

lies in our dock the skin friction resistance increases at a rate of 1A0Zo per day and this we

have found to be true for periods as long as three months.” [7],

Other areas where the increase of frictional resistance is troublesome is in the ■

water-supply and power industries. Such systems are particularly prone to biofouling,

due to an abundance of bacteria. Observations of researchers investigating biofouled

industrial systems have been qualitative with regards to biofilm structure, and

quantitative with regard to energy losses caused by the biofilms.

In a case in 1959, biofouling affected a power conduit located in North Carolina [8].

The pipeline, put into operation in 1928, was approximately 4.75 miles long, and was

composed of both steel and concrete sections. In 1928, the conduit friction head losses

were determined to be 40.6 feet at 900 cfs. In 1945, the conduit friction head losses had

increased to 55 feet, i.e. 35% greater than the original value. In 1945, inspection of the

11-ft diameter pipeline found the steel pipe portion of the line, which was painted with

bituminous paint in 1939, to be in excellent condition. The inspection also showed no

9

structural damage to the concrete lining, however, the concrete portions were found to be

covered with “something like a dense layer of soot” [8], These rough black fouling

deposits were up to 5Zg inch thick and had an average depth of 1A inch overall. Chemical

analysis showed that these deposits were composed of 87% water and organic matter,

2.5% organic matrix and 10.5% mineral matter. It was also observed that the deposits

consisted of layers that “undoubtedly represent equilibrium conditions and are obviously

laid down at yearly intervals” [8]. The researchers assumed microorganisms (i.e.

biofilm) formed the deposits, and the goal of the research was to determine the best

method of treatment. The authors concluded that mechanical cleaning at certain intervals

was the most cost-effective solution. Although knowledge of the composition of these

deposits may be helpful in determining effective temporary solutions to the biofouling

problem in closed conduits, this knowledge does nothing to associate biofilm structure to

the resulting effect on fluid dynamics and the resulting losses.

Another notable case of this type of fouling occurred in Germany in 1950. In this

instance, a 60-cm diameter, 93-kilometer long water supply line was reduced to 55% of

its original flow capacity over three years. The loss was due to a “thin, slimy layer” [9].

The layer was characterized by a “ripple-like” surface having an average thickness of .

0.25-in. The resulting energy losses could not be explained in terms of constriction or

equivalent sand roughness common to friction factor relationships.

Regarding this particular case, Characklis pointed to the rippled surface as the cause

of the unusually high losses experienced by pipelines,, using an example of a solid surface

of similar pattern that showed high frictional resistance [10]. Brauer performed

10

experiments on form stability of asphalt-lined pipes as a function of temperature of the

flowing water [11]. Brauer observed that, at higher temperatures, the asphalt mating also

assumed a rippled surface structure, which was accompanied by an unusual increase in

frictional resistance. Brauer proposed that there is a “two phase” flow, in which shear

stress (caused by the bulk fluid flow acting on the asphalt) forced the asphalt to be

dragged along the pipe wall [11]. Although this theory would explain an increase in loss,

Characklis acknowledges that the losses are not nearly as high as the losses occurring in

some biofouled systems.

Characklis continues with a general explanation of energy loss in turbulent flows

past rigid surfaces.

“The complicated nature of turbulent motion can be visualized by assuming that fluid particles moving near the wall coalesce into lumps and travel bodily together for a certain distance. If such a lump of fluid collides with the leading part of a roughness element, the fluid changes its direction and a momentum exchange takes place. Any forced motion of the fluid particles in a direction transverse to the flow corresponds to an increase in general turbulence. This phenomena causes energy loss in flow past rigid rough surfaces” [12].

Characklis then explains a possible mechanism for increased losses in biofouled

systems when the fluid interface is changed from that of a rigid surface to a rippled slime

layer.

“If the material constituting the roughness element has a low modulus of elasticity, the force exerted on the roughness element by the fluid may be sufficient to cause a temporary deformation of the element which would result in an oscillatory motion of the pliable roughness element. A resonance phenomenon could occur from the coordinated motions of the individual roughness elements.”

11

Also in 1973, Kouwen and Unny investigated the effects of surface roughness

flexibility in open channels [13]. They applied dimensional analysis to the situation and

postulated the following relationship (1.6).

U f

N y _i

mEI h d'

^ P u O y

Ik° 1.6

Where U0 is wall friction velocity; U is the free stream velocity; h0 is the roughness height

with an undeflected pliable over-layer; 5 is the height of the absolute boundary layer

(shear layer); h is the measure of the roughness height; EI is a measure of the bending

stiffness'of the pliable elements; and m is a non-dimensional value of the aerial density of

the elements. Kouwen and Unny ultimately proposed the following empirical

correlation, in which aK and bK represent numerical constants.

mEIV

-bs/ 1.7

Minkus reported another case of this phenomenon [14]. This case involved a 42-in.

diameter cast iron pipeline that had been used to transport water 7 miles to a treatment

facility for 22 years. To increase longevity of the line, a 1A-In cement lining was added to

the pipeline wall. Capacity measurements were taken after the lining was applied and

again 2 years later. The results showed a decrease in the pipeline’s capacity from 50

million gallons per day to 44 million gallons per day. Inspection of the pipeline showed

the cement lining to be in excellent condition; “however, a microbiological and chemical

film 1/32 to 1/16 in. thick was found attached to the wall” [14].

12

Minlcus also reports about a 36-in. concrete pipeline that was tested over a 5-year

period. This pipeline showed a reduction in flow of 23%, from 35 millions gallons per

day to 27 million gallons per day [14]. Immediately after emptying the pipeline, a film

was observed, and although it appeared to be smooth, Minkus states that “it could be said

that with a little imagination that there was a rougher feeling to the deposit."

In 1980 Picologlou et al. [6], presented the results of the first formal laboratory

experiments focused on the specific problem of pressure loss in biofouled closed

conduits. Picologlou et al. ran a series of experiments involving three tubular fouling

reactors (TFR) that were fouled with biofilm while pressure loss across the reactors was

monitored. In these experiments, either velocity through the reactor, or pressure loss

across the reactor, was held constant. The experiments in. which velocity was held

constant showed increases in pressure loss, while the experiments with a constant rate of

pressure reduction showed diminished flow capacity. The authors propose and evaluate

six possible mechanisms that could contribute to the increased losses.

I . Biofilm constricting the diameter and reducing flow.

To evaluate the extent and affect of diameter constriction by the biofilm, the authors

measured biofilm thickness during their experiments. Thickness measurements were

determined by dividing the biofilm volume by the surface area covered by the

biofilm. Biofilm volume was determined by removing a sample of biofilm and then

allowing excess fluid to drain from the sample. The authors acknowledge that the

thickness measurements could be as much as 36% larger for drain times of 2.5

minutes than for drain times of 10 minutes. They also found that the constriction of

13

the hydraulic diameter caused by the biofilm would account for only 10% of the total

frictional resistance.

2. Change in fluid viscosity.

The authors found that fluid viscosity never varied by more than 2% from pure

water; therefore, the authors dismissed the effect that fluctuating viscosity had on

pressure loss as negligible.

3. Viscous dissipation within the biofilm due to its creeping flow in the down stream

direction.

The authors refer to the work done by Bauer [11] regarding the asphalt-lined

conduits, but dismiss this creeping phenomenon as a possible mechanism for the

increase loss based on two reasons. First, the biofilm coverage always appeared

uniform, and second, there was no evidence of an accumulation of biofilm in areas

where creeping biofilm would collect.

4. Viscous dissipation within the biofilm due to its oscillatory response to turbulent

flow excitation.

Previous research by the authors determined the viscoelastic nature of the biofilm.

Using a Weissenberg Rheogoniometer, the biofilm was determined to have a

relatively large viscous modulus, i.e., the viscous modulus was much larger than the

elastic modulus. Picologlou et al. state that, “the possibility exists that the biofilm

draws energy from the flow, such energy being eventually dissipated through viscous

action. This situation is quite complex and defies analysis, particularly since there is

a nonlinear coupling between the structure of the turbulent flow and the biofilm

14

response.” The authors also believe that this loss mechanism is of secondary

importance because the losses they observed could quite satisfactorily be attributed to

increased film surface roughness.

5. Increased dissipation in fluid due to increased surface roughness as a result of

biofilm accumulation.

Although the frictional resistance observed could usually be adequately explained

using methods for rigid rough surfaces, the authors do not conclude that the biofilm

presents a rigid rough surface to the flow. This would be an oversimplification and

would not account for all their experimental observations. Specifically, in one

experiment the pronounced frictional resistance could not be adequately explained as

a rigid surface element.

6. Increased dissipation in fluid due to the presence of biofilm filaments.

Filaments of the biofilm were observed to flutter with a frequency related to the

bulk fluid velocity. The authors also qualitatively observed that the frictional

resistance increased with increasing filament length, although they do not define what

is meant by “streamer length”. They suggest that this loss mechanism is analogous to

increased drag in streams due to bottom vegetation, and similar phenomena occurring

in atmospheric boundary layers in the presence of natural vegetation.

The authors concluded that both frictional resistance and equivalent sand roughness

values correspond to an increase in biofilm thickness. They characterized the frictional

losses as having a lag period associated with small biofilm thickness, followed by a rapid

increase when the biofilm thickness reached a critical thickness. The authors

15

hypothesized that this may be because the biofilm thickness reaches a critical value in

relation to the viscous sublayer thickness. The response of the friction factor for a tube

with attached biofilm is similar to that of a rigid rough surface for Reynolds numbers

ranging from 5,000-48,000. They also observed a filamentous morphology of the biofilm

surface and stated that these filaments contribute to the increase in frictional resistance.

Many researchers have suggested that biofilms have an effect similar to that of a

compliant surface [6], [10], [15], [16], [17-21], Most research done in this area suggests

compliant surfaces reduce skin friction, in some cases, significantly. These findings tend

to contradict applying such a concept to biofilms. Initial research in this area was done

by Kramer, who covered underwater projectiles with rubber diaphragm [19]. He found

drag could be reduced up to 40% from that of an equivalent rigid-surface projectile at

Reynolds numbers of 15x106. Looney and Blick achieved reductions in skin friction up

to 50% using a compliant plane [20]. Pelt found reductions of 35% in friction losses for

flexible tubes lined with a variety of viscous fluids [21]. The findings of all of this

research suggest that to maximize reduction in drag, a highly viscous fluid enclosed by

the thinnest membrane should be used.

Research by Klinzing et al. [22] on frictional losses in foam-damped flexible tubes

yielded interesting but inconclusive results. They tested flow through a 15/16-in inner

diameter, 1/16-in wall silastic tube. This tube was encased by polyurethane foam, and

was hydraulically smooth. Although no pronounced decrease in frictional resistance was

found, the authors do show a definite decrease in the friction factor for the foam-damped

tube between Reynolds numbers of 10,000 and 20,000. They interpret this result as a

16

possible delay in the onset of full turbulence from the normal transition range of

Reynolds numbers between 2100 - 7000 to the 10,000 - 20,000 range.

Loeb et al. [17] studied the effects of microbial fouling films on the hydrodynamic

drag of rotating disks. This experimental study was undertaken to “evaluate the effects of

■ microbial slime under hydrodynamic conditions that reflect realistic ranges of vessel

operation”. Their work indicated that even under relatively high Reynolds numbers,

representing vessel speeds of 20 to 60 knots, the biofilms increased drag by up to 10%. It

should be noted that they tested surfaces that were either initially hydraulically smooth or

rough. Unlike Picologlou et al. [6], Loeb et al. [17] did not find an initial reduction in

drag on the hydraulically rough discs in the early stages of fouling. •

Lewkowicz and Das [23] investigated turbulent boundary layers on rough surfaces

with and without a pliable over-layer that simulated marine biofouling. They covered

two flat plates (30 cm by 92 cm) with abrasive paper and mounted tufts of fine nylon

fibers to one a plate to form a “combined roughness.” Each tuft contained approximately

300 fibers that were 2 cm long and 15 microns in diameter (the modulus of elasticity for

the nylon is approximately 2 x IO9 N/m2). The tufts were laid out on the plate at a rate of

3100 per square meter. The plate was then placed in a wind tunnel and exposed to a free

stream of air at 26 m/s. By utilizing a flexible top wall in the working section of the

tunnel, the investigators were able to keep the static pressure distribution along the plate

constant to within 2% of the inlet dynamic pressure. “Not surprisingly, the combined

roughness had a thickening effect on the boundary layer (by some 25-30%). Notably, it

affected, in that sense, the displacement thickness more than the momentum thickness as

17

the shape factor, 5* /6 ,fox the combined roughness increased on average by 30%.”

Here <5* is the displacement thickness, and O is the momentum thickness. They also

found the skin friction coefficient, Cf, was 18% higher on the plate with the combined

roughness.

Lewandowski and Stoodley [15] also observed an increase in pressure loss in

conduits fouled with biofilm. They postulated that “structural development of the biofilm

suggests that individual microcolonies behave like blunt bodies shedding vortices. The

microcolonies assume elongated forms, termed “streamers”, possibly because of an

external pressure drag force.” The vortex sheet formed around the blunt colonies

activated the streamers into motion. Energy then dissipated through the flow-induced

movements of the streamer and microcolonies. This would, in part, explain the increase

in pressure loss.

In addition, Lewandowski and Stoodley also suggest that the biofilm influences

pressure loss only above a certain critical flow velocity [15]. This pressure loss,

attributed to the biofilm, reaches a steady (or pseudo-steady) state in the reactor. The

authors conclude that the “interpretation of classical hydrodynamic parameters such as

Reynolds number, friction factor, and surface roughness as related to biofilms should be

reexamined in context to biofilm viscoelasticity and heterogeneity.” The authors also

suggest that, although the Reynolds number calculated using the reactor geometry may be

useful for predicting the overall flow stability, it should be re-evaluated to asses local

flow conditions near a biofihn.

18

To verify their hypothesis concerning the vorticies, the. investigators attempted to

relate the frequencies of streamer motion to the Strouhal number [16]. Using confocal

scanning laser microscopy, they were able to plot the position of a location on a streamer

filament as a function of time. Unfortunately, no characteristic frequencies were found,

suggesting that the streamers’ motion is more likely a response to turbulent flow rather

than to vortices formed around the blunt colonies.' . -

Recently, in 1999, Schultz and Swain investigated the effect of biofilms on

turbulent boundary layers [24]. Their experiments were conducted in a water tunnel

utilizing actual biofilms. Biofilms were grown on steel plates, 2.06-m by 0.58-m (54 mm

thick), in filtered lagoon water. The control plates were non-fouled. The biofilms were

allowed to grow for 2 to 3 weeks, and thickness measurements (performed using a wet-

film paint thickness gauge) ranged from 25 microns to 2032 microns. A fiber-optic laser

Doppler velocimeter (LDV) was used to acquire velocity measurements.

Results of the investigation showed that there was no statistically significant

difference in the turbulent boundary layer (absolute) thickness between the control and

the fouled specimens. However, statistically significant increases in the displacement

thickness and the shape factor were found for the biofouled plates. The authors also

found that the skin friction was dependent on biofilm thickness, composition, and

morphology. For example, a biofilm thickness of 160 microns increased frictional

resistance by 33%, while a thickness of 350 microns increased the resistance by 68%. In

addition, the authors state that biofilms containing a higher proportion of algae seem to

19

“draw a greater amount of momentum from the mean flow” because of “waving algae

filaments”.

The authors conclude that there is not a sufficient characteristic length scale.

associated with the complex biofilm structure to relate it to traditional methods, such as a

standard Nikuradse sand roughness.

Clearly, the problem of increased frictional resistance caused by biofilm has been

the center of a considerable amount of research. Although documentation of the problem

is extensive, there does not currently exist an accepted mechanism relating the biofouling

to the increased losses. The following research was done to provide a clearer

understanding of the relationship between this type of biofouling and increases in

frictional resistance.

APPROACHES TO PREDICTING VELOCITY PROFILES IN HETEROGENEOUS BIOFILMS.

Introduction

To model flow velocities within a biofilm, one must obtain some knowledge of

biofilm structure. There are currently three models of biofilm structure explained in

literature on biofilms grown under low flow velocities.

1. Dense, or slab, biofilm structure, in which there are no pores or voids [25].

2. Heterogeneous biofilm structure, in which microcolonies form mushroom like

structures [26] [27] [28] [29].

3. Heterogeneous mosaic biofilm structure, in which individual microcolonies

form stacks which are separated from other colonies [30].

In the dense or non-porous films, there is no convective flow to model. However,

in the heterogeneous and heterogeneous mosaic structure models, convective transport is

important. Evaluations of three possible approaches to modeling flow within these types

of films are offered in this chapter.

The first evaluation compares the biofilm’s velocity profile to actual profiles

obtained for various vegetative canopies [31]. The second comparison is based on

modeling the film structure as a highly idealized canopy composed of spherical elements

21

[32]. A third possibility of a canopy composed of idealized cylinders is then offered and

evaluated.

Velocity Profile of an Actual Biofilm

In order to evaluate the effectiveness of these approaches, an actual velocity profile

is needed for comparison. The profile that will be used for this purpose was generated at

Montana State University’s Center for Biofilm Engineering using an electrochemical

technique [33]. This technique utilizes the measuring of the limiting current of a

microelectrode. The limiting current is a function of the mass transfer boundary layer,

which is in turn a function of the local flow velocity. The micro-electrode was calibrated

using a particle tracking technique in conjunction with confocal scanning laser

microscopy. The microelectrode was then positioned in and near a heterogenous biofilm.

Xia et al. [33] determined velocity profiles at five points (see Figure 2.1) in a three

species biofilm (Psuedamonas aeruginosa. Pseudomonas flour escens and Klebsiella

pneumoniae), which was, on average, 160 pm thick. Four of the five profiles were taken

at voids in the biofilm structure, while the fifth (point E) was taken within and above a

cell cluster. In Figure 2.1, lighter areas denote voids and darker areas represent biofilm

clusters. The velocity profile at point A was selected as the profile to be used for

comparison because of its location in a void and the corresponding flow profile, shown in

Figure 2.2.

Vel

ocit

y (m

/s)

2 2

Figure 2.1 Image of 3-Specie Biofilm, Showing the Locations at which Velocity Profiles were Measured.

V e r t i c l e P o s i t i o n ( g m )

Figure 2.2 Plot of Flow Velocity as a Function of Vertical Position for the Five Locations Shown in Figure 2.1 [33].

23

Comparison of flow velocity within a biofilm to atmospheric flow in a vegetative canopy

Research and studies into flow profiles in natural canopies is abundant. Empirical

data is available for flow profiles from plant types ranging from immature com to large

trees [31]. Such data has been incorporated into the modeling of shear stresses on

vegetation caused by wind gusts, and into complicated models used to predict the

behavior of wildfires. The intent here is to evaluate whether such canopy profiles could

be used to model flow within a porous biofilm.

Arguments can be made both for and against modeling flow within biofilms as flow

within natural canopies. An argument for such a model is that certain components of the

systems are somewhat analogous: flow in both systems can be considered

incompressible, and both canopies are composed of flexible elements. Arguments can

also be made that both vegetation and bibfilms develop structures that optimize nutrient

uptake. For a vegetative system this includes such parameters as sunlight uptake by leaf

area (photosynthesis) and transport of water from the soil throughout the plant. For a

biofilm these processes include breaking down substrates through diffusion and cell

respiration.

At the same time obvious differences exist between the two systems. For instance,

the volume ratios of the two systems vary significantly. For vegetation the ratio of

vegetative volume to total volume is of the order IO'3, while volume ratios of biofilms

with their extra-cellular matrix tend to be of the order IO"1. It follows from this

volumetric relationship that the two systems will also have dramatically different surface

24

area ratios. Differences in these ratios will result in different shear and pressure drag

forces exerted on the fluid in each system.

For vegetative canopies, Cionco [31] proposed the following equation for flow

profiles within vegetative canopies,

~~—11u = Ufl * e 2.1

where u is the velocity of the fluid at a given height (z) within the total canopy height

(H). Uh is the fluid velocity at the canopy top, and “a” is an attenuation coefficient that

varies based on the type of vegetation. In order to illustrate the resulting velocity profile.

Figure 2.3 is provided for a canopy with an attenuation coefficient value, a, of two.

10.8

x 0.60.4 0.2

00 0.2 0.4 0.6 0.8 1

u/Uh

Figure 2.3 Flow Profile Generated by Equation 2.1 with an Attenuation Coefficient of 2.

Equation 2.1 is empirically based, and it was developed by fitting curves to data

such as that offered in Figure 2.4. Figure 2.4 shows velocity as a function of vertical

position inside a canopy for several types of vegetation [31]. From this graph one can

observe the similar shape of the flow profiles typical in all vegetation. Also, one should

note that this graph shows the resolution of the flow profiles into the bulk fluid flow up to

25

a height twice that of the canopy. The reader should note that equation 2.1 is only

applicable to flow within the canopy, and Uh is not the bulk fluid velocity but rather the

fluid velocity at the canopy top.

A WIND-PROFILE INDEX FOR CANOPY FLOW

CURVE — SYMBOL— ELEMENTS

1 o CORN2 » WHEAT3 + CHRISTM AS TREES4 * BUSHEL BASKETS5 * FL E X IB L E CANOPY6 * PEG CANOPY

--- 2H---

-H------- (Top o f Canopy)-

.6 0 .7 0 .8 0 .9 0 IjO

Figure 2.4 Nondimensional Wind Profiles for Various Canopies [31]

In order to compare the velocity profiles of vegetative canopies to the velocity

profiles of biofilms, a graphical technique was employed. First, the flow profile at point

A (see Figure 2.1) was made dimensionless in the same manner as is typically done with

vegetative canopies. Next, this data was plotted using a logarithmic x-axis (as in Figure

2.4). Finally, this graph was scaled to the same size as Figure 2.4, and the results of the

biofilm profile were then superimposed on Figure 2.4. The result is shown in Figure 2.5.

2 6

CURVE — SYMBOL— ELEMENTS

CORNWHEATCHRISTMAS TR EES BUSHEL BASKETS FLEX IB LE CANOPY PEG CANOPY

— — 2H --------

Biofilm- ----------H ------- (Top o f Canopy)-

.6 0 .7 0 .8 0 .9 0 1.0

U2H

Figure 2.5 Superimposing of Biofilm Flow Profile on Figure 2.4

Qualitatively, these results are quite good, with the biofilm flow profile appearing

to be very similar to a vegetative canopy composed of corn. It follows that for

quantitative purposes, the attenuation coefficient of 1.97 (Table 2.1) for corn can be used

with equation 2.1 to provide estimates of flow velocities for modeling purposes.

Table 2.1 Attenuation Coefficients for various canopies [31]

Canopy a valueOats 2.80Wheat 2.45Plastic Strips 1.67Rice 1.62Sunflower 1.32Xmas Trees 1.06Larch Trees 1.00Wooden Pegs 0.79Citrus Orchard 0.44Bushel Baskets 0.36Com 1.97Immature Com 2.82Sam pled Biofilm 1.97

27

It should be noted that although the fit for the biofilm profile is satisfying visually,

attempts to numerically fit data to equation 2.1 result in a poor correlation. This results

from the difference in the basic shape of the two curves, i.e. the velocity within the

vegetative canopies are concave up while the biofilm’s velocity profile is concave down

(see Figure 2.5).

After these considerations, three conclusions are made for comparing flow within a

biofilm to flow within a vegetative canopy:

1. This approach can be used satisfactorily to estimate the flow velocity within

and above a porous biofilm.

2. Although qualitative comparison is satisfactory, there exists a difference in

the shape of the two profiles, which makes quantitative comparison difficult.

3. Use of equation 2.1, with an attenuation coefficient of 1.97, can provide

approximate flow velocities within, biofilms modeled as heterogeneous and

heterogeneous mosaic films.

2 8

Comparison to Flow within an Artificial Canopy

The previous study compared flow within biofilm systems to flow within a

vegetative canopy. This is primarily an empirical approach to simplify the quantifying of

complicated systems; nevertheless, it does provide researchers with a simple approach for

estimating shear within biofilms.

Now the relevance of an artificial canopy, offered by Stolzenbach, will be

investigated [32]. Stolzenbach proposes a canopy similar to that illustrated in Figure 2.6.

Just as with the vegetative canopy, there exist arguments for and against this approach.

Volume and surface ratio values for this model tend to be much closer to that of a biofilm

than ratio values for a vegetative canopy. In addition, the simplicity of the approach and

the existence of an analytical solution (if the proper assumptions are made) are appealing,

but one can argue that this approach is greatly oversimplifying a very complicated natural

system.

Bulk fluid

Figure 2.6 Spherical Canopy Illustration

29

Stolzenbach proposes a biofilm model of a canopy composed of spherical elements

ranging from radii of amj„ to amax, with size distribution following the probability density

function. Even this apparently simple approach needs the following assumptions in order

to make the problem tractable. First, Stokes flow is assumed, i.e. Reynolds numbers are

low enough that there is no separation of flow. Second, there are no perturbations in the

flow, and each fluid element is affected by only one sphere at a time. Finally, one­

dimensional, incompressible flow is assumed. Following these assumptions, Stolzenbach

applies the Navier Stokes equations (2.2).

where u; are the velocity vectors, p is the density, and xj and Xj are the principle

coordinate directions. This equation is greatly simplified to equation 2.3 when the

previous assumptions (involving I-dimensional incompressible Stokes flow) are

evaluated and Cartesian notation is used. Note that if the body force is neglected the

equation becomes that of the shear-driven Couette flow.

Next the body force, or drag, caused by the spheres and acting on the fluid is

substituted by evaluating the following integral that represents the total drag force acting

on the fluid.

2.2

2.3

f x = 6* ft* pi* as * n(as )da = - Equation 2.4

30

WherertO1) is the particle size distribution function and a is the volume fraction

occupied by the spheres. This term is then substituted into the simplified Navier Stokes

equation, resulting in equation 2.5.

A S2W 9 ju*U *a0 = + ------ 2.5

To evaluate this equation, the no-slip boundary condition at the substratum is

applied along with the boundary condition of shear-driven flow at the canopy surface.

The results are illustrated in Figure 2.7.

m a x

a = volume fraction a c = sphere radius / canopy Height 5 = Canopy Height (Lf) z = y

Figure 2.7 Flow Profile Generated Using the Canopy Model Composed of Spherical Elements and Associated Variables from Stolzenbach [32].

In order to compare the results of this data with the biofilm, values for the

parameters of a, the volume fraction occupied by the spheres, and a c, the ratio of the

sphere radius to the canopy height, were needed. These values were determined from

31

measurements taken from Figyre 2.1. The value for the sphere radius (200 microns) was

taken across the colony where point E is located in Figure 2.1. The thickness of the film

5, was reported to be 160 microns [33]. The value for a, the volume fraction of the film

was estimated to be 0.375. This number was determined by averaging the volume

fraction at the substratum with the volume fraction at the biofilm surface. At the

substratum the volume ratio was estimated to be 0.75. This value is based on the ratio of

the area shown in Figure 2.1 that is covered with biofilm compared to the total area. At

the biofilm surface, the porosity is zero. The average of these, 0.375, was used for the

calculation of the volume fraction, a. To estimate oic (the ratio of the sphere radius

squared to the biofilm thickness squared) the values for the radius of the colony and

biofilm thickness were used. Using these values, a/ac is 0.24, and the estimated profile

shape for this value can be approximated as 0.25. The corresponding curve can be

viewed in Figure 2.7.

Stolzenbach1 s model compares favorably to the shape of the actual flow profile

generated from the sample data taken from the three species biofilm, and this comparison

can be seen in the 2 graphs in Figure 2.8. Here the data for the actual biofilm was

normalized in the same fashion as Stolzenbach1 s predicted profiles displayed in Figure

2.7. It also interesting to note that flow within the biofilm at point A (Figure 2.1) does

not seem to be greatly affected by the presence of the film.

When the values, which were estimated from the biofilm image, are used to

determine a /ac, the results of this evaluation are favorable. But it needs to be

acknowledged that the estimated value of 200 microns for the value of the sphere radii.

32

obtained in from the biofilm image, is greater than the thickness of the film (160

microns). This radii value is somewhat greater than originally intended by Stolzenbach.

However, it is interesting to note that although this scale differs, the model still appears

promising. This may be in part because the biofilm volume fraction, a, is actually low.

Here it was estimated to be 0.375, meaning that 37.5% of the canopy is actual biofilm. A

low volume fraction is reflected in a low value for a/ac. It would follow that the biofilm

would not greatly change the velocity profile from that of shear-driven Couette flow,

which is indeed what we observe in the actual velocity profile in Figure 2.8.

Estimated velocity profile for biofilm using Stolzenbaclfs model

max

Figure 2.8 Plot of Estimated Biofilm Profile using Stolzenbach's Model with Parameter Sizes Estimated from Figure 2.1 and the Actual Flow Profile Obtained at Point A in Figure 2.1

In conclusion, both the natural and artificial canopy models have inadequacies, but

both yield results that may be easily incorporated into the modeling of biofilms and their

associated shear stresses.

33

Based on the promising results obtained for this model and acknowledging its

insufficiencies, a more appropriate model consisting of cylinders rather than spheres is

offered and evaluated.

Evaluation of Flow Through a Model Biofilm Composed of Cylindrical Elements.

It is now proposed to model heterogeneous biofilms as a canopy model composed

of cylindrical elements protruding from the substratum as illustrated in Figure 2.9. This

model is more applicable to biofilm modeling than the spherical model proposed by

Stolzenbach [32] for the following reasons.

Unlike the canopy composed of spherical elements that are held in place by

assumption, this model’s elements protrude from the substratum completely to the bulk

fluid, and they more closely represent the actual colonies in a heterogeneous biofilm.

Figure 2.9. Proposed Cylinder Model

One can vary the size of the cylinder radii in this model to account for changes of

the biofilm porosity within the canopy height. One could even allow these cylinders to

34

form mushroom or mosaic type shapes or clusters, which some researchers argue are

common among heterogeneous biofilms [26] [27]. However, when the protruding

structures begin to approach this scale, it is not valid to assume that flow around

neighboring protrusions does not interact.

In order to determine the flow profile, a method similar to that of Stolzenbach's is

employed, and the assumptions involving Stokes flow (low Reynolds numbers) and the

independent effect of each cylinder on the flow must be made. To determine the drag

resulting from the cylinders on the flow, a linear relationship between Reynolds number

and drag coefficient is assumed to approximate the drag force acting on the fluid. This is

similar to that of Stokes flow (Figure 2.10).

Figure 2.10 Drag Coefficient as a Function of Reynolds Number for a Cylinder [2]

If the protrusions are considered to have constant radii, the resulting Navier Stokes

equations are once again reduced to the form of equation 2.3. Next, the equation for drag

caused by cylinders is substituted for fx

35

2.6Z

where p is fluid density, U is velocity as a function of y, A is the cross-section area of an

individual cylinder, Cd is the coefficient of drag (approximated as 12/Re) and N is the

number of cylinders per area. Substituting equation 2.6 into equation 2.3 and simplifying

leads to equation 2.7

d2U9 /

with

2.7

B = 6*vf*iVD

2.8

Here A and N ate previously defined, and D is the cylinder diameter. The solution to

equation 2.7, with the no slip-boundary condition at the substratum, and U=I at y=l,, was

determined analytically and is equation 2.9.

U I * e ~ B * y _|____ IeB —e B

B * y 2.9

The results of equation 2.9 for various values of B are displayed graphically in

Figure 2.11. From these graphs, a profile similar to that provided by Stolzenbach's

model is seen, and the results here compare well with the results for the actual biofilm .

profile (Figure 2.11). Conclusions for comparison of this model is that it appears to have

the all the benefits of Stolzenbach5 s yet it is more realistic for the following reasons.

36

I . The actual biofilm structure reviewed here is arguably more similar to cylinders

than to the spheres that Stolzenbach proposes. Stolzenbach’s sphere model treats the

biofilm more as a homogeneous medium rather than the heterogeneous structure

determined for the actual biofilm. Although the results of the comparison made to the

actual biofilm were favorable, perhaps Stoltzenbach’s model is applicable to biofilms

exhibiting more of a homogeneous and porous structure.

Plot for Actual Biofilm Plot for B=1

Plot for B=2 Plot for B=S

1

0.8

y 0.6 H 0.4

0.2

0

* *

D 0.2 0.4 ^ 0.6 0.8

1

0.8

y 0.6

H 0.4

0.2

0

r"*.■

/ Sill:II I l0.4 ^ 0.6 0.8

u H u H

Figure 2.11 Plot of Biofilm Flow Velocity Profile Resuting from Equation 2.9 for Various Values of B.

37

2. The cylinder model could be further refined to more closely imitate actual

biofilm structure based on the changes in porosity at different heights within the biofilm.

For instance, to mimic the blunt colonies observed in many biofilms, the cylinders’ radii

could be assumed large at the base and then taper down away from the substratum. The

colonies could also be assumed to have small radii at the substratum and larger radii

within the canopy height. Such alterations would imitate the proposed mushroom model

[35].

Conclusion

Each of these methods can be used to estimate velocity profiles in biofilms that

have a highly porous heterogeneous structure. Unfortunately, this type of biofilm

structure is usually associated with biofilms grown under low flow velocities below those

relevant to industrial piping applications. Such industrial systems typically operate at

Reynolds numbers over 10,000 and fluid velocities on .the scale of meters per second

rather than centimeters per second. Therefore, a dramatically different approach is

needed to understand the problem of increased frictional losses in industrial biofouled

systems.

38

MATERIALS AND METHODS

. Introduction

In order to gain insight into the Mctional losses exhibited by biofouled systems,

several experiments were run. The purpose and results of the specific experiments are

included and discussed in Chapters 4 through 7. Although the specific objective of each

experiment varied, all the results were obtained from experiments that shared a similar

system, which consisted of a recycle loop containing a closed conduit reactor.

Throughout this reactor, biofilm growth and the corresponding pressure loss were

monitored. Rather than defining this system and its components in each of the following

chapters, the general system and method will be defined in this chapter and the details

specific to each experiment, which vary from the general layout, are included in the '

following chapters.

System Layout

The general experimental layout used in these studies is illustrated in Figure 3.1,

and an actual system is shown in Figure 3.2. The general system is composed of a

recycle loop in which the biofilm was grown. The recycle loop consists of a mixing

chamber, tubing, pump, the reactor, and meters to monitor flow and pressure loss across

39

the reactor. The media, filtered air, and dilution water were added to the mixing chamber

that was vented to the atmosphere of the room. Flow through the recycle loop was

provided by either, a Cole Parmer pump (Model 7553-70) with an impeller pump head

attachment, or a Little Giant (Model 4-MD-HC) impeller pump (Figure 3.3). A pressure

vessel served as a pulse damper was placed after the pump in order to dampen flow

surges that may have been created by the pump. Down stream of the pulse damper, flow

MEDIA

Figure 3.1 The General System Layout used for the Experiments of Chapters 4 through 7

40

entered the reactor. Closed channel reactors were used for the experiments. Two ports

were placed across each reactor. The first port was positioned a minimum of 20

hydraulic diameters from the entrance in order to ensure fully developed flow. A second

port was placed at the end of each reactor. Cole Parmer pressure transmitters (Model

07354-05 and Model 07354-07) monitored pressure loss across the reactors.

Figure 3.2 Photo of an Actual Closed Conduit System used to Monitor Energy Losses Caused by Biofouling.

41

Growth Medium and Sterilization

The growth medium was the same for all experiments. The medium was diluted at

a twenty to one ratio with filtered tap water to a final concentration of 183 ppm

Na2HPO4, 35 ppm KH2P04, 19 ppm (NH4)2SO4, and 1.9 ppm MgSO4 * 7H20.

The reactors were sterilized by circulating a 5% bleach solution throughout for at

least 12 hours. The system was then flushed with filtered water until the pH returned to 7

and filled with media and dilution water. For the experiments involving 3 species, the

media and associated tubing were autoclaved for 2 hours at 121C in order to prevent

contamination. The media and dilution water were pumped to the system at a rate to

ensure a retention time of 20 to 30 minutes for all experiments.

Figure 3.3 An Impeller Pump also used to provide Flow through the Recycle Loop and Closed Conduit Reactor

42

Bidfouling Method

The systems were inoculated with either a mixture of three species of bacteria

{Pseudomonas aeruginosa (ATCC 700829), Pseudamonas fluorescense (ATCC 700830),

and Klebsiellapneumoniea (ATCC 700831)) or activated sludge. For the experiments

involving the three species of bacteria, the media and associated tubing were autoclaved

for 2 hours at 121C in order to prevent contamination.

All systems once inoculated were run as a batch culture for 24 to 48 hours to ensure

microbial attachment. The systems were then switched to a continuous flow in order to

wash out all suspended microorganisms.

Growth media was delivered to a mixing chamber by peristaltic pumps and

recirculated within the system. Filtered air was bubbled into the mixing chamber to

provide oxygen. The growth media was diluted at a twenty to one ratio with filtered tap

water to a final concentration of 183 ppm Na2HPO4, 3.5 ppm KH2P04, 19 ppm

(NH4)2SO4, 1.9 ppm MgSO4 x 7H20, 40 ppm glucose and 10 ppm yeast extract. This

media and dilution water were pumped to the system at a rate to ensure a retention time

of 20 to 30 minutes for all experiments.

43

Measurement of Pressure Loss and Calculation of Friction Factors

Volumetric flow rates through the reactor were held constant throughout every

experiment and monitored using either a McMillan Co 101 Flo-Sen (Model 6593) or a

Great Plains Industries electronic flow meter (Model A104GMN025NA1) (Figures 3.4

and 3.5). Pressure drop was recorded every 24 hours for each reactor. Figure 3.6 shows

graphically the data recorded from two of the experimental runs. From this graph, it is

also possible to see that the maximum frictional losses for these two experiments occur at

about 400 hours. Pressure across the reactors was monitored with Cole-Parmer pressure

transmitters (Model 07354-05 and Model 07354-07) (Figure 3.7) which were powered by

an Emco 24 Volt Power Supply (Figure 3.8). The transducers were calibrated using

water and a u-tube manometer. Friction factors were calculated by inserting the

parameters, V , fluid velocity, I , reactor length, and Ajy, the pressure loss, into the

Darcy-Weisbach equation (1.5) and solving for the friction factor.

Figure 3.4 Cole Parmer Flowmeter

44

Figure 3.5 GPI Flowmeter

Hours

Figure 3.6 Plots of Pressure Loss vs. Time for Two Experiments involving an 8 mm Square Glass Tube and the Three Species Bacteria.

45

Figure 3.7 A Cole Parmer Pressure Transducer

Figure 3.8 Emco 24 Volt Power Supply

46

CLOSED CONDUIT SYSTEM HEADLOSS AS A FUNCTION OF BIOFILM STRUCTURE

Introduction

Several researchers have suggested a connection between increased frictional losses

and filamentous biofilm morphology. Picologou et al. reported that "the filaments of the

biofilm flutter with a frequency that is a function of the average fluid velocity" and that

"frictional resistance increased with increasing filament length" [6], Lewkowicz and Das

attempted to simulate the filaments by adhering groups of fine nylon tufts to flat,,

hydraulically rough, plates that were then placed in a wind tunnel [23]. The plates

containing the filaments showed an 18% increase in frictional resistance over those

without the filaments. Schultz and Swain, who researched the effect of biofilms on

turbulent boundary layers using laser Doppler velocimetry, stated that "Waving algae

filaments seem to draw a greater amount of momentum from the flow than do slime films

alone" [24]. Stoodley and Lewandowski tracked the movement of a filament but were

unable to correlate the movement to any related frequency [15] [16]. Although all of

these observations do provide some insight into the problem of increased frictional

resistance associated with biofouling, they do not quantitatively relate biofilm structure to

energy losses associated with biofouling. The goal of this section, is to quantify a

relationship between biofilm filaments and increased frictional losses. To accomplish

47

this goal, two experiments were run. The first experiment was a porosity study, and is

explained in section 4.2. This experiment was performed in order to determine to what

extent the biofilm base acts as a constriction to a conduit diameter, i.e. while a highly

porous biofilm would have convective flow within it; a non-porous film would not. The

second was performed in order quantify a relationship between the filamentous biofilm

surface to the energy losses of the system.

Porosity Study

a) Porosity Study Materials and Method

The reactor system used for the porosity study was similar to the general

system layout of Figure 3.1. Because the method used to determine porosity

necessitated a reactor with favorable optical properties, a square borosilicate glass

tube (0.8 cm x 0.8 cm and 121 cm length, Freidrick & Douglas BST 800-80) was

used. Total system capacity was 450 ml. The system was inoculated with the 3

species of bacteria and run as a continuous culture with the medium described in

section 3.4. The three species biofilm was grown under a flow velocity of 0.81 m/s

and a Reynolds number of 6000. Porosity measurements were obtained by the

following method of biofilm imaging and image analysis.

48

b) Biofilm Imaging and Image Analysis

After 3 weeks of monitoring pressure loss and biofilm growth, I % v/v,

acridine orange (from Sigma) was added to the reactor to stain the three species

biofilm. The solution was recycled for 10 minutes to complete the staining

Following staining, confocal microscopy images were taken at 20 micron intervals

beginning at the biofilm/glass interface and then at every 20 microns through the

200 micron thick film and up to biofilm/fluid interface. The digital images of the

biofilm were takenin binary format. In these binary images, white (zero) is the area

void of biofilm while black (one) represents areas covered with biomass. The area

porosity is defined as the ratio of void area to total area. This calculation was

performed using the software Imagepro® (Media Cybernetics, Maryland).

c) Results of the Porosity Study

The graph shown in Figure 4.1 illustrates that the three species biofilm, grown

under initial non-fouled mean flow velocity of 0.81 m/s, had a dense base. This

observation agrees well with a close-up of the cross-sectioned image taken from the

second experiment that involved similar growth conditions and the vinyl tube

reactor (Figure 4.2). An isotropic view of a cross-section taken from this

experiment is shown in Figure 4.3. Because of the dense base and the relatively

small reactor sizes, for the remaining calculations and graphs, the biofilm base, as

shown in Figure 4.4, has been considered a constriction to the conduit. The

49

increased losses from greater dynamic head have been removed when determining

friction factors.

> 0.82 0.611 0.4

Base

0 50 100 150 200 250

Distance from substratum (microns)Figure 4.1 Porosity Profile Graph Showing Low Porosity at the Biofilm Base.

This finding of a low-porous base proves to be somewhat different from the

results for many films grown under open channel, laminar flow. Many biofilms

grown under low shear laminar flow conditions tend to have high porosity values

near the substratum. This high porosity suggests that hydrodynamics in laminar

films are similar to those associated with porous mediums and may be modeled

according to the methods suggested in Chapter 2. Unfortunately, such models are

ineffective when applied to the structure of biofilms grown under turbulent flow, in

which there is a definite structure difference between the surface film [35], and the

non-porous base film.

50

Figure 4.2 Close-Up Image of Substratum and Biofilm with Filaments Protruding Toward the Bulk Fluid

51

Experiment to Quantify a Relationship between Biofilm Structure and Energy Loss

a) Materials and Methods

In order to relate the biofilm structure to energy losses in closed conduit

systems, the following experiment was performed. Again, the general system

layout was that shown in Figure 3.1. For this experiment, vinyl tubing (1.27-cm

OD x 0.635-cm ID and 50 cm length) was used as the reactor. Pressure loss

readings were taken from two ports. The ports were made of "t" couplings with an

inner diameter of 1/4 inch. The ports were 0.5 meters apart. The sampling section,

from which the 2-mm cross-sections were taken, was located downstream of the

pressure loss reactor. Pressure loss readings were taken prior to removing samples.

Ten samples were taken over 28 days. To remove the samples, the system flow was

temporarily stopped. The cross section samples (Figure 4.3 and Figure 4.4) were

then sliced from the system using a Gem single edge industrial blade (No. 940161).

After being removed from the system, the samples were placed in a plastic Petri

dish and submerged in distilled water. If the sample were not submerged, the .

filaments would adhere to the base structure and could not be observed. Such

behavior is illustrated in Figure 4.5 and Figure 4.6. The submerged samples were

then imaged. Images were captured by a COHO® closed circuit camera (Model

2222-1040/0000) and Flashpoint® frame grabber (Integral technologies Inc.)

connected to a computer. From these images (Figure 4.4 and Appendix A)

measurements concerning base thickness and streamer lengths were obtained using

the computer program Imagepro™.

52

Figure 4.3 Typical Sample of 14” I D. Vinyl Tubing from Second Experiment.

Figure 4.4 Image of Sample Showing the Traces used to Measure the Filament Lengths and Base Thickness.

Fric

tion

Fact

or

53

Figure 4.5 Non-Submerged Sample with Figure 4.6 Submerged Sample with Filaments Adhered to the Base Film. Visible Filaments.

b) Results

Figure 4.7 shows graphically the relationship between the biofilm base

thickness and the friction factor exhibited by the fouled reactor. Due to the non-

porous nature of the biofilm base discovered in the porosity study, the constrictional

0.1400.120

0.100

0.0800.0600.0400.020

0.000

Biofilm T h ickn ess (m m )

Figure 4.7 Plot of Biofilm Thickness vs. Friction Factor.

54

effects have been incorporated into the calculation of the friction factor. This graph

shows that there is no direct relationship between biofilm thickness and frictional

losses.

In order to evaluate the effect of the filaments on pressure loss, the

Buckingham Pi Theorem [2] was applied using the traditional parameters: velocity,

hydraulic diameter, fluid density, viscosity, and reactor length. The roughness

element was replaced with the new parameter of total streamer length ( SI, ) per

effective volume (the effective volume resulting after considering the biofilm base

thickness a constriction to the tube diameter). This parameter is shown in equation

4.1. The motivation for choosing this parameter is included in the discussion.

M,.-------------- = 4.1

VoI-Jr

c) Dimensional Analysis

The dimensional analysis based on the parameter provided in 4,1 is as follows.

Step I . Establish the variables on which pressure loss is dependent.

Variable Representation

Ap - pressure loss across length of pipeV = mean fluid velocityDeff = Effective Diameter (Pipe Diameter less 2x the biofilm base thickness I = length of the pipep = dynamic viscosityp - fluid density

55

——---- Total streamer length per effective volume of cross section sample"olCif

Functional Fomi

SI

P h y

Dimensions

Use the primary dimensions of M (kg), L (m), ahd t (s).

Express Variables in Primary Dimensions

Apper length of pipe ->

I L■Dcjf , -> L

M 'Ii ->

L * tLV —>tM :

P ->■ L3

SI,->

IV°hff L2

Step 2. Choice of repeating variables V, Dejj, p

Step 3. Form dimensionless Pi groups based using these repeating variables.

System of equations evaluated for Ap.

A p * r * Ddr**/,'

A = M 0 * T0 * T'0

M I + 0 + 0 + c = 0 c = —l

56

L — 2 + <z + Z> + (—3c) = O 6 = 1

t — 2 + (—la) + 0 + 0 — 0 Ci = -L

System of equations evaluated for //

f0 * r0 * /-rrO

L *t t Z3

M l + 0 + 0 + c =0 C = - I

L — I + cz + 6 + (—3c) = 0 a = —I

t — I + (—a) + 0 + 0 = 0 A = - I

11I =:* Deff

n 2 =

Final system of equations be evaluated for SllK ff

SI,V„fe

I T a M c*z_ = M 0 *Z0 *!r0

r r Zj

M 0 + 0 + 0 + c =0 c = 0

L - 2 + a + b + (-3c) = 0 a = 0 simplifies Io fl3

t 0 + {-a) + 0 + 0 = 0 6 = 2

where Slj represents the crimulative,streamer length measured (per 2-mm sample),

and the denominator of this parameter is the volume of the cross section sample

thickness as shown in Figure 4.3. Measurement data is included in Appendix A.

This volume is calculated using the effective diameter i.e. the initial tube diameter

less 2 times the biofilm thickness. The final variable, I, is the length of the reactor.

tio

n F

acto

r57

The measurements and additional data pertaining to the experiment is included in

Appendix A. By applying dimensional analysis to this parameter using the

repeating variables of effective diameter, velocity and density, the resulting

dimensionless Pi group is

SI,X3 = - j - 4.2

When this new new pi group is plotted vs. friction factor the result is a favorable

linear correlation with R2= 0.789.

0.140

0.120 0.100

0.080

0.060

^ 0.040

0.020

0.000

0 2 4 6 8 10 12 14 16

S tre a m e r L en g th / C onduit L en g th

Figure 4.8 Plot of Friction Factor vs. Dimensionless Factor of Streamer Length / Conduit Length

Discussion and Conclusion

Although many researchers have correlated a dramatic increase in frictional losses

associated with filamentous biofilms, the correlations up until now have been qualitative.

58.

Other investigators have indicated that attributed rippled biofilm topography, similar to

that shown in Figure 4.5, was responsible for the losses. Such observations may have

been the result of viewing the biofouled surface in a moist but not submerged state, i.e.

viewing a pipeline just after draining it out. Viewing the biofouled system in such

conditions the streamers would not be visible and could easily be overlooked.

These results are the first attempt to quantify a relationship between biofilm

structure and pressure loss associated with biofouling. Because natural systems are

complicated, much more so than a manufactured conduit surface, such an attempt is not

trivial and the following topics arise as points of discussion.

I . The replacement of the traditional roughness element (e / D) with the parameter

SI,in the dimensional analysis is not intuitively obvious as a parameter

based simply on the streamer length, but was selected over such a length

because of the following 2 reasons. .

a. This variable accounts for, not only the lengths of streamers, but also the

number (density) of streamers, i.e. this variable accounts for the total

streamer length per volume. By using a simple reference length based on

average streamer length ( Slave), the resulting Pi group, , would not

account for changes in the population of streamers.

59

b. Unlike the traditional roughness elements that perturb flow by protruding into

the flow, the streamers arguably lie parallel to the flow along the conduit wall.

iS7 SISuch a situation is better represented by the Pi group, ~ , rather than

c. It also should be noted that other possibilities exist for the variable such as

total streamer length per area of the conduit wall or streamer length squared

per area of the conduit wall.

2. The measuring technique was done carefully as described in the section 4.2.

Possible areas of uncertainty and the effects on the results are the following,

a. Variations in the thickness of the slice of tubing would be linearly related to a

change in the measurement of Sll•. For instance, a deviation in the length

eff

of the slice of 5% would be reflected in a 5% change in the recorded value of

SI,

'

b. The scale at which the measurements were performed on the images was

approximately 20 times. Changes in this scale could result in significant

changes in the measurements of the streamers. These differences would likely

be reflected in changes of measurements for all the images, but this would not

change the R2 value for the correlation in Figure 4.8.

3. Errors in the measurement of base thickness (Figure 4.4) of the biofilm would

be reflected in Calculations of the mean flow velocity in the reactor. Such errors

would be compounded when calculating the associated friction factor, due to the

60

velocity squared term in the Darcy equation. It is acknowledged that errors in

the thickness measurements may result in up to 15% error in the resulting

friction factor.

Despite these areas of uncertainty regarding the experiment, the results

nonetheless provide a fundamental quantitative relationship between biofilm structure

and pressure loss.

61

THE PHENOMENON OF INCREASING FRICTION FACTORS WITH INCREASING REYNOLDS NUMBERS IN BIOFOULED SYSTEMS.

Introduction

In the previous chapters, it was shown that traditional approaches fall short in

predicting losses in biofouled systems. Several industrial case studies where this

occurred were offered in Chapter I . Chapter 4 illustrated how the streamers lengths

found in some biofouled systems differ from the traditional parameter of the roughness

element, and in Chapter 4 the length of these filaments was related to the energy losses

through a dimensional analysis. In this chapter, another fundamental deviation from the

traditional losses caused by rigid roughness is investigated.

In a rigid, non-fouled conduit system, the frictional resistance and corresponding

friction factor are dependent on the two dimensionless parameters of roughness element

and Reynolds number. The relationship between these two parameters is predicted by the

Moody diagram and the Colebrook correlation for various roughnesses [2].. In addition,

the Blasius solution can be used to predict losses in hydraulically smooth conduits where

s ! D < 0.00005 [2], The Moody diagram applies to flow for all Reynolds numbers while

the Colebrook and Blasius solutions apply only to turbulent flows. For laminar flow the

value of the friction factor is independent of the surface and linearly related to the

Reynolds number [2],

62

To illustrate the general relationship between friction factor and Reynolds number,

consider a piping system of average diameter with the flow slowly increasing from the no

flow condition. Initially the flow velocity is low and the Reynolds number is less than

2,100, resulting in a laminar flow. As the flow velocity increases, the Reynolds number

increases and the friction factor falls off according to

This continues until the flow is increased to a point where the Reynolds number reaches

approximately 2100, and the friction factor is reduced to near 0.03. At this point, the

inertial force of the fluid becomes too high, relative to the viscosity, to maintain laminar

flow, and the flow enters a transitional zone. The transitional zone is characterized by

periods of laminar flow interrupted with bursts of turbulence. Here equation 5.1 becomes

ineffective at predicting the friction factor.

The nature of the flow in the transition zone creates a difficulty in predicting the

values of the friction factor with high degree accuracy. As the flow velocity is increased

to a point where the corresponding Reynolds number exceeds 4,000, the flow is fully

transitioned into turbulent flow. In this flow regime the values of the friction factor

become predictable again to within 5% to 10% using the Moody diagram and related

equations [2],

Unlike the laminar regime, friction factors related to turbulent flow are dependent

upon the roughness of the conduit wall. The Moody diagram provides friction factors for

a range the roughness elements varying from a !D < 0.00005 (hydraulically smooth) up

to 0.05.

63

The different relative roughnesses are represented on the Moody diagram by the set

of curves, all of which begin near a Reynolds number of 4,000. Similar curves can also .

be generated using the Colebrook equation (5.2) for the associated range of Reynolds

numbers and relative roughnesses [2].

V7= - 2.0 logr s !D 2.51 "

3.75.2

Where f is the friction factor, s ! D is the relative roughness and Re is the Reynolds

number. Because of the implicit dependence on the friction factor, f needs to be solved

for with an iterative method.

In addition, the Blasius equation (5.3) [2] can be used to predict losses for Reynolds

numbers in the range of 4,000 to 100,000 for smooth walled conduits.

D * ISpl p * V 2

0.1582 5.3

Where D is the hydraulic diameter, ISpl is the pressure loss, p is the fluid density, V is

the mean fluid velocity, and p is the dynamic viscosity of the fluid. The Blasius solution

is more tractable than the Colebrook solution but is limited to smooth wall conduits and

Reynolds numbers of the given range. Both the Colebrook and Blasius solutions are the

result of fitting an equation to numerous experimental results.

These methods show that for rigid wall pipes, friction factors in the turbulent

regime have their maximum value at Reynolds numbers representing the transition to

. fully turbulent flow. As Reynolds numbers increase from here, the friction factors

64

decrease in the manner shown on the Moody diagram and similarly represented

numerically by Colebrook and Blasius solution.

The goal of the following experimental was to show how friction factors exhibited

by biofouled closed conduits deviate from the aforementioned behavior. A discussion of

possible reasons for the deviation is then offered.

Materials and Methods Specific to Experiment

The general experimental setup is the same as that discussed in Chapter 3. 6.4-mm

inner diameter (9.5-mm outer diameter) clear vinyl tubing was used as the reactor. The

distance between the pressure ports was 0.5-m. Two Cole Parmer impeller pumps (Model

7553-70) placed in parallel provided the flow and a McMillan Co 101 Flo-Sen (Model

6593) measure the flow rate. The flow sensor was calibrated using tap water prior to the

experiment. The biofilm was grown under a volumetric flow rate of 41.9 ml/sec,

resulting in a flow velocity in the reactor of 1.3 m/s. The Reynolds number based on this

the diameter is approximately 7,500.

The reactor was inoculated with 250 ml of activated sludge that was obtained from

the Bozeman wastewater treatment facility, located on Springhill Road in Bozeman

Montana. This was reeirulated for 24 hours in the system with the medium described in

Chapter 3 to ensure attachment before switching to a continuous culture.

Pressure loss data for this experiment was obtained with a Cole-Parmer pressure

transmitter (Model 07354-05) that was calibrated, prior to the experiment, using water

and a u-tube manometer. Pressure readings for this experiment varied from the method

65

described in Chapter 3. For the purpose of this experiment, pressure readings were taken

over a large range of Reynolds numbers based upon the tube diameter and the bulk flow

velocity. Measurements were taken as the pumps and flow were incrementally increased

from a minimum to the maximum possible flow permitted by the pumps.

Results

Prior to fouling the system, pressure measurements were taken over the range of

Reynolds numbers permitted by the flows generated by the pumps. The resulting friction

factors were then compared to those predicted for laminar flow by 64/Re, and those

predicted for turbulent flow by the Blasius solution. The results are shown in Figure 5.1.

The comparison of the actual friction factor to the predicted is favorable in the laminar

flow regime at Reynolds numbers around .1000. Then, as the flow nears the transitional

zone the actual friction factor deviates from the 64/Re. This may be due to an early onset

of the transition zone caused by entrance conditions and/or the fact that the vinyl tube is

not truly a rigid surface. As the flow is increased to Reynolds numbers above 4,000 the

comparison of the friction factor to the Blasius solution is excellent and is numerically

expressed in Table 5.1.

0.08

6 6

oro

LLCO4-»OLL

0.07

0.06

0.05

0.04

0.03

0.02

0.01

Vinyl Reactor

Blasius Solution

0.00 4------------------- —T------------------------------ T----------------------------T--------------------------- T----------------------------T--------------------------- ,

0 2,000 4,000 6,000 8,000 10,000 12,000

Reynolds NumberFigure 5.1 Graph Comparing the Friction Factor Generated by the Non-Fouled Vinyl

Reactor to the Friction Factor Predicted using Tradition Methods for a Hydraulically Smooth Pipe.

Over the following 32 days, biofilm was allowed to grow inside the reactor, and

pressure readings were taken on various days in the manner described. These results can

be seen in Figure 5.2. The spreadsheet data is included in Appendix B. When the

pressure readings were taken, cross sections of vinyl tubing were sliced and removed

from a section of the system separate from 0.5-meter reactor section across which the

pressure loss data was recorded. These cross-sections allowed the biofilm filamentous

structure to be observed.

67

Figure 5.2 shows not only the dramatic increase in the friction factor caused by the

biofilm but also a dramatic deviation from the traditional shape of the curves discussed in

the introduction.

Table 5.1 Friction Factor Values and Comparison to Blasius Solution

# #B lasius

Sf s if3553 0.0408 0.0401 i . 8 % : “4131 0.0407 0.0394 3.2%4399 0.0392 0.0388 0l5%4667 0.0390 0.0382 2.0%4935 0.0387 0.0377 2.6%5203 0.0379 0.0372 1.9%5472 0.0371 0.0367 1.1%5740 0.0370 0.0363 1.8%6008 0.0367 0.0359 2.2%6276 0.0363 0.0355 2.3%6544 0.0362 0.0351 3.1%6813 0.0357 0.0348 2.6%7483 0.0353 0.0340 3.8%8154 0.0342 0.0333 2.8%8824 0.0333 0.0326 2.1%9494 0.0327 0.0320 2.1%

10165 0.0319 0.0315 1.5%10835 0.0311 0.0310 0.5%11221 0.0314 0.0307 2.2%

Also, images of the cross-section samples taken from the system on days 5, 7, 22,

and 32 of the experiment are included (Figure 5.3). Note the similar structure to that of

the biofilm produced by activated sludge in the experiment from Chapter 4.

Fric

tion

Fact

or6 8

0.3000------Clean

Reactor

0.2500

0.2000

0.1500

0.1000

0.0500

Dayl

Day2

Day/

♦ -D a y 18

4— Day 22

— Day 26

0.0000 -------------- i--------------1--------------,--------------,--------------,--------------,0 2,000 4,000 6,000 8,000 10,000 12,000 Day 32

Reynolds Number

Figure 5.2 Plot of Friction Factor vs. Reynolds Number for Various Days of Growth of Biofilm.

Experimental Observations

Three changes with regards to the friction factor of biofouled systems are evidenced

in Figure 5.2.

I . There is a shift up for the value of the friction factors at all Reynolds numbers.

This can be observed after only 2 days of growth. This occurs not only in the turbulent

regime but also at Reynolds numbers representing laminar flow where it is traditionally

thought that frictional losses are independent of surface topography.

69

Figure 5.3 Images of the Cross-Sections Removed from the Experiment.

2 The shape of the curves representing the friction factors changes from that of

the non-fouled system. This is particularly evident at the Reynolds numbers

above 4,000. While the non-fouled reactor shows the traditional diminishing

values of the friction factor as is traditionally associated with higher Reynolds

numbers, the biofouled reactor does not.

3 Over time, the losses in the biofouled reactor appear to reach a pseudo-steady

state. Specifically, days 22 and 32 expressed similar values for the friction

factor and these were not significantly higher than those recorded on day 7,

although the film is noticeably thicker on day 32 (see Figure 5.3 and Figure

5.4)

70

5.5 Discussion

Results from this experiment are further evidence that the phenomenon of increased

friction losses associated with biofilm can not be treated as an increase in the roughness

element. Considering biofilm to function as an increase in the roughness element would

not explain the change in the friction factor observed in Figure 5.2. In particular the

following observations are noted.

I . An increase in the roughness element would not be associated with the increase

in friction factors associated with laminar flow seen in the biofouled system. This

phenomenon is perhaps the most difficult to explain. The fact that this shift can be

observed as early as day 2, when biofilm thickness is under 100 microns, is evidence that

it is not caused by constricting effects of the biofilm on the tube diameter. Traditional

fluid dynamics holds that the flow profile in such system is hyperbolic and the no slip

boundary condition at the pipe wall is independent of the relative roughness of the pipe.

71

It has been hypothesized that the motions of the filaments is responsible for

increased frictional losses [6] [15], but the findings from this experiment regarding the

filament motion and energy loss, tend to contradict this hypotheses.

Although there is no filament motion observed under Reynolds numbers indicative

of laminar flow, the biofouled reactor nonetheless exhibits pronounced frictional losses

for the laminar flow regime. These losses are of the same magnitude of increase as the

losses exhibited under turbulent flow. Therefore, such losses associated with laminar

flow seem unlikely to be largely related to streamer/filament motion. A more feasible

mechanism that would account for the losses in both the laminar and turbulent flow is the

following.

The increased loss in the laminar region is caused by the filaments at the

biofilm/fluid interface. These filaments create a layer near the conduit wall in which the

flow is dramatically slowed by the shear forces acting along the filaments. The effect of

the streamers on the flow profile is illustrated in Figure 5.5. Although the filaments do

not occupy a large volume and, therefore, do not physically constrict the flow, the effect

of the streamers could be significant enough that the resulting flow profile is similar to

that of a constricted pipe. The increased losses in the laminar flow regime would be

caused by this type of decrease in the effective hydraulic diameter and the resulting

increase in the mean flow velocity. This mechanism would be most apparent in smaller

conduits, where the magnitude of the thickness of the streamer layer is significant relative

to the hydraulic diameter. As the size of the hydraulic diameter increases relative to the

thickness of the layer of fluid effected by the streamers, this effect becomes negligible.

72

Traditional Non-Fouled Proposed VelocityParabolic Laminar Profile for Conduit FouledVelocity Profile with Filamentous Biofilm

ConduitDiameter

Filaments/Streamers

Figure 5.5 Illustration of Flow Profile without the Presence of Filamentous Biofilm and Proposed Velocity Profile for Conduit Fouled with Filamentous Biofilm.

Similarly, in turbulent flow, this effect of the filaments would account for an

increase in the apparent frictional losses of smaller diameter biofouled conduits. In

addition, under turbulent flow, the motions exhibited by the filaments further complicate

this situation, and may, in part, account for the second deviation from the traditional

relationship between Reynolds number and frictional resistance.

2. The second significant deviation from the norm seen in the friction factor curves

occurs as the bulk fluid flow is transitioned from laminar into turbulent flow. In the

biofouled system all days, except day 2, show a minimum friction factor around a

Reynolds number of 2,100. But unlike the clean conduit, the biofouled conduits do not

show evidence of a transitional period between the Reynolds numbers of 2,100 to 4,000.

The fouled tubes show a dramatically different behavior in which the friction factors

increase up to higher Reynolds numbers of 7,000 to 9,000. This behavior is illustrated in

73

Figure 5.6 where the friction factor of the clean tube is compared to that of the tube with

22 days biofilm growth.

0.16 0.14

5 0.12 S 0.1c 0.08ip 0.06if 0.04

0.02 0

0

Day 22

Clean Reactor

I I I I I I

2,000 4,000 6,000 8,000 10,000 12,000Reynolds Number

Figure 5.6 Graph Showing the Dramatic Difference in the Friction Factors for the Clean Reactor and the Reactor Fouled with 22 days of Biofilm Growth.

Two possible explanations for this behavior are offered here,

a. The Reynolds number relevant to the friction factor in biofouled systems

should be defined based on a length scale relevant to the biofilm structure

rather than the conduit diameter. A length scale relating to the filaments,

which have been repeatedly observed and related to increased frictional losses

[6] [7] [15] [23] [24] [36], should possibly be investigated. This idea was first

suggested by Lewandowski and Stoodly [15]. They suggested that in

biofouled systems the nature of the bulk fluid flow is governed by the

traditional Reynolds number based on the tube diameter, but perhaps the

74

energy losses should be evaluated with a Reynolds number related to a length

scale pertaining to the biofilm. This would change the flow relevant to losses

from internal conduit flow to the external flow around the complex biofilm.

Losses associated with the Reynolds number and of the exterior flow around

these filaments would be more likely to exhibit the behavior viewed in Figure

5.2. In line with this, is that the peaking of the friction factor at Reynolds

numbers of around 7,000 could coincide with the fact that this was the

Reynolds number of the growth conditions. At higher Reynolds numbers

some biofilm and streamers may have been sloughed off due to the shear

being increased above the growth conditions,

b. This behavior is possibly related to a delay in the onset of full turbulence.

Similar phenomena are documented in literature. Lewandowski [37] showed,

using nuclear magnetic resonant imaging, that entry conditions for conduits

coated with biofilms is shorter than that for the same entrance condition with

no fouling. In addition, Klinzing et al. found a possible delay in the onset of

full turbulence to Reynolds numbers above 10,000 in their investigation of

frictional losses in foam-damped flexible tubes [22]. Such surfaces maybe

comparable to that of a pliable biofilm.

75

Conclusions

1. The filamentous biofilm structure increases frictional resistance in closed

conduits.

2. There are several ways that the increased frictional resistance is not consistent

with an increase in the roughness element.

o Increased losses under laminar conduit flow,

o Friction factor behavior is not consistent within the traditional

transition between laminar and turbulent flow,

o Friction factor behavior at Reynolds numbers between 4,000 and 8,000

increases rather than decreases.

3. A possible explanation for the dramatic increase in losses in relatively small

diameter conduits is that the streamers dramatically slow flow in a region near

the conduit wall.

76

SIMULATING A FILAMENTOUS BIOFILM SURFACE

Introduction

Chapter 5 showed that the filamentous biofilm surface topography is related to

increased frictional losses and Chapter 6 illustrated how these losses deviate

dramatically from those of rigid wall conduits. The goal of the experiment of this

chapter was to simulate the frictional resistance exhibited by the filamentous biofilms

of the experiments presented in Chapter 5 and Chapter 6. In order to simulate a

biofouled conduit, a pipe was lined with filamentous material, subjected to flow, and

the resulting frictional resistance was recorded and compared to the frictional

resistance exhibited by biofouled systems.

Materials and Methods

Acrylic fur, acquired from a local material store, was used to simulate the

fluid/filamentous interface exhibited by some biofilms. Prior to applying the acrylic

fur to the inside of the pipe, the fur was sheared to a shorter length to simulate the

topography discussed in chapter 4. The acrylic fur was then mounted to the inside of

a closed conduit in the following way.

77

A 44" length of vinyl tygon tubing, 3/4" inner diameter (I" outer diamter) was

sliced along its length. A piece of the material that would fit along the inside of this

tube was then cut from the fabric. Multipurpose spray adhesive (3M, Super 77) was

then applied to the material backing of the material. The vinyl tubing was pried open

and the artificial fur was then applied along its inside. This tubing was inserted into a

44" length of PVC pipe, I" inner diameter and I 1/4" outer diameter. The PVC

served to provide a sealed outer structure for the acrylic lined tubing and the snug fit

minimized the effect of the seam running the length of the tubing. Figure 6.1 shows

an image of both the simulated biofouling and an image of the natural biofouling

taken from the experiment of chapter 4.

Simulated Biofilm Actual Biofilm

Figure 6.1 Images of Simulated and Natural Biofouled Conduit.

78

Pressure ports were inserted 8 inches down stream of the entrance and I"

upstream from the exit. Care was taken to ensure that the ports did not protrude

above the material into the flow.

The reactor was placed in a recycle loop similar to that shown in Figure 3.1.

An impeller pump from the Little Giant Pump Co. (Model 4-MD-HC) provided flow.

The flow was throttled using a ball valve. Volumetric flow rate was monitored using

a Great Plains Industries electronic flow meter (Model A104GMN025NA1).

Pressure measurements were taken over the range of flow allowable by the pump. A

Cole-Parmer pressure transmitter (Model 07354-05) was used to monitor the pressure

loss across the reactor.

Figure 6.2 Image of a Cross Section of the Tube Lined with the Acrylic Fur withMeasurement of the Hydraulic Diameter and Showing the Seam Cut along the Reactors Length.

79

Fhe measurement of the hydraulic diameter was made as shown in Figure 6.2.

The measurement was made from the surface of the fabric in order to be conservative

with regards to the hydraulic diameter. By measuring the hydraulic diameter in this

way, flow within fibers of the material is assumed negligible.

Friction factors were calculated by inserting the parameters of mean fluid

velocity, V , hydraulic diameter, D , pressure loss, Ap, and reactor length, I , into the

Darcy-Weisbach equation (1.3).

Results

Figure 6.3 displays the results of the friction for the pipe lined with the

filamentous material. For comparison, a graph of the results for a biofouled tube from

chapter 6 is also included in Figure 6.4.

0 . 2 5

0 . 0 5

2 0 0 0 7 0 0 0 1 2 0 0 0 1 7 0 0 0 2 2 0 0 0

R e y n o l d s N u m b e r

Figure 6.3 Plot of Friction Factor vs. Reynolds Number for the Conduit Lined with the Acrylic Fur.

8 0

0.2500Actual f day 0 ■Actual f Day 320.2000

0.1500

0.1000

0.0500

0.00002000 IO 6000 8<

Reynolds Number10000 12000

Figure 6.4 Plot of the Friction Factor vs. Reynolds Number for Biofouled and Non-fouled conduit.

Discussion

The results show that the artificially fouled pipe shows a dramatic increase in

frictional resistance; unfortunately, these losses can not be characterized as similar to

those shown by the biofouled conduit of chapter 6. The simulated biofiouling

exhibited friction factors similar to a dramatic increase in the relative roughness

rather than to the changes in friction factor caused by biofouling discussed in

chapter 6.

Although both the natural and artificial biofilms have a filamentous structure,

they do have fundamental differences that may result in the differences in the

frictional resistance.

The filaments of the simulated biofilm are more rigid and protrude from the

base into the bulk fluid flow. Even under the higher flow velocities, it is difficult to

81

imagine that the filaments of the simulated film flex or yield much to. the flow. It

would follow that the simulated surface would exhibit frictional resistance more

closely related to a rough rigid surface rather than that of the pliable filamentous

biofilm.

Increased frictional resistance in the simulated system would be attributed

primarily to the pressure drag as the filaments function as cylinders in a cross flow.

Where as the filaments in the biofouled system would more accurately be represented

as cylinders parallel to the flow, in which shear rather than pressure is the primary

mechanism of frictional resistance.

82

A RELATIONSHIP BETWEEN MEAN FLOW VELOCITY AND FRICTION FACTORS EXHIBITED BY BIOFOULED SYSTEMS

Introduction

Chapters 4 through 6 investigated several aspects of the relationship between

biofilm structure and fluid flow. Chapter 7 investigated an attempt to simulate these

losses by artificially simulating the biofilm fluid interface in a closed conduit. Many

insights into the problem of increased friction factor were gained through these

experiments but such information and knowledge cannot easily be applied to industrial

applications.

This chapter looks at the results gathered from these experiments and others, and

shows that frictional losses in biofouled closed conduit systems are more dependent on

the fluid flow velocity arid less dependent on the Reynolds number.

System Layout and Reactors

Twelve experiments were conducted. All shared a layout similar to (Figure 3.1).

The experiments were all run with the media described in Chapter 3 except for four

experiments that used the same buffer but had 240 ppm glucose. Closed channel reactors

83

were used for all experiments. Specifications for the various reactors are given in

Table 7.1.

Table 7.1 Experimental Parameters for the 12 Experiments.

10 |mm-3 species ■ 40 ppm Square Glass 8.00 1.00 4,900 0.693 species 40 ppm Square Glass 8.00 1.00 6,000 0.81Sludge 40 ppm Round Vinyl 6.40 0.50 5,000 0.87

3 species 40 ppm Square Glass 8.00 1.00 7,000 0.983 species 40 ppm Round Glass 7.60 1,22 6,800 1,00Sludge 40 ppm Round Vinyl 6.40 0.50 7,500 1.32

3 species 40 ppm Square Glass 6.00 . 1.00 8,000 1.413 species 40 ppm Round Glass 7.60 1.22 10,800 1.60

zelver tfr 1 N/A Square Glass 12.70 2.32 16,000 1.50zelver tfr 3 N/A Square Roughened 12.70 2.32 7,000 0.653 species 240 ppm Round PVC 19.10 0.94 3,000 0.183 species 240 ppm Round Vinyl 15.90 0.73 3,600 0.253 species 240 ppm Square Glass 6.00 1.00 7,500 1.123 species 240 ppm Round glass 7.60 1.00 7,500 1.40

Constrictional Effects

In order to account for the losses associated with increased fluid velocity caused by

the biofilm constricting the hydraulic diameter; thickness measurements were performed

using a light microscope. Experiments involving three species typically achieved

maximum thickness of 300 to 400 microns. For the experiments involving the activated

sludge, cross-sectioned samples of the vinyl tubing were removed in order to obtain an

accurate measurement of film thickness. This was necessary due to the fact that the

activated sludge achieved thickness up to 1400 microns, far greater than those accurately

measurable using a light microscope. This thickness was then subtracted from the

84

reactors hydraulic diameter when calculating the mean fluid velocity from the volumetric

flow rate. Figures 4.3 and 4.4 show one of these cross-sections from which thickness

measurements were made for the activated sludge.

Results

Prior to fouling, the 12 reactors produced frictional resistance within 5% of values

predicted by the Blasius equation and the Moody diagram. As the reactors became

fouled with biofilm, there were dramatic increases in the frictional resistance. This

increase peaked at a maximum before falling off due to a change in biofilm structure or

sloughing of the film from the conduit wall. The maximum frictional losses correspond

to friction factors ranging from 0.05 to 0.24. The resulting maximum friction factors

from each experiment are provided in Table 7.2.

In Figure 7.1, the maximum friction factor recorded for each of the 12 experiments

are plotted vs. the Reynolds number. These points are then fit with a linear regression

and the resulting R2 value is 0.5175.

85

Table 7.2 Reynolds number, Velocity and Maximum Friction Factor of the 12 Experiments

ReynoldsNumber Velocity (m/s) Maximum

friction factor

4,900 0.69 0.186,000 0.81 0.165,000 0.87 0.127,000 0.98 0.146,800 1.00 0.197,500 1.32 0.128,000 1.41 0.1010,800 1.60 0.0816,000 1.50 0.057,000 0.65 0.193,000 0.18 0.243,600 0.25 0.237,500 1.12 0.067,500 1.40 0.05

0.30L -Oo 0.25(9

= 0.20♦3

c 0.15U-

| 0.10 Ex 0 .05COS

0.000 5 ,000 10 ,000 15,000 20 ,000

R eyn o ld s N u m b er

Figure 7.1 Graph Illustrating the Relationship between Reynolds Number Based onConduit Hydraulic Diameter and Maximum Friction Factors Exhibited by the Fouled Closed Conduit Reactors.

y = -1 E-OSx + 0.2402 R2 = 0.5175

86

In Figure 7.2, the maximum friction factor is plotted vs. mean flow velocity for the

12 experiments. These points are then fit with a linear regression and the resulting R2

value is 0.7923.

y = -0.1294x + 0.2637 R2 = 0.7923O 0.25

c 0.20

c 0.15

S 0.05

0.00 0.25 0.50 0.75 1.00 1.25 1.50

M ean F lo w V e loc ity (m /s)

Figure 7.2 Plot Comparing Mean Flow Velocities to Maximum Friction Factors

Discussion

Figure 7.2 illustrates that the flow rate under which the biofilm is grown is related

to the maximum friction factor exhibited by such systems. The relationship shows a

favorable inverse linear correlation with an R2 value of 0.7923, over the sample range of

flow velocities from 0.18 to 1.6 meters per second. This R2 value is satisfactory for such

a correlation with a natural system and at a minimum reflects that the frictional resistance

exhibited by biofilms diminishes as the flow rate under which the biofilm is grown

87

increases. In addition, the results show that flow velocity, rather than the traditional

parameter of Reynolds number, appears to be more relevant to the frictional losses.

This finding would not have been evident had it not been for the various hydraulic

diameter reactors used in the experiments. If a single hydraulic diameter had been used

for all the experiments, then fluid velocity would be the only variable in the Reynolds

number (7.1), i.e. fluid density, p , hydraulic diameter, D , and fluid viscosity, /u would all

be constant in equation 1.3. Any change in fluid velocity would be directly reflected by a

change Reynolds number. This change would be proportionally the same as the change

in fluid velocity. This would result in both the Reynolds number and fluid velocity

having the same correlation, R2, to the maximum frictional losses.

This dependence of frictional resistance in biofouled systems on velocity suggests

that the resulting biofilm topography is primarily dependent on shear resulting from the

bulk fluid flow. This is in sharp contrast to the parameter factor of Reynolds number on

which friction factors have been traditionally based. To illustrate how this can occur,

consider the liquid solid interface of two water conduit systems with identical Reynolds

number but dramatically different mean flow velocities and hydraulic diameters. The

first system has a hydraulic diameter of 10 cm, and a mean flow rate of 0.2 m/s resulting

in a Reynolds number of 17,857. For a hydraulically smooth conduit with f equal to

0.019, the resulting shear stress is 59.2 N/m2. The second system has a hydraulic

diameter of I cm, and a mean flow rate of 2 m/s, also resulting in a Reynolds number of

17,857. For a hydraulically smooth conduit, the shear tress at the fluid solid interface,

based on a friction factor of 0.019, is 94.81 N/m2, 60% greater than the other system

88

sharing the same Reynolds number. Changes in the shear stress of this magnitude have

no effect on rigid conduit surfaces such as steel or concrete. However, the results from

this experiment suggest that changes in the shear stress of this magnitude likely effect the

topography of biofilm because of its rheological properties.

A possible hypothesis that would explain this type of behavior would be as follows.

When exposed to lower flow velocities that result in lower shear, the resulting biofilm

surface exhibits a highly filamentous morphology. As a result, the surface of the biofilm

is greatly increased, which also greatly increases the shear between the flow and the film.

As flow velocities are increased, the shear on the streamers is increased to a point greater

than that allowable by the physical properties of the film. As a result the streamers are

sloughed off. Also, if the initial flow velocity and resulting shear stress on the film

growth conditions are high, filaments will not form.

To further illustrate this point, consider the two images (Figure 7.3 and 7.4) of

biofouled tubing taken from the experiment that related pressure loss to the b i o f i lm

filaments. Figure 7.3 shows the cross section of a conduit containing the biofilm

exhibiting a significant number of filaments. The film demonstrated a friction factor of

0.116 and the mean bulk fluid velocity for this film was 1.33 m/s. Figure 7.4 shows the

cross section of the same biofilm but after two additional weeks of growth. Here the

biofilm base has grown significantly and the film exhibits few filaments. At this time,

the film demonstrated a frictional resistance of only 0.038. Due to the constrictional

effects of the biofilm base the fluid flow velocity was measured to be 2.85 m/s.

89

Figure 7.3 Biofilm Structure for Figure 7.4 Biofilm Structure forBulk Fluid Velocity of 1.33 m/s. Bulk Fluid velocity of 2.85.

This calculation was based on the volumetric flow rate and the constricted tube diameter

caused by the biofilm base thickness. At this high flow velocity the frictional resistance

created at the biofilm surface is approximately equivalent to that of a hydraulically

smooth surface.

Conclusion

The relationship between flow velocity and frictional resistance has implications for

both the practicing engineer and those further researching this problem.

For the practicing engineer who is designing hydraulic pipelines prone to

biofouling, these results demonstrate that the effects of increased pressure losses due to

biofouling will be minimized if the system is operated at higher flow velocities. If this is

not feasible then the engineer must anticipate the possibility of large frictional resistance

and incorporate additional costs of reduction in flow and treatment measures into design

analysis. Figure 7.2 may be used to conservatively estimate the maximum possible

friction losses in such systems.

• Implications involving further research suggest flow velocity or more specifically,

shear stress, plays a more important role than the traditional parameter of Reynolds

Number with respect to losses in biofouled systems.

90

91

CONCLUSION AND FUTURE WORK

For past researchers, determining a method of predicting frictional losses in closed

conduits was not trivial. To this day, methods such as the Colebrook equation, Blasius

equation and the Moody diagram which are used to predict losses in clean, new conduits,

are methods empirically based on tens of thousands of experiments. Predicting losses in

systems where the manufactured fluid-solid interface of a pipe wall has become covered

with the complex topography of a compliant biofilm is an even more daunting task. The

biofilm greatly complicates an already difficult problem. The research presented in this

thesis provides a further understanding of the problem of increased frictional resistance

associated with biofouled conduits.

In Chapter 2, it was shown that for relatively slow flows in porous biofilms,

velocity profiles can be modeled using flow profiles of a vegetative canopy, or, if enough

assumptions are made, using simple canopy models and the Navier Stokes equations.

Such methods are simplified approaches to a complicated problem, but nonetheless,

resulting profiles generated by these approaches satisfactorily compared to a flow profile

acquired from an actual biofilm. Therefore these methods may be used to estimate flow

profiles in highly porous heterogeneous biofilms.

Unfortunately, for biofilms grown under higher turbulent flow rates in closed

conduits, the situation is further complicated. As shown in Chapter 4, these biofilms

have a non-porous base from which filaments protrude. These filaments exhibit

92

movements as they interact with the turbulent flow. For these filamentous biofilms,

many past researchers have suggested that there is a correlation between the filaments

and the increased frictional losses exhibited by the system. Here in Chapter 4, for the

first time, an experiment was performed which measured the lengths of the streamers and

quantified, using dimensional analysis, a relationship between the lengths of these

filaments and the frictional losses in closed conduits.

Next, in Chapter 5, it was shown that rather than simply increasing frictional losses

in a manner similar to that of an increase in rigid roughness, losses due to biofouling

deviate from losses caused by rigid roughness in the following ways.

1. The biofouled systems exhibited increased losses in the laminar regime.

Traditionally, pressure losses for laminar flows are independent of the interface

topography. The cause of this increase was attributed to the filaments

sufficiently slowing the flow near the interface, i.e. the filaments act as a

pseudo-constriction to the flow. This effect is likely magnified due to the

relatively small scale of the laboratory experiments.

2. Although the friction factors of biofouled systems appear to show a transition

from laminar to transitional flow at Reynolds numbers near 2,100, the friction

factors were not indicative of the transition to full turbulence at a Reynolds

numbers near 4,000

3. Unlike rigid conduits in which friction factors decrease as flow becomes more

turbulent, friction factors in the biofouled systems increased well into the

93

turbulent regime. This increase was possibly attributed to movements of the

streamers and biofilm in response to the bulk fluid flow.

In order to verify that the filaments caused these deviations from the traditional

frictional losses, a simulation of the filamentous interface was attempted. Unfortunately,

the attempt was unsuccessful. The failure was attributed to the fact that the filaments of

the material used to simulate the biofilm were sufficiently stiff to protrude into the flow

rather than lie parallel to the flow as the pliable biofilm filaments were observed to do.

Future attempts to simulate the filamentous interface should focus on finding or

manufacturing a material that more accurately simulates the biofilm filaments, but

obtaining such a material most likely will not be easy.

Finally, in Chapter 7 the results of all the experiments were revisited to find that the

velocity is a more influential factor than Reynolds number on the biofilm structure and

therefore its ability to greatly increase frictional losses. Possible future work could

further verify this by growing biofilms under constant Reynolds numbers, but with

varying velocities.

In order to gain a better understanding of this complex problem, future work

investigating the increased frictional losses associated with biofouling should take a step

back from closed conduits and focus on the shear force exhibited on biofouled flat plate

(or rotating disk) reactors. By utilizing these open channel reactors, future research

would take a simplified approach to the problem by avoiding several of the problems and

pitfalls implicit to the closed conduit flow and the research contained in this thesis. Such

an approach would have the following advantages over research using closed conduits.

94

1. Ifa different characteristic length (as suggested by past researchers and

acknowledged in Chapter 6) is more relevant than hydraulic diameter in biofouled

systems, such a parameter would be more apparent in an open channel system than in

a closed conduit system. In these systems the flow velocities and shear stress could

be increased and decreased without complicating the situation by changing the nature

of the bulk fluid flow from laminar to turbulent.

2. Similarly, this type of system would not mix the problem of internal conduit flow

with the problem of external flow past a biofilm surface. The flow and shear acting

on a flat plate would be based on an external Reynolds number problem and also

when the plate became fouled it would still remain an external flow problem.

3. The fact that biofilms are much easier to observe and sample in open channel flat plat

systems will result in a more easily and better quantified biofilm structure.

4. The constrictional effects of the biofilm discussed in Chapter 4 would be eliminated.

5. From a technical viewpoint, modem techniques used to investigate fluid flow (such as

laser doppler velocimetry) are easier to apply to open channels than to closed

conduits.

Once a better understanding of the shear forces associated with biofilms on flat

plates is obtained, this knowledge could then be used to analyze the shear stresses

causing losses in biofouled closed conduits. Unfortunately, this flat plate type of reactor

does not exactly represent the problem associated with biofilms in industry but

understanding these systems would be a major step toward properly modeling the exact

mechanism of loss associated with biofouling of closed conduits.

95

A recommendation for further research on frictional losses in closed; conduits is that

the research should focus on conduit systems of larger scale. Due to the relatively small

hydraulic diameters used in the experiments of this thesis, the constrietional effects as

discussed in Chapter 5, many times dominated the losses. By increasing the scale of the

piping systems, the constrietional effects would be minimized and the frictional losses

would be highlighted. Because of the large costs associated with increasing the scale of

the experimental systems, perhaps future investigation should focus on actual industrial

systems that are plagued with the problem of increased frictional resistance caused by

biofouling. Previous studies into industrial systems with this problem have inspected the

biofilms only after draining out the system, by doing this one hides the filaments and

overlooks the real cause of the losses. By determining and utilizing a method of in-situ

sampling of industrial systems, future research could better quantify a relationship

between the filamentous biofilm structure and the losses for which they are responsible.

96

REFERENCES CITED

1. Currie, I.G., “Fundamental Mechanics of Fluids" McGraw-Hill, Inc., New York, New York, 1993.

2. Munson, B. R., Young, D. F., Okiishi, T. H. “Fundamentals of Fluid Mechanics,” John Wiley and Sons, 1994, New York, New York.

3. Pigott, R. J. S., "The Flow of Fluids in Closed Conduits," Mechanical Engineering, vol. 55, 1933, pp. 497-501.

4. Nikkuradse, J . ,"Stromungsgesetze in Rauhen Rohen," V.D.I Forschungsheft 361, Berlin, 1933, pp. 1-22.

5. Moody, L. F., “Friction Factors for Pipe Flow, ” Transactions o f the A.S.M.E. November 1944, pp.671-684.

6. Picologlou B. F., N. Zelver, and W. G. Characklis, (1980) “Biofilm Growth and Hydraulic Performance” Journal o f the Hydraulic Division, ASCE, vol 106, No HY5, 1980, pp. 733-746.

7. McEntee, W., “Variation of Frictional Resistance of Ships with Condition of Wetted Surface, ” Transactions Society o f Naval Architects and Marine Engineers, Vol. 24,1915, pp. 37-42.

8. Pollard, A. L., House, H. E., “An Unusual Deposit in a Hydraulic Tunnel,” Journal o f the Power Division, Proceedings o f the American Society o f Civil Engineers, December 1959.

9. Seifert, R., Kruger, W,. “Uberraschend hohe Reibungsziffer einerFemwasserleitung,” Verein Deutscher Ingenieure Zeitschrift, Vol.92, 1950, pp.189-191.

10. Characklis, W. G., “Attached Microbial Growth - II. Frictional Resistance Due To Microbial Slimes,” Water Research, Permagon Press Vol. 7, 1973, pp 1249- 1258.

97

11. Bauer H., "Flow Resistance in Pipes with Rippled Roughness," Chem. ztg., 1963, 87, 199.

12. Schlicting H., Boundary Layer Theory, 4th edn. McGraw-Hill, 1960, New York

13. Kouwen, N. and Unny, T.E., “Flexible Roughness in open channels”. Journal o f the Hydraulics Division ASCE, Vol.99, No.HYS, May 1973, pp.713-728.

14. Minkus, A. J., "Determination of Hydrualic Capacity of Pipelines," J. New Engl. Wat. WksAss., 1956, 68, pp. 1-10.

15. Lewandowski Z. and P. Stoodly “Flow Induced Vibrations, Drag Force and Pressure Drop in Closed Conduits Covered with Biofilm,” Wat. Sci Tech1, 32(8), 1995, pp. 19-26.

16. Stoodley, P., Lewandowski, Z., Boyle, J.D., and. Lappin-Scott, H.M., "Oscillation Characteristics of Biofilm Streamers in Turbulent Flowing Water as Related to Drag and Pressure Drop," Biotechnology and Bioengineering. 57, 1998, pp. 536- 544.

17. Loeb, Gd., Taster, D., Gracik, T., (1984) “The influence of Microbial Fouling Films on Hydrodynamic Drag of Rotating Discs, ” Marine Biodeteriation, An Interdisciplinary Study, J.D. Costlow and R.Tipper, eds., Naval Institute Press, Annapolis, MD, 1984, pp.88-94.

18. Klinzing, G. E., J. Biomed. Mater. Res., in press, 1962.

19. Kramer, M., Naval Eng.J. 74, 341,1962.

20. Looney, W.R., Blick, E. F., J. Space Rockets 3, No. 10, 1562,1966.

21. Pelt, R. J., Ph.D. dissertation, University of Pittsburgh, 1964.

22. Klinzing, G.E., Kubovcik R.J., Marmo, J. F. “Frictional Losses in Foam-Damped Flexible Tubes,” I&EC Process Design and Development, Vol. No. I January 1969, pp.112-114.

23. Lewkowicz, A. K., Das, D. K., “Turbulent Boundary Layers on Rough Surfaces With and Without a Pliable Overlayer: A Simulation af Marine Biofouling,” Proceedings o f the A.S.M.E/A.S.C.E. Bioengineering, Fluid Engineering, and Applied Mechanics Conference, 1981, pp. 174-186.

24. Schultz, M. P., Swain, G. W., “The Effect of Biofilms on Turbulent Boundary ' Layers,” Journal o f Fluids Engineering, March 1999, Vol. 121.

98

25. Wimpenny, J. W., Colasanti, R., "A Unifying Hypothesis for the Structure of Microbial Biofilms Based on Celluar Automanton Models," FEMS Microbiology Ecology, 1997, 22, pp. 1-16.

26. De Beer, D., Stoodley, P., Lewandowski, Z., "Measurement of Local Diffusion Coefficients in Biofilms by Microinjection and Confocal Microscopy," Biotechnol Bioeng., 1997, 53, pp.151-158.

27. Picioreanu, C., van Loosdrecht, M. C. M., Heijnen, J. J., "A New Combined Differential-Discrete Cellular Automaton Approach for Biofilm Modeling: Application for growth in Gel Beads," Biotechnol Bioeng, 1998, 57, pp. 718- 731.

28. Bishop, P.L., "Biofilm Structure and Kinetics," Wat. Sci Tech, 1997, 36, pp.287-294.

29. Yang, S., Lewandowski, Z. "Measurement of Local Mass Transfer Coefficient in Biofilms," Biotechnol Bioeng, 1995, 48, pp.737-744.

30. Keevil, C. W., Dowsett, A. B., Rogers, I , "Legionella Biofilms and their Control," Society for applied Bacterilogy technical series, Microbial biofilm,1993,pp.201-215.

31. Cionco R. M., “A wind-profile index for canopy flow,” Bound. Layer Meteor., 3, 1972, pp. 255-263.

32. Stolzenbach K. D. “Particle Transport and Attachment,” Berlin, Germany: Dahlem Konferenzen, 1989, pp. 33-46, Springer-Verlag, Berlin.

33. Xia, F., Beyenal, H., Lewandowski, Z., “An Electrochemical Technique to Measure Local Flow Velocities in Biofilms” Wat. Res. Vol 32, No.12,1998, pp.3631-3636.

34. Cionco R. M., “A Mathematical Model for Air Flow in a Vegetative Canopy,” Journal of Applied Meteorology, Vol. 4, August 1965, pp. 517-522.

35. Characklis, W. G., Marshall, K. C. “Biofilms,” John Wiley and Sons, New York, New York.

36. Haslbeck, E. G. Bohlander, G., “Microbial Biofilm Effects on Drag - Lab and Field,” Proceedings, 1992, S.N.A.M.E Ship Production Symposium

99

37. Lewandowski Z., Altobelli S. Water flow in a Narrow Conduit Covered with Biofilm.. International Association on Water Quality Research Seminar "Biological Degradation of Organic Chemical Pollutants in Biofilm Systems. 19- 21 May 1994, Kollekolle, Copenhagen, Denmark,

100

APPENDICES

101

APPENDIX A

EXPERIMENTAL DATA RELEVANT TO CHAPTER 4

102

Spreadsheet Data / Measurements of Streamer Lengths for the Experiment of Chapter 4

L e n g th c o r r f a c to r L e n g th c o r r f a c to r L e n g th c o r r f a c to r L e n g th c o r r f a c to r L e n g th c o r r f a c to r L e n g th c o r r f a c to r L e n g th c o r r f a c to r L e n g th L e n g t h

0 .7 8 2 0 5 8 0 .9 9 1 8 6 6 0 .1 1 6 5 8 8 0 .866891 0 .8 1 9 6 3 5 0 .8 5 7 0 2 4d .1 8 1238 0 .038101 0 .2 /3 6 8 6 0 .048781 0 .3 6 2 7 0 5 0 .6 8 7 0 8 4

0 .O844O9 0 .6 2 7 6 1 3 0 .0 4 4 8 9 4 0 .3 2 8 6 0 3 0 .4 8 8 8 9 31 .02 4 0 1 6 0 .3 8 7 5 3 8 1.035893 0 .0 4 6 0 4 4 0 .085551 0 .6 2 6 2 2 5

3 0 .9 0 0 5 8 0 .0 7 0 2 8 6 0 .5 0 4 8 8 2 0 .4 0 4 7 0 8 0 .0 2 8 2 2 90 .1 1 0 8 5 3 0 .0 9 1 1 6 6 0 .077801 0 .0 8 5 1 7 2 0 .6 1 1 8 1 6 0 .1 5 2 3 8 8 0 .0 7 0 2 7 5 0 .4 9 7 0 9 3 0 .0 6 0 0 4 6 0 .4 3 9 5 2 9

0 .0 4 9 1 0 3 0 .4 0 2 3 5 7 0 .0 8 0 1 8 5 0 .0 7 8 9 5 6 0 .4 4 0 7 6 3 0 .0 6 0 9 2 90 .0 4 5 3 0 3 0 .9 5 0 1 5 8 0 .0 7 2 0 7 9 0 .0 6 2017 0 .0 6 7 5 3 7 0 .0 2 2 7 1 6 '0 .1 6 6 2 8 2

0 .026271 0 .2 1 5 0 5 3 0 .0 7 7 2 2 6 0 .0 8 1 1 3 9 0 .6 0 3 2 9 8 0 .0 4 0 9 9 6 0 .3 0 0 0 7 2 0 .0 3 8 5 8 2 0 .2 0 6 5 6 70 .8 4 5 9 0 7 0 .2 8 6 5 3 8 0 .8 8 2 8 7 5

# M e a s u r e # M e a s u r e c 0 .0 2 0 /9 8 b .0839 9 6 0 .6 0 3 3 6 4 0 .0 2 6 2 8 3 0 .0 2 2 5 7 6 0 .0 8 3 7 1 5 0 .0 1 9 3 4 90 .3 0 9 9 8 6 0 .0 9 8 9 2 6 0 .0 6 2 0 6 5 0 .3 0 6 3 7 5 0 .0 7 8 7 4 9 0 .557031 0 .0 3 0 4 6 4

0 .85 8 2 0 1 0 .0 8 5 4 0 3 0 .6 3 5 0 0 2 0 .3 3 8 6 9 3 0 .0 6 3 5 7 4 0 .0 4 1 7 0 7 0 .3 0 5 2 9 30 .8 6 0 1 4 5 0 .4 5 6 0 9 7 0 .063201 0 .4 6 2 5 9 8 0 .3 9 /2 9 5

0 .1 5 9 9 3 8 0 .5 1 4 8 0 5 0 .0 5 8 7 9 2 0 .0 6 0 6 8 9 S M e a s u r e d0 .8 0 5 9 6 4 0 .0 7 4 9 0 7 0 .5 3 8 0 7 5 0 .0 6 1 8 0 9

0 .0 8 6 6 6 2 0 .0 5 5 3 9 4 0 .3 9 7 9 0 7 0 .1 3 1 0 7 9 0 .9 7 4 6 2 8 0 .0 2 5 5 7 9 0 .0 7 0 7 2 5 0 .5 0 0 2 7 30 .0 9 2 4 5 8 0 .0 7 3 0 4 2 0 .5 8 6 3 6 2 0 .0 9 6 6 0 8 0 .6 8 3 3 5 50 .0 4 6 5 5 9 0 .0 6 0 9 2 9 0 .0 6 8 2 5 9 0 .4 8 2 8 2 90 .0 3 2 0 7 3 0 .2 3 0 3 8 6 0 .0 9 8 8 8 2 0 .0 8 3 0 6 9 0 .6 0 8 0 2

0 .1 0 1 8 3 8 0 .0 5 0 7 2 9 0 .5 3 8 5 5 2 0 .0 6 1 3 9 5 0 .0 7 6 8 5 0 .5 4 3 5 9 90 .0 5 8 5 0 2 0 .0 8 9 5 9 6 0 .0 9 9 0 7 2 0 .0 5 8 7 9 2 0 .4 3 0 3 2 9 0 .0 5 1 9 3 20 .0 7 6 3 1 9 0 .5 5 0 8 3 7 0 .0 4 4 6 8 5 0 .3 2 0 9 8 7 0 .1 0 7 7 9 50 .0 7 9 2 5 3 0 .0 3 4 6 6 2 0 .2 4 8 9 8 3 0 .776201 0 .1 0 1 0 70 .0 6 6 4 9 9 0 4 8 6 9 6 6 0 .0 9 5 7 5 4 0 .0 2 8 1 1 2 0 .1 9 8 8 5 4

0 .0 /6 8 5 1 0 .0 8 2 9 6 6 0 .6 1 6 8 8 5 0 .6 7 8 7 0 2 0 .0 3 8 2 6 60 .1 2 5 0 4 6 0 .5 3 1 7 0 5 0 .4 9 9 0 7 6 0 .1 2 6 1 8 5 0 .9 2 3 6 0 90.099951 0 .0 5 3 5 5 6 0 .3 8 4 7 0 6 0 .0 2 4 8 8 9 0 .1 8 5 0 5 6 0 .0 2 5 8 9 7 0 .1 8 3 1 8 50 .0 5 9 4 8 3 0 .3 8 /6 3 8 0 .0 9 0 0 3 2 0 .0 4 3 9 0 50 .0 6 2 8 2 9 0 .0 8 0 9 4 3 0 .63 7 6 0 1 0 .5 5 0 9 5 40 .0 7 7 3 6 8 0 .5 5 8 4 0 8 0 .0 6 6 1 6 4 0 .1 0 9 3 2 3 0 .81 2 8 6 2 0 .0 8 8 0 6 6 0 .0 3 6 2 4 80 .0 8 6 9 3 2 0 .0 7 7 8 9 8 0 .5 7 0 1 7 6

0 .0 6 4 0 4 6 0 .4 6 0 0 5 8 0 .1 6 7 9 8 7 0 .0 5 8 6 2 8 # M e a s u r e d0 .6 3 9 6 6 6 0 .9 0 8 9 9 7 0 .1 0 5 7 0 3

# M e a s u r e a 0 .0 8 8 2 6 0 .6 4 6 0 1 60 .1 0 9 3 4 9 0 .7 8 5 4 8 6 0 .0 9 2 2 6 30 .0 3 3 6 2 6 0 .2 8 4 6 4 6 0 .896561 0 .0 3 8 5 9 4 0 .2 8 2 4 9 20 .0 9 3 1 1 7 0 .6 6 8 8 8 6 0 .8 5 2 8 0 4 0 .0 2 0 2 6 8 0 .1 4 8 3 5 3

0 .0 7 1 9 1 4 0 .5 3 4 7 0 8 0 .0 4 9 4 6 4 0 .36 2 0 5 10 .076851

0 .0 2 4 9 3 5 0 .1 8 5 3 9 8 0 .4 7 6 2 2 60 .4 3 6 0 9 9 0 .0 2 7 8 0 5 I tM e a s u r e d

0 .0 5 9 8 7 6 0 .4 3 0 1 0 90 .0 8 0 4 7 7 0 .5 /8 0 8 8 0 .0 2 2 8 1 6

0 .4 3 6 0 9 9 0 .0 8 3 1 3 8 0 .6 1 8 1 6 40 .050221 0 .3 6 0 7 5 3 0 .0 9 1 7 1 9 0 .6 8 1 9 6 80 .044301

0 .302181 # M e a s u r e d0 .0 5 1 6 8 60 .0 3 8 0 8 9

0 .3 2 8 9 9 50 .0 3 3 2 9 5

0 .5 0 9 8 0 6I tM e a s u r e d

5 3 |

Spreadsheet Data for the Experiment of Chapter 4 Calculations of Parameters of the Darcy Equation

Date Day

Number of Streamers Measured

AverageLength

BiofilmThickness(mm)

Velocity(m/s) Diameter

FrictionFactor KPa

Total Streamer Length / Sample Length

21-May a 5 5 5 0 500 5 5 7 5735 0.039 2.23 5 5 525-May 5 3 0737 07T5 0755 5705 0.043 3.0/ TT3528-May TZ 5 5 2 5 2 0 T700 5755 OOtiO 4 /2 57T7

3-Jun 3T 5 5 3 0727 1.04 5752 0.082 7.20 T575310-Jun 25 53 55T 0730 1.07 5773 0.088 8.00 27752Iti-Jun 3T 33 5 5 3 5 5 0 1.33 57T3 O.llti 18 88 2775321-Jun 35 35 5 3 5 0755 TT32 3755 0.113 21.53 18.1030-Jun 35 25 5 3 3 0757 T752 3730 0.0/0 24./ti T2755IikJuI 55 ^TT 5 3 0 1.25 2735 3753 0.039 28.95 373721-Jul 55 T 3 0735 1.42 2755] 3752 0.038 40./5 57TT

103

Sample of Images Taken for the Experiment of Chapter 4

Week 1 Week 2

Week 3

Week 5 Week 6

104

APPENDIX B

EXPERIMENTAL DATA RELEVANT TO CHAPTER 5

105

Spreadsheet Data for Chapter 5 ExperimentDay I

8t Q8tFlow R ate (m A3/sec) VEL

Im er D iam eter 1/4 inch

R e #

Voltage#4

P ressu reT ransducer

P ressu reDrop

(KPa)

Actualf

BlasiusEq-

Predicted. f

P ercen tDifference(Actual from Predicted)

" 20 ^ 3 ^ 5 9 9 4 3.5994 O t 15666 0.00635 .. 644 0.944 6.044 0.0857 6.6993 14 Vo30 5 .0 9 /4 5 .0 9 /4 0 .160958 0.00635 5TB 0.947 0 .0 /0 0.0684 0.0701 2%30 6.5954 6.5954 0.208259 0.00635 1181 0.950 0.096 0.0562 0.0542 4%50 8.0934 8.0934 0.255561 0.00635 1449 0.954 0.131 0.0509 0.0442 15%60 9.5914 9.5914 0.302862 0.00635 1 /1 / 0.959 0 .1 /4 0.0483 0.0373 3D%fO 11.0894 11.0894 0.350164 0.00635 1985 0,964 0.218 0.0452 0.0322 40%BO 12 .5 8 /4 1 2 .5 8 /4 0 .3 9 /4 6 5 0.00635 2253 0 .9 /0 0 .2 /0 0.0435 0.0459 5%90 14.0854 14.0854 0.444767 0.00635 2522 0 .9 /7 0.331 0.0426 0.0446 4%

TOO 15.5834 15.5834 0.492068 0.00635 2 /9 0 0.986 0.410 0.0431 0.0435 T%17.0814 0.53937 0.00635 3058 0.997 0.505 0.0442 0.0425 4%

T20 18.5794 1 8 .5 /94 0 .5866 /1 0.00635 3326 1.010 0.619 0.0458 0.0416 10%T30 2 0 .0 / /4 2 0 .0 7 /4 0 .6 3 3 9 /3 0.00635 3594 1.024 0. /41 0.0469 0 .0 4 0 8 T5%T30 2 1 .5 /5 4 2 1 .5 /5 4 0 .6 8 1 2 /4 0.00635 3863 1.040 0.880 0.0483 0.0401 20%150 0.728576 0.00635 4131 1.055 1.011 0.0485 0.0394 23%TOO 24.5714 24.5714 0 . / / 5 8 /7 0.00635 4399 1.0/1 1.150 0.0486 0.0388 25%T70 26.0694 26.0694 0.823179 0.00635 4 6 6 / 1.090 1.316 0.0494 0.0382 29%TBO 2 / 5 6 / 4 27.5674 0.87048 0.00635 4935 1.111 1.499 0.0504 0.0377 34%T90 29.0654 29.0654 0 .9 1 7 /8 2 0.00635 5203 1.133 1.691 0.0511 0.0372 37%200 30.5634 30.5634 0.965083 0.00635 5 4 /2 1.155 1.883 0.0514 0.0367 40%2TU 32.0614 32.0614 1.012385 0.00635 5 /4 0 1 .1 /8 2.083 0.0517 0.0363 42%220 33.5594 33.5594 1.059686 0.00635 6008 1.200 2 .2 /5 0.0516 0.0359 44%230 35.0574 35.0574 1.106988 0.00635 6276 1.228 2.519 0.0523 0.0355 47%230 36.5554 36.5554 1. 154289 0.00635 6544 1 .2 5 / 2. / / 2 0.0529 0.0351 51%250 38.0534 38.0534 1.201591 0.00635 6813 1 .2 /8 2.955 0.0521 0.0348 50%275 4 1 ./9 8 4 41.7984 1.319844 0.00635 7483 1.356 3.634 0.0531 0.0340 55%300 45.5434 45.5434 1.438098 0.00635 8154 1.425 4.236 0.0521 0.0333 57%325 49.2884 49.2884 1.556352 0.00635 8824 1.531 5.160 0.0542 0.0326 55%350 53.0334 53.0334 1 .6 /4 6 0 5 0.00635 9494 1.616 5.901 0.0536 0.0320 67%375 5 6 . / /8 4 5 6 .7 /5 4 I . /92859 0.00635 10165 1.606 5.813 0.0460 0.0315 46%300 60.5234 60.5234 1.911113 0.00635 10835 1 .6 /0 6 .3 /1 0.0444 0.0310 43%350 53.0334 53.0334 1 .6 /4 6 0 5 0.00635 9494 1.699 6.624 0.0601 0.0320 88%

Data for Chapter 5 Experiment Day 2

81 081 Flow R ate (m A3/sec) VEL

Inner D iam eter 1/4 inch

R e #

V oltage#4

P ressu reT ran sd u ce

: r ■

P ressu reDrop

(KPa)

Actualf

BlasiusEq-

Predictedf

P ercen tD ifference(Actual from Predicted)

20 3.5994 3.5994 6.113856 8.00635 644 0.949 0.687 0 .1 )1 6 0.0993 ... 73%30 5 .0 9 /4 5 .0 9 /4 0.160958 0.00635 3T3 0.952 0.113 0.1112 0.0701 55%40 6.5954 6.5954 0.208259 0.00635 1181 0.958 0.166 0.0971 0.0542 75%50 8.0934 8.0934 0.255561 0.00635 1449 0.964 0.218 0.0849 0.0442 52%BO 9.5914 9.5914 0.302862 0.00635 1717 0.970 0.270 075750 0.0373 101%70 11.0894 11.0894 0.350164 0.00635 1985 0.979 0.349 0.0724 0.0322 124%80 1 2 .5 8 /4 12.5874 0.397465 0.00635 2253 0.993 0.471 0.0758 0.0459 55%SO 14.0854 14.0854 0.444767 0.00635 2522 1.007 0.593 0.0762 0.0446 7T%

TOO 15.5834 15.5834 0.492068 0.00635 2790 1.023 0.732 0 .0 /6 9 0.0435 77%TTO 17.0814 17.0814 0.53937 0.00635 3058 1.041 0 .889 0.0778 0.0425 83%T20 18.5794 18.5794 0.586671 0.00635 3326 1.058 1.037 0.0767 0.0416 84%T30 20.0774 20.0774 0.633973 0.00635 3594 1.076 1.194 0.0756 0.0408 85%T40 21.5754 21.5754 0.681274 0.00635 3863 1.098 1.386 0.0760 0.0401 90%T50 23.0734 23.0734 0.728576 0.00635 4131 1.123 1.604 0.0769 0.0394 55%TBa 24.5714 24.5714 0.775877 0.00635 4399 1.160 1.926 0.0814 0.0388 110%TTO1 26.0694 26.0694 0.8231 /9 0 .00635 4667 1.183 2TT77 0.0799 0.0382 109%TBO1 27.5674 2 / .5 6 /4 0 .87048 0.00635 4935 1.220 2.449 0.0823 0.0377 118%TSO1 29.0654 29.0654 0.917782 0.00635 5203 1.249 2.702 0.0816 0.0372 119%203 30.8634 80.5634 0.965083 0.00635 5472 1.279 2.963 0.0810 0.0367 120%2TTT 32.0614 32.0614 1(112385 U.UU635 57401 1.275 2.928 0.0727 0.0363 100%223 33.5594 33.5594 1.059686 0.00635 6008 1.306 3 .199 0.0725 0.0359 102%23G 35.0574 35.0574 1.106988 0.00635 6 2 /6 1.330 3 .408 0.0708 0.0355 55%243 36.5554 36.5554 1.154289 0.00635 6544 1.362 3.687 0.0704 0.0351 100%253 38.0534 38.0534 1.201591 0.00635 6813 1357 3 5 5 7 0.0704 0.0348 102%275' 41.7984 41.7984 1.319844 0.00635 7483 1.488 4 .785 0.0699 0.0340 106%30ff 45.5434 45.5434 1.438098 0.00635 5T54 1330 57035 0.0694 0.0333 109%325 49 .2884 49.2884 1.556352 0.00635 8824 1.690 6 .546 0.0688 0.0326 111%353 53.0334 53.0334 1.674605 U.UUO&L 9494 1.780 7.330 0.0665 0.0320 108%

Data for Chapter 5 ExperimentDay 7

106

Bt QBtFlow R a te (m A3 /se c )

VELInner

D ia m e te r 1 /4 inch

R e #

V oltage#4

P re ss u reT ra n sd u c e

:: I r

P re s s u reD rop

(K P a)

A ctualf

B las iu sEq.

P re d ic te df

P e rc e n t D iffe ren ce (Actual from Predicted)

.Io1 3 .5 9 9 4 3 .5 9 9 4 0 .1 1 3 6 5 6 ~trnoB35 644 cT9S2 0 .1 2 7 0.2499 0.0993 152%3U" 5 .0 9 7 4 5 .0 9 7 4 0 .1 6 0 9 5 8 0 .0 0 6 3 b 9TT 0 .9 5 6 0 .1 4 8 0 .1 4 5 5 0.0701 T07%30 6 5 9 5 4 6 .5 9 5 4 0 .2 0 8 2 5 9 0 .0 0 6 3 5 TTET 0 .9 6 3 0 .2 0 9 0 .1 2 2 7 0 .0 5 4 2 126%50 8 .0 9 3 4 8 .0 9 3 4 0 .255561 0 .0 0 6 3 5 1449 0 .9 7 0 0 .2 7 0 0 .1 0 5 3 0 .0 4 4 2 138%50 9 .5 9 1 4 9 .5 9 1 4 0 .3 0 2 8 6 2 0 .0 0 6 3 5 1717 0.981 0 .3 6 6 0 .1 0 1 6 0 .0 3 7 3 172%70 1 1 .0 8 9 4 11 .0 8 9 4 0 .3 5 0 1 6 4 0 .0 0 6 3 5 1985 0 .9 9 4 0 .4 7 9 0 .0 9 9 5 0 .0 3 2 2 209%50 1 2 .5 8 7 4 12 .5 8 7 4 0 .3 9 7 4 6 5 0 .0 0 6 3 5 2 253 1 .010 0 .6 1 9 0 .0 9 9 7 0 .0 4 5 9 117%50 1 4 .0 8 5 4 14 .0 8 5 4 0 .4 4 4 7 6 7 0 .0 0 6 3 5 2 522 1 .029 0 .7 8 4 0 .1 0 0 9 0 .0 4 4 6 126%

TOO 1 5 .5 8 3 4 15 .5 8 3 4 0 .4 9 2 0 6 8 0 .0 0 6 3 5 2 7 9 0 1.050 0 .9 6 7 0 .1 0 1 7 0 .0 4 3 5 1 34%TTO 1 7 .0 8 1 4 17 .0 8 1 4 0 .5 3 9 3 7 0 .0 0 6 3 5 305 8 1.075 1 .185 0 .1 0 3 7 0 .0 4 2 5 144%TZO 1 8 .5 7 9 4 18 .5 7 9 4 0 .586671 0 .0 0 6 3 5 3 3 2 6 1 .104 1.438 0 .1 0 6 3 0 .0 4 1 6 156%TOO 2 0 .0 7 7 4 2 0 .0 7 7 4 0 .6 3 3 9 7 3 0 .0 0 6 3 5 359 4 T7T35 1 .717 0 .1 0 8 7 0 .0 4 0 8 166%T30 2 1 .5 7 5 4 2 1 .5 7 5 4 0 .6 8 1 2 7 4 0 .0 0 6 3 5 386 3 1 .164 1.961 0 .1 0 7 5 0 .0401 168%T50 2 3 .0 7 3 4 2 3 .0 7 3 4 0 .7 2 8 5 7 6 0 .0 0 6 3 5 4131 1 .205 2 .3 1 8 0 .1 1 1 2 0 .0 3 9 4 182%T50 2 4 .5 7 1 4 2 4 .5 7 1 4 0 .7 7 5 8 7 7 0 .0 0 6 3 5 4 3 9 9 1 .240 2 .6 2 3 0 .1 1 0 9 0 .0 3 8 8 186%T70 2 6 .0 6 9 4 2 6 .0 6 9 4 0 .8 2 3 1 7 9 0 .0 0 6 3 5 466 7 1 .284 3 .0 0 7 0 .1 1 2 9 0 .0 3 8 2 195%TOO 2 7 .5 6 7 4 2 7 .5 6 7 4 0 .8 7 0 4 8 0 .0 0 6 3 5 4 9 3 5 1 .319 3 .3 1 2 0 .1 1 1 2 0 .0 3 7 7 19 5 %TOO 2 9 .0 6 5 4 2 9 .0 6 5 4 0 .9 1 7 7 8 2 0 .0 0 6 3 5 520 3 1.371 3 .7 6 5 0 .1 1 3 8 0 .0 3 7 2 206%ZOO 3 0 .5 6 3 4 3 0 .5 6 3 4 0 .9 6 5 0 8 3 0 .0 0 6 3 5 5472 1 .425 4 .2 3 6 0 .1 1 5 7 0 .0 3 6 7 215%ZTO 3 2 .0 6 1 4 3 2 .0 6 1 4 1 .0 1 2 3 8 5 0 .0 0 6 3 5 574 0 1 .465 4 .5 8 4 0 .1 1 3 8 0 .0 3 6 3 214%ZZO 3 3 .5 5 9 4 3 3 .5 5 9 4 1 .0 5 9 6 8 6 0 .0 0 6 3 5 600 8 1 .510 4 .9 7 7 0 .1 1 2 8 0 .0 3 5 9 214%Z30 3 5 .0 5 7 4 3 5 .0 5 7 4 1 .1 0 6 9 8 8 0 .0 0 6 3 5 6 2 7 6 1 .579 5 .5 7 8 0 .1 1 5 9 0 .0 3 5 5 226%230 3 6 .5 5 5 4 3 6 .5 5 5 4 1 .15 4 2 8 9 0 .0 0 6 3 5 6 5 4 4 1 .635 6 .0 6 6 0 .1 1 5 9 0 .0351 230%250 3 8 .0 5 3 4 3 8 .0 5 3 4 1 .201591 0 .0 0 6 3 5 6 8 1 3 1 .700 6 .6 3 3 0 .1 1 6 9 0 .0 3 4 8 236%275 4 1 .7 9 8 4 4 1 V 9 8 4 1 .3 1 9 8 4 4 0 .0 0 6 3 5 7483 1.870 8 .1 1 4 0 .1 1 8 6 0 .0 3 4 0 249%300 4 5 .5 4 3 4 4 5 .5 4 3 4 1.4 3 8 0 9 8 0 .0 0 6 3 5 815 4 2 .0 4 0 9 .5 9 6 0.1181 0 .0 3 3 3 255%325 4 9 .2 8 8 4 4 9 .2 8 8 4 1 .556 3 5 2 0 .0 0 6 3 5 882 4 2 .2 5 0 1 1 .426 0.1201 0 .0 3 2 6 268%350 5 3 .0 3 3 4 5 3 .0 3 3 4 1 .674 6 0 5 0 .0 0 6 3 5 949 4 2 .4 3 0 12 .995 0 .1 1 7 9 0 .0 3 2 0 268%375 5 6 .7 7 8 4 5 6 .7 7 8 4 1 .7 9 2 8 5 9 0 .0 0 6 3 5 10165 2 .0 1 0 9 .3 3 5 0 .0 7 3 9 0 .0 3 1 5 135%

Data for Chapter 5 Experiment Day 18

8 t Q S t F l o w R a t e ( m A 3 / s e c ) V E L

I n n e r D i a m e t e r

1 / 4 i n c hR e #

o f f s e tf o r

m a m o m e te r

c m

V o l t a g e# 4

P r e s s u r eT r a n s d u c e

r

P r e s s u r eD r o p

( K P a )

A c t u a lf

P T 4

B l a s i u sE q -

P r e d i c t e df

P e r c e n tD i f f e r e n c e(Actual from Predcted)

20 3.5994 3.5994 0.113656 6.60335 OT 5b 0.951 0T l8 0 .2 3 2 8 0 .0 9 9 3 1 3 4 %3 0 5.0974 5 .0 9 /4 0.160958 0.00635 B T B 2.50 E T S i S 0.148 0 .1 4 5 5 0 .0 /0 1 1 0 7 %4 0 6.5954 6.5954 0 208259 0.00635 1181 2.50 0.962 0.200 0 .1 1 7 6 0 .0 5 4 2 117%5 0 8.0934 8.0934 0.255561 0.00635 1449 2.50 0.978 0.340 0 .1 3 2 4 0 .0 4 4 2 200%B O 9.5914 9.5914 0.302862 0.00635 1717 2.50 0.990 0.444 0 .1 2 3 3 0 .0 3 7 3 231%

7 0 11.0894 11.0894 0.350164 0.00635 1985 2.50 1.006 0.584 0 .1 2 1 2 0 .0 3 2 2 276%B O 12.5874 1 2 .58 /4 0 .3 9 /4 6 5 0.00635 2253 2.50 1.027 0.767 0 .1 2 3 6 0 .0 4 5 9 169%5 0 14.0854 14.0854 0 .4 4 4 /6 / 0.00635 2522 2.50 1.045 0.924 0 .1 1 8 9 0 .0 4 4 6 167%

T O O 15.5834 15.5834 0.492068 0.00635 2790 2.50 1.075 1.185 0 .1 2 4 6 0 .0 4 3 5 187%T T O j 1/.UB14 17.0814 0.53937 0.00635 3 0 5 8 2.50 1.095 1.360 0 .1 1 8 9 0 .0 4 2 5 180%T Z O - T E T B T S R 18.5794 0.586671 0.00635 3326 2.50 1.129 1.656 0 .1 2 2 5 0 .0 4 1 6 194%T 3 0 - 20.0774 2 0 .0 /7 4 0.633973 0.00635 3594 2.50 1.158 1.909 0 .1 2 0 9 0 .0 4 0 8 196%T 4 0 " 2 1 .5 /5 4 21.5754 0 .6 8 1 2 /4 0.00635 3863 2.50 1.195 2.231 0 .1 2 2 3 0 .0 4 0 1 205%T 5 0 " 2B.U/3 4 I 23.0734 0.728576 0.00635 4131 2.50 1.229 2.528 0 .1 2 1 2 0 .0 3 9 4 207%T B O - 2 4 .5 /1 4 2 4 .5 /1 4 0.775877 0.00635 4399 2.50 1.275 2.928 0 .1 2 3 8 0 .0 3 8 8 219%T T O - 26.0694 26.0694 0.8231 79 0.00635 4667 2.50 1.318 3.303 0 .1241 0 .0 3 8 2 225%T B B 27.5674 27.5674 0 .8 /0 4 8 0.00635 4935 2.50 1.362 3.687 0 .1 2 3 8 0 .0 3 7 7 228%T W 29.0654 29.0654 0.917782 0.00635 5203 2.50 1.408 4.088 0 .1 2 3 5 0 .0 3 7 2 2 3 2 %

Z O B 30.5634 30.5634 0.965083 0.00635 5472 2.50 1.462 4.558 0 .1 2 4 6 0 .0 3 6 7 2 3 5 %2 W 32.0614 32.0614 1.012385 0.00635 5740 2.50 1.510 3 T 3 7 7 0 .1 2 3 6 0 .0 3 6 3 240%Z Z B 3 3 . 5 5 9 4 33.5594 1.059686 0.00635 6008 2.50 1.559 5.404 0 .1 2 2 5 0 . 0 3 5 9 241%2 3 1 T 35.0574 35.0574 1.106988 0.00635 6276 2.50 1.620 5.935 0 .1 2 3 3 0 .0 3 5 5 247%2 4 B 36.5554 36.5554 1.154289 0.00635 6544 2.50 T T S G S 6.336 0 .1 2 1 0 0 .0 3 5 1 245%Z 5 B 38.0534 38.0534 1.201591 0. UlXxii 6813 2.50 1.727 6.868 0.1211 0 .0 3 4 8 248%275 41.'/984 41.7984 1.319844 0.00635 7483 2.50 1.890 8.289 0 .1211 0 .0 3 4 0 256%B O B 45.5434 45.5434 1 . 4 3 a U 9 H O.UlXxJb B T B T 2.50 2.060 9.7 70 0 .1 2 0 2 0 .0 3 3 3 262%B Z B 49.4884 49.2884 1.556352 0.00635 8824 2.50 2.240 11.339 0.1191 0 .0 3 2 6 265%3 5 B 53.0334 53.0334 1.674605 0.00635 9494 2.50 2.430 12.995 0 .1 1 7 9 0 .0 3 2 0 268%3 7 B 56.7784 56.7784 1.792859 0.00635 10165 2.50 2.560 14.128 0 .1 1 1 9 0 .0 3 1 5 255%

Data for Chapter 5 ExperimentDay 22

107

8 t Q StFlow R a te (m A3 /se c )

VELInner

D iam eter 1/4 inch

R e #

V oltage#4

P re ssu reT ra n sd u c e

r

P re ss u reDrop

(K Pa)

Actual fou led f

B lasiu sEq.

P red ic tedf

P e rc e n tD iffe rence(Actual from Predicted)

" Io 3 .5 9 9 4 3 .5 9 9 4 0.138234 ~ThtoB35 772 olSo U.109 0.1501 0.0829 -gnpj£

30 8 .0 9 /4 5 .0 9 /4 0 .1 9 2 9 3 2 0 .0 0 6 3 5 1094 0 .9 5 4 0.131 0 .0893 0 .0 5 8 5 53%30 6 .5 9 8 4 6 .5 9 5 4 0 .2 4 9 6 2 9 0 .0 0 6 3 5 1415 0 .962 0 .2 0 0 0 .0818 0.0452 81%50 8 .0 9 3 4 8 .0 9 3 4 0 .3 0 6 3 2 / 0 .0 0 6 3 5 1 /3 7 0 .9 /6 0 .3 2 2 0 .0 8 /4 0 .0 3 6 9 13/Vo50 9 .5 9 1 4 9 .5 9 1 4 0 .3 6 3 0 2 5 0 .0 0 6 3 5 2058 0 .9 9 5 0 .4 8 8 0 .0942 0.0311 2 0 3 Vo70 11 .0 8 9 4 11 .0894 0 .4 1 9 /2 3 0 .0 0 6 3 5 238 0 1.014 0 .6 5 4 0 .0 9 4 4 0 .0 2 6 9 251%50 12.58 /4 12 .5874 0 .4 /6 4 2 1 0 .0 0 6 3 5 2701 1.043 0 .906 0 .1016 0 .0 4 3 8 132%go 14 .0854 14 .0854 0 .5 3 3 1 1 8 0 .0 0 6 3 5 3 023 1 .0 /5 1.185 0.1061 0 .0 4 2 6 149%

TOO 15 .5 8 3 4 15 .5 8 3 4 0 .5 8 9 8 1 6 0 .0 0 6 3 5 3344 1 .112 1.508 0 .1103 0.0416 165%TTO 17 .0 8 1 4 17 .0 8 1 4 0 .6 4 6 5 1 4 0 .0 0 6 3 5 3 666 1.152 1.856 0 .1130 0 .0 4 0 6 178%TZO 1 8 .5 /9 4 1 8 .5 /9 4 0 ./0 3 2 1 2 O.Oueas 3987 1.197 2 .2 4 9 0 .1157 0 .0 3 9 8 191 VoTOO 2 0 .0 / / 4 2 0 .0 / / 4 0 ./5 9 9 1 0 .0 0 6 3 5 4 308 1.247 2 .684 0 .1183 0 .0 3 9 0 203%T30 2 1 .5 /5 4 2 1 .5 /5 4 0 .8 1 6 6 0 / 0 .0 0 6 3 5 4 630 1.303 3 .1 /3 0.1211 0.0383 216%T50 2 3 .0 /3 4 Z 3 .0 /3 4 0 .8 /3 3 0 5 0 .0 0 6 3 5 4951 1.372 3 . / / 4 0 .1259 0.0377 234%TOO 2 4 .5 /1 4 2 4 .5 /1 4 0 .9 3 0 0 0 3 0 .0 0 6 3 5 5 2 /3 1.433 4 .3 0 6 0 .1267 0.0371 242%TTO 2 6 .0 6 9 4 26 .0 6 9 4 0 .9 8 6 /0 1 0 .00 6 3 5 5594 1.495 4 .8 4 6 0 .1267 0 .0 3 6 5 247%TBO 2 / 5 6 / 4 2 / 5 6 / 4 1 .043399 0 .00 6 3 5 5 916 1 .5 /3 5 .526 0 .1292 0.0360 259%TOO 2 9 .0 6 5 4 2 9 .0 6 5 4 1 .1 0 0 0 9 / 0 .00 6 3 5 6 2 3 / 1.642 6 .1 2 / 0 .1289 0.0356 262%ZOO 3 0 .5 6 3 4 3 0 .5 6 3 4 1 .1 5 6 /9 4 0 .00 6 3 5 6 559 I . / 3 / 6 .9 5 5 0 .1323 0.0351 2 //V oZTO 3 2 .0 6 1 4 3 2 .0 6 1 4 1.213492 0 .00 6 3 5 6 880 1 .8 2 / / . / 4 0 0 .1338 0 .0 3 4 / 286%ZZO 3 3 .5 5 9 4 3 3 .5 5 9 4 1 .2 /0 1 9 0 .00 6 3 5 /2 0 2 1.913 8 .489 0 .1339 0 .0 3 4 3 290%ZOO 3 5 .0 5 /4 3 5 .0 5 /4 1 .326888 0 .00 6 3 5 /5 2 3 2 .0 0 0 9 .2 4 8 0 .1 3 3 / 0 .0 3 3 9 294%Z30 3 6 .5 5 5 4 3 6 .5 5 5 4 1.38 3 5 8 6 0 .00 6 3 5 /8 4 4 2 .100 10.119 0.1345 0 .0 3 3 6 3 0 1 Vo250 3 8 .0 5 3 4 3 8 .0 5 3 4 1 .440283 0 .0 0 6 3 5 8166 2 .1 9 0 10.904 0.1338 0 .0 3 3 2 302%Z75 4 1 / 9 8 4 4 1 / 9 8 4 1.582028 0 .0 0 6 3 5 8 9 /0 2 .4 0 0 12 . /3 4 0.1295 0 .0 3 2 5 299%300 4 5 .5 4 3 4 4 5 .5 4 3 4 I . / 2 3 / / 2 O.QOBSb 9 / / 3 2 .6 8 0 15.1 /4 0 .1300 0 .0318 309%

Data for Chapter 5 Experiment Day 26

Bt QStFlow R a te (m A3/sec) V E L

Inner D iam eter 1/4 inch

R e #

V oltage#4

P re ssu reT ran sd u ce

r

P re ssu reD rop(K Pa)

Actual f Day 26

BlasiusEq-

P red ic tedf

P e rce n t D ifference ( A c tu a l f r o m

P r e d i c t e d )

' 20 3 .5994 3 .5994 0.1 T l 652 0 .00636 BOB 6.952 0 .113 0.1438 0 .0 / 9 7 " 5 0 %30 5 .0 9 /4 5 .0 9 /4 0 .200463 0 .00635 1137 0 .960 0 .183 0.1159 0 .0563 106%TO 6.5954 6 .5954 0.259374 0.00635 1471 0 .968 0 .253 0.0956 0 .0435 120%50 S.U9S4 8 .0934 0.318286 0.00635 T805 0 .9 /6 0 .322 0.0810 0 .0355 1 28%BO 9.5 9 1 4 9 .5914 0 .377197 0.00635 2139 0 .986 0 .410 0 .0 /3 3 0 .0299 145%70 11.0894 11.0894 0.436108 0.00635 2473 1.000 0 .532 0.0711 0 .0259 175%80 12.5874 12.5874 0.495019 0.00635 2807 1.015 0 .662 0.0688 0 .0434 5 5 %go 14.US54 14.0854 0.55393 0.UUS3S 3141 1.039 0 .872 0.0723 0 .0422 7T%

TOO 15.5834 15.5834 0 .612842 0.00635 3475 1 .0 /4 1.177 0.0797 0 .0412 94%TTO 17.0814 17.0814 0 .6 /1 7 5 3 0.00635 3809 1.103 1.429 0.0806 0 .0402 100%TZO 18.5794 1 8 .5 /9 4 0 ./3 0 6 6 4 0.00635 4143 1.146 1.804 0.0860 0 .0 3 9 4 118%TOO 20.0774 2 0 .0 7 /4 0. /8 9 5 /5 0 .00635 4 4 7 / 1 .1 8 / 2.161 0.0882 0 .0386 128%TTO 21.5754 21 .5754 0.848486 0.00635 4811 TT227 ZTBTff 0.0887 O T B B T B T33%T50 23 .0 7 3 4 23 .0734 0 .9 0 /3 9 / 0 .00635 5145 1.286 3.024 0.0935 0 .0373 151%TBO 24.8714 2 4 .5 /1 4 0 .966309 0.00635 5479 1.336 3.460 0.0943 0 .0367 157%TTO 26 .0694 26 .0694 1 .02522 0.00635 5813 1.396 BTBBB 0.0964 0 .0362 166%TBO 2 7 .5 6 /4 2 7 .5 6 /4 1.084131 0.00635 6147 1.466 TBBB 0.0995 0.0357 179%TOO 29.0654 29 .0664 1.143042 0.00635 6481 1.519 5.055 0.0985 0 .0352 180%ZOO 30.5 6 3 4 30.5634 1 .201953 0.00635 6815 1.SB4 5.622 0.0990 0 .0348 185%ZTO 32.0 6 1 4 32.0614 1 .260865 0.00635 7T39 TTBBff 6 .284 0.1006 0.0344 193%ZZOj 33 .5 5 9 4 33.5594 1 .3 1 9 /7 6 0.00635 /4 8 3 1.734 6.929 0.1012 0 .0340 198%Z30" 35 .0574 35.0574 1.378687 0.00635 / 8 1 / 1 .8 0 / 7.565 0.1013 0.0336 201%ZTO- 36 .5554 36 .5554 1 .437598 0.00635 8151 1.884 8.236 0.1014 0.0333 205%Z 5 a 38 .0 5 3 4 38.0534 1.4965U9 0.00635 8485 1.967 8.960 0.1018 0.0329 209%275 41 .7 9 8 4 41 .7984 1 .6 4 3 /8 7 0.00635 9320 2.180 10.816 0.1019 0.0322 217%3 0 0 45 .5 4 3 4 45 .5434 1.791065 0.00635 10155 2 .3 /0 12.472 0.0990 0.0315 214%325 49 .2 8 8 4 49 .2 8 8 4 1.938343 0.00635 10990 2.600 14.477 0.0981 0.0309 218%350“ 53 .0334 53 .0334 2.085621 0 .00635 11825 2.850 16.656 0.0975 0.0303 222%

108

Data for Chapter 5 ExperimentDay 28

081 Flow Rate (m'‘S/sec) VEL

Inner Diameter 1/4 inch

R e#

Voltage# 4

PressureTransduce

r

PressureDrop(KPa)

Actual f Day 32

BlasiusEq-

Predictedf

Percent Difference

( A c t u a l f r o m P r e d i c t e d )

3 . 5 6 9 4 5 . 1 5 4 7 4 7 0 . 0 0 6 3 5 # 7 0 . 9 5 7 0 . 1 5 7 0.1667 0.0729 i 28%5 . 0 9 7 4 5 . 0 9 7 4 0 . 2 1 9 1 5 0 . 0 0 6 3 5 I W 0 . 9 6 1 0 . 1 9 2 0.1016 0.0515 97%6 . 5 9 5 4 6 . 5 9 5 4 0 . 2 8 3 5 5 3 0 . 0 0 6 3 5 TBM 0 . 9 6 8 0 . 2 5 3 0.O8OO 0.0398 101%8 . 0 9 3 4 8 . 0 9 3 4 0 . 3 4 7 9 5 6 0 . 0 0 6 3 5 T W 0 . 9 8 4 0 . 3 9 2 0.0824 0.0324 154%9 . 5 9 1 4 9 . 5 9 1 4 0 . 4 1 2 3 5 9 0 . 0 0 6 3 5 2 3 3 8 1 . 0 0 0 0 . 5 3 2 0.0796 0.0274 191%

1 1 . 0 8 9 4 1 1 . 0 8 9 4 0 . 4 7 6 7 6 2 0 . 0 0 6 3 5 2 7 0 3 1 . 0 2 1 0 . 7 1 5 0.0800 0.0237 238%1 2 . 5 8 7 4 1 2 . 5 8 7 4 0 . 5 4 1 1 6 5 0 . 0 0 6 3 5 3 0 6 8 1 . 0 4 4 0 . 9 1 5 0.0795 0.0425 87%1 4 . 0 8 5 4 1 4 . 0 8 5 4 0 . 6 0 5 5 6 8 0 . 0 0 6 3 5 3 4 3 3 1 . 0 8 4 1 . 2 6 4 0.0877 0.0413 112%1 5 . 5 8 3 4 1 5 . 5 8 3 4 0 . 6 6 9 9 7 0 . 0 0 6 3 5 3 7 9 8 1 . 1 4 0 1 . 7 5 2 0.0993 0.0403 147%1 7 . 0 8 1 4 1 7 . 0 8 1 4 0 . 7 3 4 3 7 3 0 . 0 0 6 3 5 4 1 6 4 1 . 1 9 6 2 . 2 4 0 0.1057 0.0393 169%1 8 . 5 7 9 4 1 8 . 5 7 9 4 0 . 7 9 8 7 7 6 0 . 0 0 6 3 5 4 5 2 9 1 . 2 6 0 2 . 7 9 8 0.1116 0.0385 190%

2 0 . 0 7 7 4 2 0 . 0 7 7 4 0 . 8 6 3 1 7 9 0 . 0 0 6 3 5 1 . 3 2 5 3 . 3 6 4 0.1149 0.0378 204%2 1 . 5 7 5 4 2 1 . 5 7 5 4 0 . 9 2 7 5 8 2 0 . 0 0 6 3 5 5 2 5 9 1 . 4 0 5 4 . 0 6 2 0.1201 0.0371 224%2 3 . 0 7 3 4 2 3 . 0 7 3 4 0 . 9 9 1 9 8 5 0 . 0 0 6 3 5 5 6 2 4 1 . 4 8 0 4 . 7 1 5 0.1220 0.0365 234%2 4 . 5 7 1 4 2 4 . 5 7 1 4 1 . 0 5 6 3 8 8 0 . 0 0 6 3 5 5 9 8 9 1 . 5 7 1 5 . 5 0 8 0.1256 0.0359 250%2 6 . 0 6 9 4 2 6 . 0 6 9 4 1 . 1 2 0 7 9 1 0 . 0 0 6 3 5 6 3 5 4 1 . 6 6 5 6 . 3 2 8 0.1282 0.0354 262%2 7 . 5 6 7 4 2 7 . 5 6 7 4 1 . 1 8 5 1 9 3 0 . 0 0 6 3 5 6 7 2 0 1 . 7 5 5 7 . 1 1 2 0.1289 0.0349 269%2 9 . 0 6 5 4 2 9 . 0 6 5 4 1 . 2 4 9 5 9 6 0 . 0 0 6 3 5 7 0 5 5 TBBT 8 . 0 0 1 0.1304 0.0344 279%3 0 . 5 6 3 4 3 0 . 5 6 3 4 1 . 3 1 3 9 9 9 0 . 0 0 6 3 5 7 4 5 0 " 1 . 9 6 0 8 . 8 9 9 0.1312 0.0340 286%3 2 . 0 6 1 4 3 2 . 0 6 1 4 1 . 3 7 8 4 0 2 0 . 0 0 6 3 5 7 8 1 5 2 . 0 6 0 9 . 7 7 0 0.1309 0.0336 289%3 3 . 5 5 9 4 3 3 . 5 5 9 4 1 . 4 4 2 8 0 5 0 . 0 0 6 3 5 BTM 2 . 1 8 0 1 0 . 8 1 6 0.1322 0.0332 298%3 5 . 0 5 7 4 3 5 . 0 5 7 4 1 . 5 0 7 2 0 8 0 . 0 0 6 3 5 8 5 4 5 2 . 2 9 0 1 1 . 7 7 5 0.1319 0.0329 301%3 6 . 5 5 5 4 3 6 . 5 5 5 4 1 . 5 7 1 6 1 1 0 . 0 0 6 3 5 S W 2 . 4 2 0 1 2 . 9 0 8 0.1330 0.0325 309%3 8 . 0 5 3 4 3 8 . 0 5 3 4 1 6 3 6 0 1 3 0 . 0 0 6 3 5 9 2 7 6 2 . 5 3 0 1 3 . 8 6 7 0.1319 0.0322 310%4 1 . 7 9 8 4 4 1 . 7 9 8 4 1 . 7 9 7 0 2 1 0 . 0 0 6 3 5 1 0 1 8 8 2 . 9 0 0 1 7 . 0 9 2 0.1347 0.0315 328%4 5 . 5 4 3 4 4 5 . 5 4 3 4 1 . 9 5 8 0 2 8 0 . 0 0 6 3 5 1 1 1 0 1 3 . 1 7 0 1 9 . 4 4 5 0.1291 0.0308 319%

STATE - BOZEMAN