a study of non-newtonian viscosit y and yield stress of

236
A Study of Non-Newtonian Viscosity and Yield Stress of Blood in a Scanning Capillary-Tube Rheometer A Thesis Submitted to the Faculty of Drexel University by Sangho Kim in partial fulfillment of the requirements for the degree of Doctor of Philosophy December 2002

Upload: others

Post on 15-Oct-2021

7 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: A Study of Non-Newtonian Viscosit y and Yield Stress of

A Study of Non-Newtonian Viscosity and Yield Stress of Blood

in a Scanning Capillary-Tube Rheometer

A Thesis

Submitted to the Faculty

of

Drexel University

by

Sangho Kim

in partial fulfillment of the

requirements for the degree

of

Doctor of Philosophy

December 2002

Page 2: A Study of Non-Newtonian Viscosit y and Yield Stress of

ii

Acknowledgments

I wish to express my sincere gratitude to Dr. Young I. Cho, for his guidance

and inspiration during my entire tenure in graduate school. His experience and idea

have proven to be invaluable. I also wish to thank Dr. David M. Wootton for serving

as my co-advisor, and for his valuable suggestions and guidance on Biofluid

Dynamics.

I wish to express my appreciation to the members of my dissertation

committee, including: Dr. Ken Choi and Dr. Alan Lau from the MEM Department,

and Dr. Peter Lelkes from the School of Biomedical Engineering.

I am deeply indebted to Dr. Kenneth Kensey, Mr. William Hogenauer, and

Dr. Larry Goldstein from Rheologics, Inc. for providing valuable comments on the

test methods and data reduction procedure.

A sincere appreciation is extended to several colleagues whose friendship I

have cherished during my graduate studies, including: Dr.Wontae Kim, Dr. Sunghyuk

Lee, Chagbeom Kim, Giyoung Tak, Dohyung Lim, and Jinyong Wee.

Last but not least, I wish to thank my parents for their unbounded support

throughout my life. Their reliable provision of emotional, spiritual, and financial

support has allowed me to accomplish tasks that would have otherwise been

impossible.

Page 3: A Study of Non-Newtonian Viscosit y and Yield Stress of

iii

Table of Contents

LIST OF TABLES.....................................................................................................viii

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

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

CHAPTER 1 INTRODUCTION .............................................................................. 1

1.1 Clinical Significance of Blood Viscosity.................................................... 1

1.2 Motivation of the Present Study ................................................................. 3

1.3 Objectives of the Present Study .................................................................. 3

1.4 Outline of the Dissertation .......................................................................... 4

CHAPTER 2 CONSTITUTIVE MODELS.............................................................. 5

2.1 Newtonian Fluid.......................................................................................... 5

2.2 Non-Newtonian Fluid ................................................................................. 10

2.2.1 General Non-Newtonian Fluid........................................................... 10

2.2.1.1 Power-law Model...................................................................... 11

2.2.1.2 Cross Model .............................................................................. 12

2.2.2 Viscoplastic Fluid .............................................................................. 13

2.2.2.1 Bingham Plastic Model............................................................. 13

2.2.2.2 Casson Model............................................................................ 14

2.2.2.3 Herschel-Bulkley Model........................................................... 15

2.3 Rheology of Blood...................................................................................... 19

2.3.1 Determination of Blood Viscosity ..................................................... 19

Page 4: A Study of Non-Newtonian Viscosit y and Yield Stress of

iv

2.3.1.1 Plasma Viscosity....................................................................... 20

2.3.1.2 Hematocrit................................................................................. 20

2.3.1.3 RBC Deformability................................................................... 21

2.3.1.4 RBC Aggregation - Major Factor of Shear-Thinning Characteristics........................................................................... 21

2.3.1.5 Temperature .............................................................................. 22

2.3.2 Yield Stress and Thixopropy ............................................................. 23

2.3.2.1 Yield Stress ............................................................................... 23

2.3.2.2 Thixotropy - Time Dependence ................................................ 24

CHAPTER 3 CONVENTIONAL RHEOMETRY: STATE-OF-THE-ART ........... 30

3.1 Introduction................................................................................................. 30

3.2 Rotational Viscometer ................................................................................ 34

3.2.1 Rotational Coaxial-Cylinder (Couette Type)..................................... 34

3.2.2 Cone-and-Plate................................................................................... 35

3.3 Capillary-Tube Viscometer......................................................................... 38

3.4 Yield Stress Measurement .......................................................................... 41

3.4.1 Indirect Method.................................................................................. 42

3.4.1.1 Direct Data Extrapolation ......................................................... 42

3.4.1.2 Extrapolation Using Constitutive Models................................. 43

3.4.2 Direct Method .................................................................................... 44

3.5 Problems with Conventional Viscometers for Clinical Applications ......... 46

3.5.1 Problems with Rotational Viscometers.............................................. 46

3.5.2 Problems with Capillary-Tube Viscometers...................................... 48

Page 5: A Study of Non-Newtonian Viscosit y and Yield Stress of

v

CHAPTER 4 THEORY OF SCANNING CAPILLARY-TUBE RHEOMETER.... 49

4.1 Scanning Capillary-Tube Rheometer (SCTR) ............................................ 49

4.1.1 U-Shaped Tube Set ............................................................................ 50

4.1.2 Energy Balance .................................................................................. 51

4.2 Mathematical Procedure for Data Reduction.............................................. 60

4.2.1 Power-law Model............................................................................... 60

4.2.2 Casson Model..................................................................................... 66

4.2.3 Herschel-Bulkley (H-B) Model ......................................................... 72

CHAPTER 5 CONSIDERATIONS FOR EXPERIMENTAL STUDY................... 81

5.1 Unsteady Effect ........................................................................................... 82

5.2 End Effect.................................................................................................... 87

5.3 Wall Effect (Fahraeus-Lindqvist Effect) ..................................................... 90

5.4 Other Effects................................................................................................ 95

5.4.1 Pressure Drop at Riser Tube .............................................................. 95

5.4.2 Effect of Density Variation................................................................ 96

5.4.3 Aggregation Rate of RBCs - Thixotropy........................................... 97

5.5 Temperature Considerations for Viscosity Measurement of Human Blood..........................................................................................101 5.6 Effect of Dye Concentration on Viscosity of Water ...................................104

5.6.1 Introduction........................................................................................104

5.6.2 Experimental Method.........................................................................106

5.6.3 Results and Discussion ......................................................................107

CHAPTER 6 EXPERIMENTAL STUDY WITH SCTR.........................................112

6.1 Experiments with SCTR (with Precision Glass Riser Tubes) ....................112

Page 6: A Study of Non-Newtonian Viscosit y and Yield Stress of

vi

6.1.1 Description of Instrument ..................................................................113

6.1.2 Testing Procedure ..............................................................................114

6.1.3 Data Reduction with Power-law Model.............................................116

6.1.4 Results and Discussion ......................................................................117

6.2 Experiments with SCTR (with Plastic Riser Tubes)...................................130

6.2.1 Description of Instrument ..................................................................131

6.2.2 Testing Procedure ..............................................................................132

6.2.3 Data Reduction with Casson Mocel...................................................133

6.2.3.1 Curve Fitting .............................................................................134

6.2.3.2 Results and Discussion .............................................................135

6.2.4 Data Reduction with Herschel-Bulkley (H-B) Model .......................139

6.3 Comparison of Non-Newtonian Constitutive Models ................................158

6.3.1 Comparison of Viscosity Results.......................................................159

6.3.2 Comparison of Yield Stress Results ..................................................162

6.3.3 Effects of Yield Stress on Flow Patterns ...........................................164

CHAPTER 7 CONCLUSIONS AND RECOMMENDATIONS .............................180

LIST OF REFERENCES...........................................................................................184

APPENDIX A: NOMENCLATURE.........................................................................194

APPENDIX B: FALLING OBJECT VISCOMETER - LITERATURE REVIEW..............................................................197 APPENDIX C: SPECIFICATION OF CCD AND LED ARRAY............................200

APPENDIX D: BIOCOATING OF CAPILLARY TUBE........................................202

APPENDIX E: MICROSOFT EXCEL SOLVER.....................................................204

APPENDIX F: NEWTON’S METHOD OF ITERATION.......................................206

Page 7: A Study of Non-Newtonian Viscosit y and Yield Stress of

vii

APPENDIX G: REPEATABILITY STUDY WITH DISTILLED WATER ............208

APPENDIX H: EXPERIMENTAL DATA...............................................................210

VITA..........................................................................................................................221

Page 8: A Study of Non-Newtonian Viscosit y and Yield Stress of

viii

List of Tables

2-1. Viscosity of some familiar materials at room temperature............................... 8 2-2. Range of shear rates of some familiar materials and processes ........................ 9 5-1. Comparison of unsteadyP∆ and cP∆ for distilled water ........................................ 84 5-2. Comparison of unsteadyP∆ and cP∆ for bovine blood .......................................... 86 5-3. Density estimation............................................................................................. 99 6-1. Comparison of initial guess and resulting value using power-law model.........124 6-2. Comparison of initial guess and resulting value using Casson model ..............144 6-3. Comparison of initial guess and resulting value using Herschel-Bulkley model ..........................................................................155 6-4. Comparison of four unknowns determined with Herschel-Bulkley model for three consecutive tests..................................................................................157 6-5. Various physiological studies with non-Newtonian constitutive models .........167 6-6. Measurements of water viscosity ......................................................................169 6-7. Measurements of bovine blood viscosity ..........................................................171 6-8. Measurements of human blood viscosity ..........................................................173 6-9. Comparison of model constants, yh∆ and yτ ...................................................175 6-10. Comparison of ∞=∆ th and yst hh ∆+∆ ..............................................................176 H-1. A typical experimental data set of human blood obtained by a scanning capillary-tube rheometer with precision glass riser tubes.................................210 H-2. A typical experimental data set of distilled water obtained by a scanning capillary-tube rheometer with plastic riser tubes..............................................213 H-3. A typical experimental data set of bovine blood obtained by a scanning capillary-tube rheometer with plastic riser tubes..............................................215

Page 9: A Study of Non-Newtonian Viscosit y and Yield Stress of

ix

H-4. A typical experimental data set of human blood obtained by a scanning capillary-tube rheometer with plastic riser tubes..............................................218

Page 10: A Study of Non-Newtonian Viscosit y and Yield Stress of

x

List of Figures

2-1. Flow curve of a Newtonian fluid....................................................................... 7 2-2. Flow curve of power-law fluids......................................................................... 16 2-3. Flow curve of a Casson model .......................................................................... 17 2-4. Flow curve of viscoplastic fluids....................................................................... 18 2-5. Comparison of Newtonian plasma viscosity and shear-thinning whole blood viscosity ............................................................... 26 2-6. Variation of the relative viscosity of blood and suspension with rigid spheres at a shear rate > 100 s-1 ..................................................................................... 27 2-7. Rouleaux formation of human red blood cells photographed on a microscope slide showing single linear and branched aggregates and a network................ 28 2-8. Elevated blood viscosity at low shear rates indicates RBC aggregation........... 29 3-1. Rheometers ........................................................................................................ 33 3-2. Schematic diagram of a concentric cylinder viscometer ................................... 36 3-3. Schematic diagram of a con-and-plate viscometer............................................ 37 3-4. Schematic diagram of a capillary-tube viscometer............................................ 40 3-5. Determination of yield stress by extrapolation.................................................. 45 4-1. Schematic diagram of a U-shaped tube set........................................................ 56 4-2. Fluid-level variation in a U-shaped tube set during a test ................................. 57 4-3. Typical fluid-level variation measured by a SCTR........................................... 58 4-4. Liquid-solid interface condition for each fluid column of a U-shaped tube set........................................................................................ 59 4-5. Fluid element in a capillary tube at time t ........................................................ 79 4-6. Velocity profile of plug flow of blood in a capillary tube................................. 80

Page 11: A Study of Non-Newtonian Viscosit y and Yield Stress of

xi

5-1. Pressure drop estimation for distilled water ...................................................... 83 5-2. Pressure drop estimation for bovine blood ........................................................ 85 5-3. Flow-pattern changes due to end effects ........................................................... 89 5-4. Migration of cells toward to the center of lumen (wall effect).......................... 92 5-5. Fahraeus-Lindquist effect due to the reduction in hematocrit in a tube with a small diameter and the tendency of erythrocytes to migrate toward the center of the tube......................................................................................... 93 5-6. Viscosity measurements for bovine blood with three different capillary tubes with ID of 0.797 mm (with length = 100 mm), 1.0 mm (with length = 130 mm), and 1.2 mm (with length = 156 mm) ................................................................ 94 5-7. Viscosity results for human blood with two different capillary tubes with length of 100 mm (with ID = 0.797 mm) and 125 mm (ID = 0.797 mm) ........100 5-8. Schematic diagram of a U-shaped tube set for temperature measurement........102 5-9. Temperature measurement at a capillary tube during a viscosity test ...............103 5-10. Schematic diagram of a scanning capillary-tube rheometer (SCTR) system................................................................................109 5-11. Variations of both power-law index and consistency index of dye-water solution due to effects of dye concentrations...................................................110 5-12. Viscosity data for dye-water solution with 6 different dye concentrations at 25..............................................................................................................111 6-1. Schematic diagram of a scanning capillary-tube rheometer with precision glass riser tubes .........................................................................121 6-2. Curve-fitting procedure with power-law model for mineral oil ........................122 6-3. Curve-fitting procedure with power-law model for human blood ....................123 6-4. Height variation in each riser tube vs. time for mineral oil ...............................125 6-5. Viscosity measurement for mineral oil at 25 with a scanning capillary-tube rheometer (SCTR) .....................................................................126 6-6. Height variation in each riser tube vs. time for human blood at 37. .............127

Page 12: A Study of Non-Newtonian Viscosit y and Yield Stress of

xii

6-7. Viscosity measurement (log-log scale) for human blood at 37 with rotating viscometer (RV) and scanning capillary-tube rheometer (SCTR) ......128 6-8. Viscosity measurement (log-log scale) of unadulterated human blood at 37, measured with scanning capillary-tube rheometer (SCTR) and cone-and-plate rotating viscometer (RV), for two different donors .................129 6-9. Picture of a SCTR with plastic riser tubes.........................................................141 6-10. Heating pad for a test with unadulterated human blood...................................142 6-11. Curve-fitting procedure with Casson model for distilled water .......................143 6-12. Curve-fitting procedure with Casson model for donor 1..................................145 6-13. Curve-fitting procedure with Casson model for donor 2..................................146 6-14. Height variation in each riser tube vs. time for distilled water at 25............147 6-15. Viscosity measurement for distilled water at 25 ..........................................148 6-16. Height variation in each riser tube vs. time for bovine blood with 7.5% EDTA at 25.................................................................................149 6-17. Viscosity measurement for bovine blood with 7.5% EDTA at 25 using both rotating viscometer (RV) and scanning capillary-tube Rheometer (SCTR) ..........................................................................................150 6-18. Height variation in each riser tube vs. time for human blood at 37 .............151 6-19. Viscosity measurement for human blood (2 different donors) at 37 ............152 6-20. Shear-stress variation vs. shear rate for human blood (from 2 different donors) at 37.....................................................................153 6-21. Curve-fitting procedure with Herschel-Bulkley model for bovine blood ........154 6-22. Viscosity measurements of bovine blood with 7.5% EDTA at 25, analyzed with Herschel-Bulkley model.............................................156 6-23. Test with distilled water at 25.......................................................................168 6-24. Test with bovine blood at 25 ........................................................................170

Page 13: A Study of Non-Newtonian Viscosit y and Yield Stress of

xiii

6-25. Test with unadulterated human blood at 37..................................................172 6-26. Wall shear stress at a capillary tube vs. shear rate............................................174 6-27. Variations of a plug-flow region at a capillary tube as a function of time for bovine blood with 7.5% EDTA at 25 .....................................................177 6-28. Velocity profiles at a capillary tube for bovine blood with 7.5% EDTA at 25.................................................................................178 6-29. (a) Viscosity, (b) wall shear rate, and (c) wall shear stress Plotted as a function of mean velocity at a capillary tube using three non-Newtonian models for bovine blood with 7.5% EDTA ..................179 B-1. Falling cylinder viscometers .............................................................................199 C-1. Cross sectional view of SV352A8-01 module..................................................201 G-1. Repeatability study #1 ......................................................................................208 G-2. Repeatability study #2 ......................................................................................209

Page 14: A Study of Non-Newtonian Viscosit y and Yield Stress of

xiv

Abstract

A Study of Non-Newtonian Viscosity and Yield Stress of Blood in a Scanning Capillary-Tube Rheometer

Sangho Kim Professors Young I. Cho and David M. Wootton

The study of hemorheology has been of great interest in the fields of

biomedical engineering and medical researches for many years. Although a number

of researchers have investigated correlations between whole blood viscosity and

arterial diseases, stroke, hypertension, diabetes, smoking, aging, and gender, the

medical community has been slow in realizing the significance of the whole blood

viscosity, which can be partly attributed to the lack of an uncomplicated and clinically

practical rheometer.

The objectives of the present study were to investigate the theoretical

principles of a scanning capillary-tube rheometer used for measuring both the

viscosity and yield stress of blood without any anticoagulant, to experimentally

validate the scanning capillary-tube rheometer using disposable tube sets designed for

daily clinical use in measuring whole blood viscosity, and to investigate the effect of

non-Newtonian constitutive models on the blood rheology and flow patterns in the

scanning capillary-tube rheometer.

The present study introduced detailed mathematical procedures for data

reduction in the scanning capillary-tube rheometer for both viscosity and yield-stress

measurements of whole blood. Power-law, Casson, and Herschel-Bulkley models

were examined as the constitutive models for blood in the study. Both Casson and

Herschel-Bulkley models gave blood viscosity results which were in good agreement

Page 15: A Study of Non-Newtonian Viscosit y and Yield Stress of

xv

with each other as well as with the results obtained by a conventional rotating

viscometer, whereas the power-law model seemed to produce inaccurate viscosities at

low shear rates.

The yield stress values obtained from the Casson and Herschel-Bulkley

models for unadulterated human blood were measured to be 13.8 and 17.5 mPa,

respectively. The two models showed some discrepancies in the yield-stress values.

In the study, the wall shear stress was found to be almost independent of the

constitutive model, whereas the size of the plug flow region in the capillary tube

varies substantially with the selected model, altering the values of the wall shear rate

at a given mean velocity. The model constants and the method of the shear stress

calculation given in the study can be useful in the diagnostics and treatment of

cardiovascular diseases.

Page 16: A Study of Non-Newtonian Viscosit y and Yield Stress of

1

CHAPTER 1. INTRODUCTION

1.1. Clinical Significance of Blood Viscosity

The study of hemorheology has been of great interest in the fields of

biomedical engineering and medical research for many years. Hemorheology plays

an important role in atherosclerosis [Craveri et al., 1987; Resch et al., 1991; Lee et al.,

1998; Kensey and Cho, 2001]. Hemorheological properties of blood include whole

blood viscosity, plasma viscosity, hematocrit, RBC deformability and aggregation,

and fibrinogen concentration in plasma. Although a number of parameters such as

pressure, lumen diameter, whole blood viscosity, compliance of vessels, peripheral

vascular resistance are well-known physiological parameters that affect the blood

flow, the whole blood viscosity is also an important key physiological parameter.

However, its significance has not been fully appreciated yet.

A number of researchers measured blood viscosities in patients with coronary

arterial disease such as ischemic heart disease and myocardial infarction [Jan et al.,

1975; Lowe et al., 1980; Most et al., 1986; Ernst et al., 1988; Rosenson, 1993]. They

found that the viscosity of whole blood might be associated with coronary arterial

diseases. In addition, a group of researchers reported that whole blood viscosity was

significantly higher in patients with peripheral arterial disease than that in healthy

controls [Ciuffetti et al., 1989; Lowe et al., 1993; Fowkes et al., 1994].

Page 17: A Study of Non-Newtonian Viscosit y and Yield Stress of

2

Other researchers investigated correlation between the hemorheological

parameters and stroke [Grotta et al., 1985; Coull et al., 1991; Fisher and Meiselman,

1991; Briley et al., 1994]. They reported that stroke patients showed two or more

elevated rheological parameters, which included whole blood viscosity, plasma

viscosity, red blood cell (RBC) and plate aggregation, RBC rigidity, and hematocrit.

It was also reported that both whole blood viscosity and plasma viscosity were

significantly higher in patients with essential hypertension than in healthy people

[Letcher et al., 1981, 1983; Persson et al., 1991; Sharp et al., 1996; Tsuda et al., 1997;

Toth et al., 1999]. In diabetics, whole blood viscosity, plasma viscosity, and

hematocrit were elevated, whereas RBC deformability was decreased [Hoare et al.,

1976; Dintenfass, 1977; Hill et al., 1982; Poon et al., 1982; Leiper et al., 1982].

Others conducted hemorheological studies to determine the relationships

between whole blood viscosity and smoking, age, and gender [Levenson et al., 1987;

Bowdler and Foster, 1987; Fowkes et al., 1994; Ernst, 1995; Ajmani and Rifkind,

1998; Kameneva et al., 1998; Yarnell et al., 2000]. They found that smoking and

aging might cause the elevated blood viscosity. In addition, it was reported that male

blood possessed higher blood viscosity, RBC aggregability, and RBC rigidity than

premenopausal female blood, which may be attributed to monthly blood-loss

[Kameneva et al., 1998].

Page 18: A Study of Non-Newtonian Viscosit y and Yield Stress of

3

1.2. Motivation of the Present Study

The medical community has been slow in realizing the significance of whole

blood viscosity, which can be attributed partly to the lack of an uncomplicated and

clinically practical method of measuring whole blood viscosity. In most clinical

studies, mainly two types of viscometer have been available for general use:

rotational viscometers and capillary tube viscometer, as will be discussed in Chapter

3. These viscometers are used at laboratory only, and are not used in a clinical

environment. Until recently, the most immediate difficulty has been the lack of an

instrument that is specially designed for daily clinical use in measuring whole blood

viscosity.

1.3. Objectives of the Present Study

The objectives of the present study were 1) to investigate the theoretical

principles of a scanning capillary-tube rheometer (SCTR), which is capable of

measuring the viscosity and yield stress of blood without adding any anticoagulant, 2)

to validate the SCTR using disposable tube sets for clinical applications, and 3) to

investigate the effect of non-Newtonian constitutive models on the blood rheology

and flow patterns in the SCTR.

The present study introduced detailed mathematical procedures for data

reduction in the SCTR for both viscosity and yield-stress measurements of blood. In

Page 19: A Study of Non-Newtonian Viscosit y and Yield Stress of

4

experimental studies, distilled water (Newtonian fluid), bovine blood (non-Newtonian

fluid) with 7.5% EDTA, and unadulterated human blood (non-Newtonian fluid) were

used for the measurements of both viscosity and yield stress. Power-law, Casson, and

Herschel-Bulkley models were examined as constitutive models for blood in the study.

1.4. Outline of the Dissertation

Chapter 2 reviews the constitutive models applicable for non-Newtonian

characteristics including shear-thinning and yield stress. Chapter 3 reviews the

conventional rheometers that measure either the viscosity or yield stress of a fluid. In

this chapter, only rheometers that can be applicable to clinical applications are

discussed. Chapter 4 introduces the theory of a scanning capillary-tube rheometer.

Chapter 5 discusses the considerations for the experimental study, which include

unsteady effect, end effect, wall effect, temperature analysis, dye concentration effect,

and other possible factors. Chapter 6 presents the results of experimental studies

performed with a scanning capillary-tube rheometer. Chapter 6 also reports the effect

of non-Newtonian constitutive models on the rheological measurements and flow

patterns of blood in a capillary tube. Chapter 7 gives conclusions of the study and

recommendations for future study.

Page 20: A Study of Non-Newtonian Viscosit y and Yield Stress of

5

CHAPTER 2. CONSTITUTIVE MODELS

This chapter reviews literature on non-Newtonian constitutive models, which

are applicable to the study of blood rheology. Viscous liquids including whole blood

can be divided in terms of rheological properties into Newtonian, general non-

Newtonian, and viscoplastic fluids. The characteristics of blood, which include

shear-thinning, yield stress, and thixotropy, are discussed in this chapter.

2.1. Newtonian Fluid

Fluid such as water, air, ethanol, and benzene are Newtonian. This means that

when shear stress is plotted against shear rate at a given temperature, the plot shows a

straight line with a constant slope that is independent of shear rate (see Fig. 2-1).

This slope is called the viscosity of the fluid. All gases are Newtonian, and common

liquids such as water and glycerin are also Newtonian. Also, low molecular weight

liquids and solutions of low molecular weight substances in liquids are usually

Newtonian. Some examples are aqueous solutions of sugar or salt.

The simplest constitutive equation is Newton’s law of viscosity [Middleman,

1968; Bird et al., 1987; Munson et al., 1998]:

γµτ &= (2-1)

where µ is the Newtonian viscosity and γ& is the shear rate or the rate of strain.

Page 21: A Study of Non-Newtonian Viscosit y and Yield Stress of

6

The Newtonian fluid is the basis for classical fluid mechanics. Gases and

liquids like water and mineral oils exhibit characteristics of Newtonian viscosity.

However, many important fluids, such as blood, polymers, paint, and foods, show

non-Newtonian viscosity.

Table 2-1 shows the wide viscosity range for common materials. Different

instruments are required to measure the viscosity over this wide range. One

centipoise, 1 cP (= 10-3 Pa·s or 1 mPa·s), is approximately the viscosity of water at

room temperature. Shear rates corresponding to many industrial processes can also

vary over a wide range, as indicated in Table 2-2.

Page 22: A Study of Non-Newtonian Viscosit y and Yield Stress of

7

(a)

(b)

Fig. 2-1. Flow curves of a Newtonian fluid. (a) Shear stress vs. Shear rate.

(b) Viscosity vs. Shear rate.

0

50

100

0 50 100 150

Shear rate

Shea

r stre

ss

0

10

0 50 100

Shear rate

Visc

osity

Page 23: A Study of Non-Newtonian Viscosit y and Yield Stress of

8

Table 2-1. Viscosity of some familiar materials at room temperature [Barnes et al., 1989].

Liquid Approximate Viscosity (Pa·s)

Glass 1040

Asphalt 108

Molten polymers 103

Heavy syrup 102

Honey 101

Glycerin 100

Olive oil 10-1

Light oil 10-2

Water 10-3

Air 10-5

Page 24: A Study of Non-Newtonian Viscosit y and Yield Stress of

9

Table 2-2. Range of shear rates of some familiar materials and processes [Barnes et al., 1989].

Process Range of

Shear Rates (s-1) Application

Sedimentation of fine powders in a suspending liquid

10-6 – 10-4 Medicines, paints

Leveling due to surface tension 10-2 – 10-1 Paints, printing inks

Draining under gravity 10-1 – 101 Painting, coating

Screw extruders 100 – 102 Polymer melts, dough

Chewing and swallowing 101 – 102 Foods

Dip coating 101 – 102 Paints, confectionery

Mixing and stirring 101 – 103 Manufacturing liquids

Pipe flow 100 – 103 Pumping, blood flow

Spraying and brushing 103 – 104 Fuel atomization, painting

Rubbing 104 – 105 Application of creams and

lotions to the skin

Injection mold gate 104 – 105 Polymer melts

Milling pigments in fluid bases 103 – 105 Paints, printing inks

Blade coating 105 – 106 Paper

Lubrication 103 – 107 Gasoline engines

Page 25: A Study of Non-Newtonian Viscosit y and Yield Stress of

10

2.2. Non-Newtonian Fluid

Any fluids that do not obey the Newtonian relationship between shear stress

and shear rate are non-Newtonian. The subject of rheology is devoted to the study of

the behavior of such fluids. Aqueous solutions of high molecular weight polymers or

polymer melts, and suspensions of fine particles are usually non-Newtonian.

2.2.1. General Non-Newtonian Fluid

In the case of general non-Newtonian fluids, the slope of shear stress versus

shear rate curve is not constant. When the viscosity of a fluid decreases with

increasing shear rate, the fluid is called shear-thinning. In the opposite case where the

viscosity increases as the fluid is subjected to a high shear rate, the fluid is called

shear-thickening. The shear-thinning behavior is more common than the shear-

thickening.

In general, the Newtonian constitutive equation accurately describes the

rheological behavior of low molecular weight polymer solutions and even high

molecular weight polymer solutions at very slow rates of deformation. However,

viscosity can be a strong function of the shear rate for polymeric liquids, emulsions,

and concentrated suspensions.

Page 26: A Study of Non-Newtonian Viscosit y and Yield Stress of

11

2.2.1.1. Power-law Model

One of the most widely used forms of the general non-Newtonian constitutive

relation is a power-law model, which can be described as [Middleman, 1968; Bird et

al., 1987; Munson et al., 1998]:

nmγτ &= (2-2)

where m and n are power-law model constants. The constant, m , is a measure of

the consistency of the fluid: the higher the m is, the more viscous the fluid is. n is a

measure of the degree of non-Newtonian behavior: the greater the departure from the

unity, the more pronounced the non-Newtonian properties of the fluid are.

The viscosity for the power-law fluid can be expressed as [Middleman, 1968;

Bird et al., 1987; Munson et al., 1998]:

1−= nmγη & (2-3)

where η is non-Newtonian apparent viscosity. It is well known that the power-law

model does not have the capability to handle the yield stress. If n < 1, a shear-

thinning fluid is obtained, which is characterized by a progressively decreasing

apparent viscosity with increasing shear rate. If n > 1, we have a shear-thickening

fluid in which the apparent viscosity increases progressively with increasing shear

rate. When n = 1, a Newtonian fluid is obtained. These three types of power-law

models are illustrated in Fig. 2-2.

One of the obvious disadvantages of the power-law model is that it fails to

describe the viscosity of many non-Newtonian fluids in very low and very high shear

rate regions. Since n is usually less than one, η goes to infinity at a very low shear

Page 27: A Study of Non-Newtonian Viscosit y and Yield Stress of

12

rate (see Fig. 2-2) rather than to a constant, 0η , as is often observed experimentally.

Viscosity for many suspensions and dilute polymer solutions becomes constant at a

very high shear rate, a phenomenon that cannot be described by the power-law model.

2.2.1.2. Cross Model

As discussed in the previous section, the power-law model does not have the

capability of handling Newtonian regions of shear-thinning fluids at very low and

high shear rates. In order to overcome this drawback of the power-law model, Cross

(1965) proposed a model that can be described as [Ferguson and Kemblowski, 1991;

Cho and Kensey, 1991; Macosko, 1994]:

+−

+= ∞∞ nmγ

ηηηγτ

&&

10 (2-4)

where

0η and ∞η = viscosities at very low and high shear rates, respectively

m and n = model constants.

At an intermediate shear rate, the Cross model behaves like a power-law model as

shown in Fig. 2-3. However, unlike the power-law model, the Cross model produces

Newtonian viscosities ( 0η and ∞η ) at both very low and high shear rates.

Page 28: A Study of Non-Newtonian Viscosit y and Yield Stress of

13

2.2.2. Viscoplastic Fluid

The other important class of non-Newtonian fluids is a viscoplastic fluid.

This is a fluid which will not flow when a very small shear stress is applied. The

shear stress must exceed a critical value known as the yield stress for the fluid to flow.

For example, when opening a tube of toothpaste, we need to apply an adequate force

in order to make the toothpaste start to flow. Therefore, viscoplastic fluids behave

like solids when the applied shear stress is less than the yield stress. Once the applied

shear stress exceeds the yield stress, the viscoplastic fluid flows just like a normal

fluid. Examples of viscoplastic fluids are blood, drilling mud, mayonnaise,

toothpaste, grease, some lubricants, and nuclear fuel slurries.

2.2.2.1. Bingham Plastic Model

Many types of food stuffs exhibit a yield stress and are said to show a plastic

or viscoplastic behavior. One of the simplest viscoplastic models is the Bingham

plastic model, and it can be expressed as follows [Bird et al., 1987; Ferguson and

Kemblowski, 1991; Macosko, 1994]:

yBm τγτ += & when yττ ≥ , (2-5)

0=γ& when yττ ≤ , (2-6)

where

yτ = a constant that is interpreted as yield stress

Page 29: A Study of Non-Newtonian Viscosit y and Yield Stress of

14

Bm = a model constant that is interpreted as plastic viscosity.

Basically, the Bingham plastic model can describe the viscosity characteristics of a

fluid with yield stress whose viscosity is independent of shear rate as shown in Fig. 2-

4. Therefore, the Bingham plastic model does not have the ability to handle the

shear-thinning characteristics of non-Newtonian fluids.

2.2.2.2. Casson Model

This model was originally introduced by Casson (1959) for the prediction of

the flow behavior of pigment-oil suspensions. The Casson model is based on a

structure model of the interactive behavior of solid and liquid phases of a two-phase

suspension [Casson, 1959]. The model describes the flow of viscoplastic fluids that

can be mathematically described as follows [Bird et al., 1987; Ferguson and

Kemblowski, 1991; Cho and Kensey, 1991; Macosko, 1994]:

γττ &ky += when yττ ≥ , (2-7)

0=γ& when yττ ≤ , (2-8)

where k is a Casson model constant.

The Casson model shows both yield stress and shear-thinning non-Newtonian

viscosity. For materials such as blood and food products, it provides better fit than

the Bingham plastic model [Fung 1990; Cho and Kensey, 1991; Nguyen and Boger,

1992; Fung, 1993].

Page 30: A Study of Non-Newtonian Viscosit y and Yield Stress of

15

2.2.2.3. Herschel-Bulkley Model

The Herschel-Bulkley model extends the simple power-law model to include a

yield stress as follows [Herschel and Bulkley, 1926; Tanner, 1985; Ferguson and

Kemblowski, 1991; Holdsworth, 1993]:

ynm τγτ += & when yττ ≥ , (2-9)

0=γ& when yττ ≤ , (2-10)

where m and n are model constants.

Like the Casson model, it shows both yield stress and shear-thinning non-

Newtonian viscosity, and is used to describe the rheological behavior of food

products and biological liquids [Ferguson and Kemblowski, 1991; Holdsworth, 1993].

In addition, the Herschel-Bulkley model also gives better fit for many biological

fluids and food products than both power-law and Bingham plastic models.

Page 31: A Study of Non-Newtonian Viscosit y and Yield Stress of

16

Fig. 2-2. Flow curves of power-law fluids. (a) shear-thinning fluid (n < 1).

(b) Newtonian fluid (n = 1). (c) shear-thickening fluid (n > 1).

0

5

10

0 50 100 150

Shear rate

(b)

(c)

Vis

cosi

ty

(a)

0

50

100

0 50 100 150

(b)

(c)

Shea

r stre

ss

(a)

Page 32: A Study of Non-Newtonian Viscosit y and Yield Stress of

17

Fig. 2-3. Flow curve of a Cross model.

Shear rate (log)

Viscosity (log)

∞η

Power-law region

Newtonian regions

Page 33: A Study of Non-Newtonian Viscosit y and Yield Stress of

18

Fig. 2-4. Flow curves of viscoplastic fluids. (a) Casson or Herschel-Bulkley fluid.

(b) Bingham plastic fluid.

0

50

100

0 50 100 150

(b)

Shea

r stre

ss

(a)

Bm

Shear rate

1

Page 34: A Study of Non-Newtonian Viscosit y and Yield Stress of

19

2.3. Rheology of Blood

Blood behaves like a non-Newtonian fluid whose viscosity varies with shear

rate. The non-Newtonian characteristics of blood come from the presence of various

cells in the blood (typically making up 45% of the blood’s volume), which make

blood a suspension of particles [Fung, 1993; Guyton and Hall, 1996]. When the

blood begins to move, these particles (or cells) interact with plasma and among

themselves. Hemorheologic parameters of blood include whole blood viscosity,

plasma viscosity, red cell aggregation, and red cell deformability (or rigidity).

2.3.1. Determinants of Blood Viscosity

Much research has been performed to formulate a theory that accounts

completely for the viscous properties of blood, and some of the key determinants

have been identified [Dinnar, 1981; Chien et al., 1987; Guyton and Hall, 1996]. The

four main determinants of whole blood viscosity are (1) plasma viscosity, (2)

hematocrit, (3) RBC deformability and aggregation, and (4) temperature. The first

three factors are parameters of physiologic concern because they pertain to changes in

whole blood viscosity in the body. Especially, the second and third factors,

hematocrit and RBC aggregations, mainly contribute to the non-Newtonian

characteristics of shear-thinning viscosity and yield stress.

Page 35: A Study of Non-Newtonian Viscosit y and Yield Stress of

20

2.3.1.1. Plasma Viscosity

Plasma is blood from which all cellular elements have been removed. It has

been well established that plasma behaves like a Newtonian fluid. Careful tests

conducted using both rotating and capillary tube viscometers over a range of shear

rates (i.e., from 0.1 to 1200 s-1) found no significant departures from linearity.

Therefore, its viscosity is independent of shear rate. Figure 2-5 illustrates this clearly

in the horizontal viscosity line for plasma [Dintenfass, 1971; Dinnar, 1981]. Since

blood is a suspension of cells in plasma, the plasma viscosity affects whole blood

viscosity, particularly at high shear rates.

2.3.1.2. Hematocrit

Hematocrit is the volume percentage of red blood cells in whole blood. Since

studies have shown normal plasma to be a Newtonian fluid [Fung, 1993], the non-

Newtonian features of human blood undoubtedly come from suspended cells in blood.

The rheological properties of suspensions correlate highly with the concentrations of

suspended particles. In blood, the most important suspended particles are the red

blood cells (RBC). Hematocrit is the most important determinant of whole blood

viscosity [Benis et al., 1970; Thurston, 1978; Fung, 1993; Picart et al., 1998; Cinar et

al., 1999]. The effect of hematocrit on blood viscosity has been well documented.

All studies have shown that the viscosity of whole blood varies directly with

Page 36: A Study of Non-Newtonian Viscosit y and Yield Stress of

21

hematocrit at all cell concentrations above 10%. In general, the higher the hematocrit,

the greater the value of whole blood viscosity [Dintenfass, 1971; Dinnar, 1981; Chien

et al., 1987; Guyton and Hall, 1996].

2.3.1.3. RBC Deformability

Deformability is a term used to describe the structural response of a body or

cell to applied forces. The effect of RBC deformability in influencing general fluidity

of whole blood is clearly revealed in Fig. 2-6. This figure shows the relative viscosity

of blood at a shear rate >100 s-1 (at which particle aggregation is negligible, isolating

RBC deformability) compared with that of suspensions with rigid spheres. At 50%

concentration, the viscosity of a suspension of rigid spheres reaches almost infinity so

that the suspension is not able to flow. On the contrary, normal blood remains fluid

even at a hematocrit of 98%, on account of the deformability of its RBCs [Fung,

1993].

2.3.1.4. RBC Aggregation - Major Factor of Shear-Thinning Characteristic

Since red cells do not have a nucleus, they behave like a fluid drop [Dinnar,

1981]. Hence, when a number of red cells cluster together as in the flow of a low

shear rate, they aggregate together. Accordingly, human RBCs have the ability to

Page 37: A Study of Non-Newtonian Viscosit y and Yield Stress of

22

form aggregates known as rouleaux. Rouleaux formation is highly dependent on the

concentration of fibrinogen and globulin in plasma. Note that bovine blood does not

form rouleaux because of absence of fibrinogen and globulin in plasma [Fung, 1993].

Various degrees and numbers of rouleaux in linear array and branched network are

pictured in Fig. 2-7.

Figure 2-8 shows the relationship between blood viscosity and rouleaux

formation. Rouleaux formation of healthy red cells increases at decreasing shear

rates. As red cells form rouleaux, they will tumble while flowing in large vessels.

The tumbling disturbs the flow and requires the consumption of energy, thus

increasing blood viscosity at low shear [Fung, 1993]. As shear rate increases, blood

aggregates tend to be broken up, resulting in drop in blood viscosity (see Fig. 2-8). In

short, rouleaux formation increases blood viscosity, whereas breaking up rouleaux

decreases blood viscosity.

2.3.1.5. Temperature

Temperature has a dramatic effect on the viscosity of any liquid, including

whole blood and plasma. As in most fluids, blood viscosity increases as temperature

decreases [Fung, 1993; Guyton and Hall, 1996]. In blood, reduced RBC

deformability and increased plasma viscosity particularly elevate whole blood

viscosity at low temperatures [Barbee, 1973]. Consequently, precise control of the

sample temperature is necessary to measure viscosity accurately in vitro. It is

Page 38: A Study of Non-Newtonian Viscosit y and Yield Stress of

23

preferable and is a standard in hemorheologic studies to carry out blood viscosity

measurements at body temperature of 37. Typically, blood viscosity increases less

than 2% for each decrease in temperature [Barbee, 1973].

2.3.2. Yield Stress and Thixotropy

2.3.2.1. Yield Stress

In addition to non-Newtonian viscosity, blood also exhibits a yield stress. The

source of the yield stress is the presence of cells in blood, particularly red cells.

When such a huge amount (40-45% by volume) of red cells of 8-10 microns in

diameter is suspended in plasma, cohesive forces among the cells are not negligible.

The forces existing between particles are van der Waals-London forces and

Coulombic forces [Cheng and Evans, 1965; Mewis and Spaull, 1976]. Hence, in

order to initiate a flow from rest, one needs to have a force which is large enough to

break up the particle-particle links among the cells.

However, blood contains 40-45% red cells and still moves relatively easily.

The healthy red cells behave like liquid drops because the membranes of red cells are

so elastic and flexible. Note that in a fluid with no suspended particles, the fluid

starts to move as soon as an infinitesimally small amount of force is applied. Such a

fluid is called a fluid without yield stress. Examples of fluid with no yield stress

include water, air, mineral oils, and vegetable oils. Examples of fluids having the

Page 39: A Study of Non-Newtonian Viscosit y and Yield Stress of

24

yield stress include blood, ketchup, salad dressings, grease, paint, and cosmetic

liquids.

The magnitude of the yield stress of human blood appears to be at the order of

0.05 dyne/cm2 (or 5 mPa) [Schmid- nbeinoSch && and Wells, 1971; Walawender et al.,

1975; Nakamura and Sawada, 1988; Fung, 1993; Stoltz et al., 1999] and is almost

independent of temperature in the range of 10-37 [Barbee, 1973].

2.3.2.2. Thixotropy - Time Dependence

The phenomenon of thixotropy in a liquid results from the microstructure of

the liquid system. Thixotropy may be explained as a consequence of aggregation of

suspended particles. If the suspension is at rest, the particle aggregation can form,

whereas if the suspension is sheared, the weak physical bonds among particles are

ruptured, and the network among them breaks down into separate aggregates, which

can disintegrate further into smaller fragments [Barnes, 1997].

After some time at a given shear rate, a dynamic equilibrium is established

between aggregate destruction and growth, and at higher shear rates, the equilibrium

is shifted in the direction of greater dispersion. The relatively long time required for

the microstructure to stabilize following a rapid change in the rate of flow makes

blood thixotropy readily observable [How, 1996].

This effect on viscosity has been studied using a steady flow [Huang et al.,

1975]. At high shear rates, structural change occurs more rapidly than that at low

Page 40: A Study of Non-Newtonian Viscosit y and Yield Stress of

25

shear rates. In their study, the first step was from the no-flow condition to a shear

rate of 10 s-1. They found that blood viscosity decreased over a period of

approximately 20 seconds at the shear rate of 10 s-1 before the final state was attained.

Next, when the shear rate stepped from 10 to 100 s-1, almost no time was required to

reach the microstructual equilibrium after the change of shear rate.

Gaspar-Rosas and Thurston (1988) also investigated on erythrocyte aggregate

rheology by varying shear rate from 500 s-1 to zero. Based on their results, it can be

concluded that the recovery of quiescent structure requires approximately 50 seconds

while the high shear rate structure is attained in a few seconds. In other words, in

order to minimize the effect of the thixotropic characteristic of blood on the viscosity

measurement between the shear rates of 500 and 1 s-1, at least 50 seconds should be

allowed during the test to have the fully aggregated quiescent state at a shear rate near

1 s-1.

Page 41: A Study of Non-Newtonian Viscosit y and Yield Stress of

26

Fig. 2-5. Comparison of Newtonian plasma viscosity and shear-thinning whole blood viscosity.

Shear rate (s-1)400 10

4

1

Vis

cosi

ty (c

P)

100

Whole blood

Plasma

Page 42: A Study of Non-Newtonian Viscosit y and Yield Stress of

27

Fig. 2-6. Variation of the relative viscosity of blood and suspension with rigid spheres at a shear rate > 100 s-1 [Goldsmith, 1972].

Particle volume fraction

0.8 0.2

10

1

Rel

ativ

e vi

scos

ity

0.4

100

0.6

Normal blood

Suspension with rigid spheres

Page 43: A Study of Non-Newtonian Viscosit y and Yield Stress of

28

Fig. 2-7. Rouleaux formation of human red blood cells photographed on a microscope slide showing single linear and branched aggregates (left part) and a network (right part). The number of cells in linear array are 2, 4, 9, 15 and 36 in a, b, c, d, and f, respectively. [Fung, 1993; Goldsmith, 1972]

Page 44: A Study of Non-Newtonian Viscosit y and Yield Stress of

29

Fig. 2-8. Elevated blood viscosity at low shear rates indicates RBC aggregation (rouleaux formation). Blood viscosity decreases with increasing shear rates as RBC aggregations breaks up to individual red cells.

1

Rel

ativ

e vi

scos

ity

10

Normal blood

Shear rate (s-1)4001 10

Page 45: A Study of Non-Newtonian Viscosit y and Yield Stress of

30

CHAPTER 3. CONVENTIONAL RHEOMETRY: STATE-OF-THE-ART

This chapter reviews literature on conventional rheometries. Section 3.1

briefly introduces conventional rheometers. In sections 3.2 and 3.3, viscometers

commonly used for the viscosity measurements of fluids, which have been used for

hemorheology studies, are demonstrated. Section 3.4 provides conventional methods

of measuring yield stresses of fluids. Section 3.5 presents the drawbacks of

conventional viscometers for clinical applications.

3.1. Introduction

Numerous types of rheometers have been used to measure the viscosity and

yield stress of materials [Tanner, 1985; Ferguson and Kemblowski, 1991; Macosko,

1994]. In the present study, rheometer refers to a device that can measure both

viscosity and yield stress of a material, whereas viscometer can measure only the

viscosity of the material. In addition, only shear viscometers will be discussed in the

study since the other type, extensional viscometers, are not very applicable to

relatively low viscous fluids, such as water and whole blood.

Typically, shear viscometers can be divided into two groups [Macosko, 1994]:

drag flows, in which shear is generated between a moving and a stationary solid

surface, and pressure-driven flows, in which shear is generated by a pressure

difference over a capillary tube. The commonly utilized members of these groups are

Page 46: A Study of Non-Newtonian Viscosit y and Yield Stress of

31

shown in Fig. 3-1. Numerous techniques have been developed for determining the

yield stress of fluids both directly and indirectly.

Most of these viscometers can produce viscosity measurements at a specified,

constant shear rate. Therefore, in order to measure the viscosity over a range of shear

rates, one needs to repeat the measurement by varying either the pressure in the

reservoir tank of capillary tube viscometers, the rotating speed of the cone or cup in

rotating viscometers, or the density of the falling objects. Such operations make

viscosity measurements difficult and labor intensive. In addition, these viscometers

require anticoagulants in blood to prevent blood clotting. Hence, the viscosity results

include the effects of anticoagulants, which may increase or decrease blood viscosity

depending on the type of anticoagulant [Rosenblum, 1968; Crouch et al., 1986;

Reinhart et al., 1990; Kamaneva et al., 1994].

Drag-flow type of viscometers includes a falling object (ball or cylinder)

viscometer and a rotational viscometer. However, the falling object viscometer is not

very convenient to use for clinical applications. In the case of the falling object

viscometer, the relatively large amount of a test fluid is required for the viscosity

measurement. In addition, since the testing fluid is at a stationary state initially, the

type of viscometer is not very applicable to a thixotropic fluid like whole blood. The

principle of the falling object viscometer is provided in Appendix B.

For the yield measurement of blood, most researchers have used indirect

methods rather than direct methods for practical reasons [Nguyen and Boger, 1983;

de Kee et al., 1986; Magnin and Piau, 1990]. Thus, the details of direct methods will

Page 47: A Study of Non-Newtonian Viscosit y and Yield Stress of

32

not be discussed in this chapter. As indirect methods, data extrapolation and

extrapolation using constitutive models are introduced and discussed in this chapter.

Page 48: A Study of Non-Newtonian Viscosit y and Yield Stress of

33

Fig. 3-1. Rheometers.

Rheometers

Viscosity Measurements

Yield Stress Measurements

Drag Flows

Pressure- Driven Flows

Indirect Methods

Direct Methods

Falling/ Rolling Object

Viscometer

Capillary-Tube

Viscometer

Data Extrapolation

Extrapolation using

Constitutive Models

Rotational Viscometer

Page 49: A Study of Non-Newtonian Viscosit y and Yield Stress of

34

3.2. Rotational Viscometer

In a rotational viscometer, the fluid sample is sheared as a result of the

rotation of a cylinder or cone. The shearing occurs in a narrow gap between two

surfaces, usually one rotating and the other stationary. Two frequently used

geometries are Couette (Fig. 3-2) and cone-and-plate (Fig. 3-3).

3.2.1. Rotational Coaxial-Cylinder (Couette Type)

In a coaxial-cylinder system, the inner cylinder is often referred to as bob, and

the external one as cup. The shear rate is determined by geometrical dimensions and

the speed of rotation. The shear stress is calculated from the torque and the

geometrical dimensions. By changing the speed of the rotating element, one is able to

collect different torques, which are used for the determination of the shear stress-

shear rate curve. Figure 3-2 shows a typical coaxial-cylinder system that has a fluid

confined within a narrow gap ( 99.0≥o

i

RR

) between the inner cylinder rotating at Ω

and the stationary outer cylinder.

Once the torque exerting on either inner or outer cylinder is measured, the

shear stress and shear rate can be calculated as follow [Macosko, 1994]:

HRMR

i

ii 22)(

πτ = or

HRMR

o

oo 22)(

πτ = (3-1)

Page 50: A Study of Non-Newtonian Viscosit y and Yield Stress of

35

iooi RR

RRR−Ω

=≅ )()( γγ && when 99.01 ≥>o

i

RR

(3-2)

where

iR and oR = radii of inner and outer cylinders, respectively

2oi RRR +

=

iM and oM = torques exerting on inner and outer cylinders, respectively

H = height of inner cylinder

Ω = angular velocity.

3.2.2. Cone-and-Plate

The common feature of a cone-and-plate viscometer is that the fluid is sheared

between a flat plate and a cone with a low angle; see Fig. 3-3. The cone-and-plate

system produces a flow in which the shear rate is very nearly uniform. Let’s consider

a fluid, which is contained in the gap between a plate and a cone with an angle of β .

Typically, the gap angle, β , is very small ( o4≤ ). The shear rate of the fluid depends

on the gap angle, β , and the linear speed of the plate. Assuming that the cone is

stationary and the plate rotates with a constant angular velocity of Ω , the shear stress

and shear rate can be calculated from experimentally measured torque, M , and given

geometric dimensions (see Fig. 3-3) as follows [Macosko, 1994]:

323

RMπ

τ = and β

γ Ω=& . (3-3)

Page 51: A Study of Non-Newtonian Viscosit y and Yield Stress of

36

Fig. 3-2. Schematic diagram of a concentric cylinder viscometer.

HoR

iR

Ω

Page 52: A Study of Non-Newtonian Viscosit y and Yield Stress of

37

Fig. 3-3. Schematic diagram of a cone-and-plate viscometer.

Ω

R

β

Cone

Plate

Torque measurement device

Fluid

Page 53: A Study of Non-Newtonian Viscosit y and Yield Stress of

38

3.3. Capillary-Tube Viscometer

The principle of a capillary tube viscometer is based on the Hagen-Poiseuille

Equation which is valid for Newtonian fluids. Basically, one needs to measure both

pressure drop and flow rate independently in order to measure the viscosity with the

capillary tube viscometer. Since the viscosity of a Newtonian fluid does not vary

with flow or shear, one needs to have one measurement at any flow velocity.

However, for non-Newtonian fluids, it is more complicated because the viscosity

varies with flow velocity (or shear rate).

In a capillary-tube viscometer, the fluid is forced through a cylindrical

capillary tube with a smooth inner surface. The flow parameters have to be chosen in

such a way that the flow may be regarded as steady-state, isothermal, and laminar.

Knowing the dimensions of the capillary tube (i.e., its inner diameter and length), one

can determine the functional dependence between the volumetric flow rate and the

pressure drop due to friction. If the measurements are carried out so that it is possible

to establish this dependence for various values of pressure drop or flow rate, then one

is able to determine the flow curve of the fluid.

For non-Newtonian fluids, since the viscosity varies with shear rate, one needs

to vary the pressure in the reservoir in order to change the shear rate, a procedure that

is highly time-consuming. After each run, the reservoir pressure should be reset to a

new value to obtain the relation between flow rate and pressure drop. In order to

determine the flow curve of a non-Newtonian fluid, one needs to establish the

functional dependence of shear stress on shear rate in a wide range of these variables.

Page 54: A Study of Non-Newtonian Viscosit y and Yield Stress of

39

Figure 3-4 shows the schematic diagram of a typical capillary-tube viscometer,

which has the capillary tube with an inner radius of cR and a length of cL . It is

assumed that the ratio of the capillary length to its inner radius is so large that one

may neglect the so-called end effects occurring in the entrance and exit regions of the

capillary tube. Then, the shear stress at the tube wall can be obtained as follows:

c

c

LPr

2∆

=τ (3-4)

c

ccw L

PR2∆

=τ (3-5)

where

τ and wτ = shear stresses at distance r and at tube wall, respectively

r = distance from the capillary axis

cP∆ = pressure drop across a capillary tube.

It is of note that the shear stress distribution is valid for fluids of any rheological

properties.

In the case of a Newtonian fluid, the shear rate at tube wall can be expressed

by taking advantage of the well-known Hagen-Poiseuille Equation as:

ccw R

VRQ 44

3 ==π

γ& (3-6)

where

wγ& = wall shear rate

VRL

PRQ cc

cc ⋅=∆

= 24

µπ

= volumetric flow rate (Hagen-Poiseuille Equation)

V = mean velocity.

Page 55: A Study of Non-Newtonian Viscosit y and Yield Stress of

40

Fig. 3-4. Schematic diagram of a capillary-tube viscometer.

Test fluid

Capillary tube

Air

Balance

Collected test fluid

cR2cL

Reservoir tank

Compressedair

Page 56: A Study of Non-Newtonian Viscosit y and Yield Stress of

41

3.4. Yield Stress Measurement

Whether yield stress is a true material property or not is still a controversial

issue [Barnes and Walters, 1985]. However, there is generally an acceptance of its

practical usefulness in engineering design and operation of processes where handling

and transport of industrial suspensions are involved. The minimum pump pressure

required to start a slurry pipeline, the leveling and holding ability of paint, and the

entrapment of air in thick pastes are typical problems where the knowledge of the

yield stress is essential.

Numerous techniques have been developed for determining the yield stress

both directly and indirectly based on the general definition of the yield stress as the

stress limit between flow and non-flow conditions. Indirect methods simply involve

the extrapolation of shear stress-shear rate data to zero shear rate with or without the

help of a rheological model. Direct measurements generally rely on some

independent assessment of yield stress as the critical shear stress at which the fluid

yields or starts to flow.

The value obtained by the extrapolation of a flow curve is known as

“extrapolated” or “apparent” yield stress, whereas yield stress measured directly,

usually under a near static condition, is termed “static” or “true” yield value.

Page 57: A Study of Non-Newtonian Viscosit y and Yield Stress of

42

3.4.1. Indirect Method

Indirect determination of the yield stress simply involves the extrapolation of

experimental shear stress-shear rate data at zero shear rate (see Fig. 3-5). The

extrapolation may be performed graphically or numerically, or can be fitted to a

suitable rheological model representing the fluid and the yield stress parameter in the

model is determined.

3.4.1.1. Direct Data Extrapolation

One of most common procedures is to extend the flow curve at low shear rates

to zero shear rate, and take the shear stress intercept as the yield stress value. The

technique is relatively straightforward only if the shear stress-shear rate data are

linear. With nonlinear flow curves, as shown in Fig. 3-5, the data may have to be

fitted to a polynomial equation followed by the extrapolation of the resulting curve fit

to zero shear rate. The yield stress value obtained obviously depends on the lowest

shear rate data available and used in the extrapolation. This shear rate dependence of

the extrapolated yield stress has been demonstrated by Barnes and Walters (1985)

with a well-known yield stress fluid, Carbopol (carboxylpolymethylene). They

concluded that this fluid would have no detectable yield stress even if measurement

was made at very low shear rates of 10-5 s-1 or less. This finding should be viewed

with caution, however, since virtually all viscometric instruments suffer wall slip and

Page 58: A Study of Non-Newtonian Viscosit y and Yield Stress of

43

other defects which tend to be more pronounced at low shear rates especially with

yield stress fluids and particulate systems [Wildermuth and Williams, 1985; Magnin

and Piau, 1990]. Thus, it is imperative that some checking procedure should be

carried out to ascertain the reliability of the low shear rate data before extrapolation is

made.

3.4.1.2. Extrapolation Using Constitutive Models

A more convenient extrapolation technique is to approximate the experimental

data with one of the viscoplastic flow models. Many workers appear to prefer the

Bingham model which postulates a linear relationship between shear stress and shear

rate. However, since a large number of yield stress fluids including suspensions are

not Bingham plastic except at very high shear rates, the use of the Bingham plastic

model can lead to unnecessary overprediction of the yield stress as shown in Fig. 3-5

[Nguyen and Boger, 1983; de Kee et al., 1986]. Extrapolation by means of nonlinear

Casson model can be used from a linear plot of 21

τ versus 21

γ& . The application of

Herschel-Bulkley model is less certain although systematic procedures for

determining the yield stress value and the other model parameters are available

[Heywood and Cheng, 1984].

Even with the most suitable model and appropriate technique, the yield stress

value obtained cannot be regarded as an absolute material property because its

accuracy depends on the model used and the range and reliability of the experimental

Page 59: A Study of Non-Newtonian Viscosit y and Yield Stress of

44

data available. Several studies have shown that a given fluid can be described equally

well by more than one model and hence can have different yield stress values

[Keentok, 1982; Nguyen and Boger, 1983; Uhlherr, 1986].

3.4.2. Direct Method

Various techniques have been introduced for measuring the yield stress

directly and independently of shear stress-shear rate data. Although the general

principle of the yield stress as the stress limit between flow and non-flow conditions

is often used, the specific criterion employed for defining the yield stress seems to

vary among these techniques. Furthermore, each technique appears to have its own

limitations and sensitivity so that no single technique can be considered versatile or

accurate enough to cover the whole range of yield stress and fluid characteristics.

Usually, the direct methods are used for fluids having yield stresses of greater than

approximately 10 Pa [Nguyen and Boger, 1983]. Therefore, as mentioned earlier, the

direct method is not very convenient to use for the yield stress measurement of blood

since the yield stress of human blood is approximately 1 to 30 mPa [Picart et al.,

1998].

Page 60: A Study of Non-Newtonian Viscosit y and Yield Stress of

45

Fig. 3-5. Determination of yield stress by extrapolation [Nguyen and Boger, 1983].

Page 61: A Study of Non-Newtonian Viscosit y and Yield Stress of

46

3.5. Problems with Conventional Viscometers for Clinical Applications

3.5.1. Problems with Rotational Viscometers

Over the years, rotational viscometers have been the standard in clinical

studies investigating rheological properties of blood and other body fluids. Despite

their popularity, rotational viscometers have some drawbacks that limit their clinical

applicability in measuring whole blood viscosity. They include the need to calibrate a

torque-measuring sensor, handling of blood, surface tensions effects, and the range of

reliability.

The torque-measuring sensor can be a conventional spring or a more

sophisticated electronic transducer. In either case, the sensor requires a periodic

calibration because repeated use of the sensor can alter its spring constant. The

calibration procedure is often carried out at manufacturer’s laboratory because it

requires an extremely careful and elaborate protocol, requiring the viscometer unit to

be returned for service.

Another concern is the need to work with contaminated blood specimens.

After each measurement, the blood sample must be removed from the test section,

and the test section must be cleaned manually. Not only is this procedure time-

consuming, but also it poses a potential risk for contact with contaminated blood.

Surface tension effects arise in the use of the coaxial-cylinder viscometer

because surface tension is relatively high for blood and macromolecular solutions.

The contact area between the blood and an inner cylinder is not uniform along the

Page 62: A Study of Non-Newtonian Viscosit y and Yield Stress of

47

periphery. The bob (inner cylinder) is pulled in different directions and revealed in

fluctuating torque readings, introducing serious errors in viscosity measurement.

Another inherent difficulty in measuring whole blood viscosity using

rotational viscometers is the limited shear rate range. In the extremes of the reputed

range (whether high shear or low shear, depending on the instrument), the detected

torque values do not have sufficient accuracy. Usually, manufacturers recommend

discarding viscosity data if the torque is less than 10% of the maximum value of the

sensor. This restriction is a major concern. For example, in the case of Brookfield

rotational viscometer, the minimum shear rate is often limited at approximately 30-50

s-1 due to the 10% restriction.

There are other clinical, practical considerations in using the rotational

viscometer. For example, it is usually necessary to treat the blood sample with a

measurable amount of anticoagulant, such as ethylenediaminetetraacetic acid (EDTA)

or heparin, to prevent coagulation during viscosity measurements. The reason for this

is that the contact area among blood, rotational viscometer component, and air is

relatively large for the size of the blood sample, and it usually takes a relatively long

time to complete viscosity measurements over a range of shear rates. Treating blood

with such anticoagulants results in an altered sample, and subsequent viscosity

measurements do not reflect the intrinsic values of unadulterated blood.

Page 63: A Study of Non-Newtonian Viscosit y and Yield Stress of

48

3.5.2. Problems with Capillary-Tube Viscometers

There are some drawbacks in the use of conventional capillary-tube

viscometers for clinical applications. The range of shear rate is limited to high shears

over 100 s-1. Although one can produce viscosity data at lower shear rates below 100

s-1 with a sophisticated vacuum system, the capillary tube system is basically designed

and operated to obtain viscosity at the high shear range. Since it is essential to obtain

blood viscosity at low shear rates below 10 s-1, the traditional capillary tube

viscometer is not suitable for measuring the viscosity at low shear rates. However,

capillary-tube viscometer is simple in its design and uses gravity field to drive test

fluid such that there is no need for calibration.

It takes a relatively long time to complete viscosity measurements over a

range of shear rates because at each shear rate, a sufficient quantity of a fluid sample

must be collected for an accurate measurement of flow velocity. After the

measurement at one shear rate, the pressure at the reservoir tank must be readjusted to

either increase or decrease shear rate. Then, the next shear rate case resumes. Thus,

anticoagulants must be added to whole blood for the viscosity measurement over a

range of shear rates.

Page 64: A Study of Non-Newtonian Viscosit y and Yield Stress of

49

CHAPTER 4. THEORY OF SCANNING CAPILLARY-TUBE RHEOMETER

Chapter 4 presents the theory of scanning capillary-tube rheometer (SCTR).

Mathematical procedures for both viscosity and yield-stress measurements were

demonstrated in detail using power-law, Casson, and Herschel-Bulkley (H-B) models.

Section 4.1 provides a brief introduction to the SCTR. In section 4.1.1, the

description of a U-shaped tube set is reported. In addition, this section shows how the

dimensions of the disposable tube set were determined. Section 4.1.2 demonstrates

the equations for the energy balance in the disposable tube set.

Section 4.2 provides the mathematical details of data reduction for both

viscosity and yield-stress measurements. Sections 4.2.1, 4.2.2, and 4.2.3 deal with

the mathematical modeling in the data reduction by using the power-law, Casson, and

H-B models, respectively. Especially, in sections 4.2.2 and 4.2.3, the yield stress as

well as the viscosity of blood was considered in the data reduction.

4.1 Scanning Capillary-Tube Rheometer (SCTR)

One of the drawbacks of using conventional capillary viscometers is that one

needs to change the pressure in the reservoir tank in order to measure the viscosity at

a different shear rate. Viscosity can only be measured at one shear rate at a time in

the conventional system. Similarly, in other types of viscometers such as rotating

viscometers and falling object viscometers, the rotating speed has to be changed or

Page 65: A Study of Non-Newtonian Viscosit y and Yield Stress of

50

the density of the falling object has to be changed in order to vary shear rate as

mentioned in Chapter 3. Such operations can make viscosity measurements time

consuming and labor intensive. Because of the time required to measure viscosity

over a range of shear rates, it is necessary to add anticoagulants to blood to prevent

clotting during viscosity measurements with these conventional viscometers. The

present study introduces an innovative concept of a new capillary tube rheometer that

is capable of measuring yield stress and viscosity of whole blood continuously over a

wide range of shear rates without adding any anticoagulants.

4.1.1 U-Shaped Tube Set

Figure 4-1 shows a schematic diagram of a U-shaped tube set, which consists

of two riser tubes, a capillary tube, and a stopcock. The inside diameter of the riser

tubes in the present study is 3.2 mm. The inside diameter and length of the capillary-

tube are 0.797 and 100 mm, respectively. The small diameter of the capillary tube,

compared with that of the riser tubes, was chosen to ensure that the pressure drops at

the riser tubes and connecting fittings were negligibly small compared to the pressure

drop at the capillary tube [Kim et al., 2000a, 2000b, and 2002].

Furthermore, the inside diameter of the capillary tube was chosen to minimize

the wall effect which is often known as Fahraeus-Lindqvist effect [Fahraeus and

Lingqvist, 1931]. The details of the wall effect will be discussed in Chapter 5. In the

present study, the wall effect was found to be negligibly small.

Page 66: A Study of Non-Newtonian Viscosit y and Yield Stress of

51

The length of the capillary tube (i.e., cL = 100 mm) in the U-shaped tube set

was selected to ensure that the end effects would be negligible [Kim et al., 2000a,

2000b, and 2002]. The end effects at the capillary tube will be also reported in

Chapter 5. In addition, the capillary-tube dimensions in the SCTR were selected to

complete one measurement within 2-3 min, a condition that is desirable when

measuring the viscosity of unadulterated whole blood in a clinical environment.

Figure 4-2 shows sketches of the fluid levels in the U-shaped tube set as time

goes on. The fluid level in the right-side riser tube decreases whereas that in the left-

side riser tube increases. As time goes to infinity, the two fluid levels never become

equal due to the surface tension and yield stress effects as shown in Fig. 4-2(c) (i.e.,

∞=∆ th > 0). While a test fluid travels through the capillary tube between riser tubes 1

and 2, the pressure drop caused by the friction at the capillary tube can be obtained by

measuring the fluid levels at riser tubes 1 and 2. In Fig. 4-3, a typical fluid-level

variation measured by the SCTR is shown. Points (a), (b), and (c) represent the three

moments indicated in Fig. 4-2 (i.e., at 0=t , t > 0, and ∞=t , respectively).

4.1.2 Energy Balance

Figure 4-4 shows the liquid-solid interface condition for each fluid column of

a U-shaped tube. A falling column (right side) always has a fully wet surface

condition, while a rising column (left side) has an almost perfectly dry surface

condition at the liquid-solid interface during the entire test. Therefore, the surface

Page 67: A Study of Non-Newtonian Viscosit y and Yield Stress of

52

tension at the right side was consistently greater than that at the left side since the

surface tension of a liquid is strongly dependent on the wetting condition of the tube

at the liquid-solid interface [Jacobs, 1966; Mardles, 1969; Kim et al., 2002]. The

height difference caused by the surface tension at the two riser tubes was one order of

magnitude greater than the experimental resolution desired for accurate viscosity

measurements. Thus, it is extremely important to take into account the effect of the

surface tension on the viscosity measurement using the disposable tube set.

The mathematical model of the flow analysis began with the equation of the

conservation of energy in the form of pressure unit, where the surface-tension effect

was considered between the two top points of the fluid columns at the riser tubes (see

Fig. 4-4). Assuming that the surface tension for the liquid-solid interface at each riser

tube remains constant during the test, one may write the governing equations as [Bird

et al., 1987; Munson et al., 1998]:

dstVhgPghVPghVP

s

stc ∫ ∂∂

+∆+∆+++=++ ∞=2

1

22

2212

11 21

21 ρρρρρρ , (4-1)

where

1P and 2P = static pressures at two top points

ρ = density of fluid

g = gravitational acceleration

1V and 2V = flow velocities at two riser tubes

1h and 2h = fluid levels at two riser tubes

)(tPc∆ = pressure drop across capillary tube

∞=∆ th = additional height difference

Page 68: A Study of Non-Newtonian Viscosit y and Yield Stress of

53

V = flow velocity

t = time

s = distance measured along streamline from some arbitrary initial point.

In Eq. (4-1), the energy emitted from LEDs was ignored since the energy transferred

from the LEDs, which can affect the temperature of a test fluid, was negligible small.

In order to ensure that the amount of the heat emitted from the LEDs is very small,

the temperature of bovine blood was measured during a room-temperature test. The

results showed no changes in temperature during the test, indicating that the energy

emitted from LEDs might be negligibly small.

For the convenience of data-reduction procedure, the unsteady term in Eq. (4-

1), dstVs

s∫ ∂∂2

1

ρ , may be ignored under the assumption of a quasi-steady state. In order

to make the assumption, one should make sure that the pressure drop due to the

unsteady effect is very small compared with that due to the friction estimated from

the steady Poiseuille flow in a capillary tube.

The unsteady term can be broken into three integrations that represent the

pressure drops due to the unsteady flow along the streamlines at riser tube 1, capillary

tube, and riser tube 2 as [Munson et al., 1998]:

++=

∂∂

∫ ∫∫∫′

1

1

2

2

2

1

2

1

s

s

s

srs

s

crs

sds

dtVdds

dtVd

dsdtVdds

tV ρρ , (4-2)

where rV and cV are mean flow velocities at riser and capillary tubes, respectively.

Since the term of tV∂∂ is independent of streamlines, one can simplify the equation as:

Page 69: A Study of Non-Newtonian Viscosit y and Yield Stress of

54

( )dtVd

lldtVd

LldtVd

LdtVd

ldtVd

dstV rc

cr

ccrs

s 2121

2

1

++=

++=

∂∂

∫ ρρρρ , (4-3)

where 1l and 2l are lengths of the liquid columns whereas cL is the length of the

capillary tube as shown in Fig. 4-5. Using the mass conservation, rrcc VRVR ⋅=⋅ 22 ππ ,

the pressure drop due to the unsteady effect can be reduced as:

dtVdll

RRLds

tVP r

c

rc

s

sunsteady

++

=

∂∂

=∆ ∫ 21

2

2

1

ρρ , (4-4)

where

unsteadyP∆ = pressure drop due to the unsteady flow

rR and cR = radii of riser and capillary tubes, respectively.

In the present experimental set up, 1l , 2l , and cL are measured to be

approximately 12, 4, and 10 cm, respectively. Since )(1 th and )(2 th are strongly

dependent on each other by the conservation of mass for incompressible fluids, rV

must be equal to dt

tdh )(1 and dt

tdh )(2 . In order to calculate the term of dtVd r from the

experimental values, one could use the following central differential method:

[ ] [ ]2

2222

111 )()(2)()()(2)(t

tththttht

tththtthdtVd r

∆∆−+−∆+

=∆

∆−+−∆+= . (4-5)

For the comparison of unsteadyP∆ with cP∆ , unsteadyP∆ was estimated through a curve-

fitting process. In order to obtain a smooth curve from raw data, the following

exponential equation was used.

2

⋅−= −btr ea

dtVd

Error . (4-6)

Page 70: A Study of Non-Newtonian Viscosit y and Yield Stress of

55

Two constants, a and b , were obtained through a curve-fitting process, a least-

square method, which minimized the sum of error for all experimental data points

obtained in each test.

Typical results showed that the magnitude of the pressure drop due to the

unsteady flow, unsteadyP∆ , was always less than 1% of that of pressure drop at capillary

tube, cP∆ , over the entire shear-rate range. This confirms that the assumption of a

quasi-steady state could be used for the present data procedure. The details of

experimental results will be discussed in Chapter 5.

Assuming a quasi-steady flow behavior, one may rewrite Eq. (4-1) as follows

[Bird et al., 1987; Munson et al., 1998]:

∞=∆+∆+++=++ tc hgtPtghVPtghVP ρρρρρ )()(21)(

21

22

2212

11 . (4-7)

Since atmPPP == 21 and 21 VV = , Eq. (3-7) can be reduced as:

[ ]∞=∆−−=∆ tc hththgtP )()()( 21ρ . (4-8)

Note that h∆ at ∞=t contains a height difference due to the surface tension, sth∆ ,

and an additional height difference due to the yield stress, yh∆ , for the case of blood

(i.e., see Fig. 4-3). The next section addresses the mathematical procedure of

handling the yield stress.

Page 71: A Study of Non-Newtonian Viscosit y and Yield Stress of

56

Fig. 4-1. Schematic diagram of a U-shaped tube set.

3.2 mm

0.797 mm

100 mm

Riser tubes

Capillary tubeStopcock

Open to air

Page 72: A Study of Non-Newtonian Viscosit y and Yield Stress of

57

(a) at 0=t (b) at 0>t (c) at ∞=t

Fig. 4-2. Fluid-level variation in a U-shaped tube set during a test.

Riser tube 2 Riser tube 1

∞=∆ th

Page 73: A Study of Non-Newtonian Viscosit y and Yield Stress of

58

Fig. 4-3. Typical fluid-level variation measured by a SCTR. (a) at 0=t , (b) at 0>t , and (c) at ∞=t .

Hei

ght

Time(a) (b) (c)

)(1 th

)(2 th

∞=∆ th

Page 74: A Study of Non-Newtonian Viscosit y and Yield Stress of

59

Fig. 4-4. Liquid-solid interface conditions for fluid columns of a U-shaped tube set.

2

2' 1'

1

Dry surface condition

Wet surface condition

1l

cL2l

Page 75: A Study of Non-Newtonian Viscosit y and Yield Stress of

60

4.2 Mathematical Procedure for Data Reduction

In Chapter 2, we discussed the non-Newtonian characteristics of whole blood.

This section deals with non-Newtonian constitutive models for blood and their

applications to the SCTR. Since blood has both shear-thinning (pseudo-plastic) and

yield stress characteristics, three different constitutive models were used for the

viscosity and/or yield-stress measurements of blood in this study. Power-law model

was chosen to demonstrate the shear-thinning behavior of blood. Casson and

Herschel-Bulkley (H-B) models were selected to measure both shear thinning

viscosity and yield stress of blood.

For the purpose of clinical applications, disposable tube sets can be used for

the viscosity and yield-stress measurements of blood. Since the disposable tube sets

have different surface conditions at riser tube 1 and 2 during the test, one needs to

mathematically handle surface tension and yield stress effects in order to measure the

viscosity and yield stress of blood using Casson or H-B model. The details of

mathematical method of isolating those two effects are shown in this section.

4.2.1 Power-law Model

It is well known that power-law model does not have the capability to handle

yield stress. As provided in Chapter 2, the relation among shear stress, shear rate, and

viscosity in power-law fluids may be written as follows:

Page 76: A Study of Non-Newtonian Viscosit y and Yield Stress of

61

nmγτ &= , (4-9)

1−= nmγη & . (4-10)

Since n < 1 for pseudo-plastics, the viscosity function decreases as the shear rate

increases. This type of behavior is characteristic of high polymers, polymer solutions,

and many suspensions including whole blood.

We consider the fluid element in the capillary tube at time t as is shown in

Fig. 4-5. The Hagen-Poiseuille flow may be used to derive the following relationship

for the pressure drop at the capillary tube as a function of capillary tube geometry,

fluid viscosity, and flow rate [Fung, 1990; Munson et al., 1998]:

dtdh

RRL

RQL

RL

RL

rlP

c

rc

c

c

c

wcw

c

cc 4

2

4

88222 µπµγµ

ττ =====∆&

, (4-11)

where

r = radial distance

l = length of fluid element

τ and wτ = shear stress and wall shear stress, respectively

3

4

cw R

γ =& = wall shear rate

µ = Newtonian apparent viscosity

tubecapillary oflength =cL

dtdhR

dtdhR

dtdhRQ rr ⋅=⋅=⋅= 22212

r πππ = volumetric flow rate.

The above relationship is valid for Newtonian fluids whose viscosities are

independent of shear rate. For non-Newtonian fluids, the viscosities vary with shear

Page 77: A Study of Non-Newtonian Viscosit y and Yield Stress of

62

rate. However, the Hagen-Poiseuille flow within the capillary tube still holds for a

quasi-steady laminar flow. When applying a non-Newtonian power-law model to

whole blood, the pressure drop at the capillary tube can be described as follows

[Middleman, 1968; Bird et al., 1987; Fung, 1990]:

n

c

r

c

c

n

cc

c

c

nwc

c

wcwc

dtdh

RR

nn

RmL

RQ

nn

RmL

RmL

RL

P

+

=

+

===∆

3

2

3

132

13222π

γγη &&

, (4-12)

where

wη = power-law apparent viscosity

3

13 c

w RQ

nn

πγ

+

=& .

It is of note that if 1=n , Eq. (4-12) yields to Eq.(4-11). Applying Eqs. (4-8), (4-11),

and (4-12), one can rewrite the energy conservation equation as follows:

dtdh

RRL

hththgc

rct 4

2

218

)()(µ

ρ =∆−− ∞= for Newtonian fluids, (4-13)

132)()( 3

2

21

n

c

r

c

ct dt

dhRR

nn

RmL

hththg

+

=∆−− ∞=ρ

for power-law fluids. (4-14)

For convenience, one may define a new function, ∞=∆−−= thththt )()()( 21θ so that

Eqs. (4-13) and (4-14) become as follows:

dtd αθθ

−= for Newtonian fluids, (4-15)

Page 78: A Study of Non-Newtonian Viscosit y and Yield Stress of

63

dtd

n

1 βθ

θ−= for power-law fluids, (4-16)

where

dtdh

dtdh

dtdh

dtd 221 2−=−=θ

2

4

4 rc

c

RLgRµρ

α =

+

=

3

2

1

213

2

c

r

n

c

c

RR

nn

mLgRρ

β .

The above equations are the first-order linear differential equations. Since α and β

are constants, these equations can be integrated as follows:

)0()( tet αθθ −= for Newtonian fluids, (4-17)

1)0()(11 −−

−=nn

nn

tn

nt βθθ for power-law fluids, (4-18)

where ∞=∆−−= thhh )0()0()0( 21θ : initial condition.

Equation (4-18) can be used for curve fitting of the experimental data (i.e.,

)(1 th and )(2 th ) to determine ∞=∆ th , the power-law index, n , and the consistency

index, m . A least-square method was used for the curve fitting. The data reduction

procedure adopted is as follows:

1. Conduct a test and acquire all data, )(1 th and )(2 th .

2. Guess values for m , n , and ∞=∆ th .

3. Calculate the following error values for all data points:

Page 79: A Study of Non-Newtonian Viscosit y and Yield Stress of

64

[ ]2 )()( valuelTheoreticavaluealExperiment ttError θθ −= . (4-19)

4. Sum the error values for all data points.

5. Iterate to determine the values of m , n , and ∞=∆ th that minimize the sum of

error.

6. Let the computer determine whether a test fluid is Newtonian or not.

7. Calculate shear rate and viscosity for all data points as follows:

)(22

tL

gRP

LR

c

cc

c

cw θ

µρ

µγ =∆=& for Newtonian fluids, (4-20)

n

c

cn

cc

cw t

mLgR

PmLR

11

)(22

=

∆= θ

ργ& for power-law fluids. (4-21)

When n becomes 1 (± 0.001), µ is equal to m , whereas when 0< n <1, the viscosity

is calculated from Eq. (4-10).

In order to obtain the velocity profile at the capillary tube, which changes with

time, using a power-law model, Eq. (4-21) can be used to derive it. Since drdV

−=γ& ,

the velocity profile can be expressed as follows:

n

cc

tPmLr

drrtdV

1

)(2

),(

∆−= ,

Crn

nmL

tPdrr

mLtP

rtV nnn

c

cnn

c

c +⋅

+⋅

∆−=⋅

∆−=

+

∫1

11

1

12)(

2)(

),( , (4-22)

where C is a constant. Using no-slip condition on the capillary wall, 0),( =cRtV ,

the constant can be obtained as:

Page 80: A Study of Non-Newtonian Viscosit y and Yield Stress of

65

nn

c

n

c

c Rn

nmL

tPC

11

12)( +

+⋅

∆= . (4-23)

Finally, the velocity profile within the capillary tube can be expressed as follows:

−⋅

∆−−⋅

+=

−⋅

∆⋅

+=

++∞=

++

nn

nn

c

n

c

t

nn

nn

c

n

c

cc

rRmL

hththn

n

rRmL

tPn

nrtV

111

21

111

2)()(

1

2)(

1),(

(4-24)

where [ ]∞=∆−−=∆ tc hththgtP )()()( 21ρ . Note that if power-law index becomes zero,

1=n , then the above equation yields to the equation for the Newtonian velocity

profile as:

( )22

4)(

),( rRL

tPrtV c

c

cc −⋅

∆=

µ. (4-25)

In order to determine the mean flow velocity at the riser tube, one has to find

the flow rate at the capillary tube first. The flow rate can be obtained by integrating

the velocity profile over the cross-sectional area of the capillary tube as follows:

[ ] nn

c

n

c

t

nn

c

n

c

c

R

c

RmL

hththgnn

RmL

tPnn

rdrrtVtQ c

131

21

131

0

2)()(

13

2)(

13

),(2)(

+∞=

+

∆−−⋅

+=

∆⋅

+=

= ∫

ρπ

π

π

(4-26)

Since )()( 2 tVRtQ rrπ= , the mean flow velocity at the riser tube can be determined by

the following equation:

Page 81: A Study of Non-Newtonian Viscosit y and Yield Stress of

66

[ ]2

131

21

2

131

2)()(

13

2)(

13)(

r

nn

cn

c

t

r

nn

cn

c

cr

RR

mLhththg

nn

RR

mLtP

nntV

+

∞=

+

∆−−⋅

+=

∆⋅

+=

ρ

(4-27)

where rR is the radius of the riser tube.

4.2.2 Casson Model

The Casson model can handle both yield stress and shear-thinning

characteristics of blood, and can be described as follows [Barbee and Cokelet, 1971;

Benis et al., 1971; Reinhart et al., 1990]:

γττ &ky += when yττ ≥ , (4-28)

0=γ& when yττ ≤ , (4-29)

where

τ and γ& = shear stress and shear rate, respectively

yτ = a constant that is interpreted as the yield stress

k = a Casson model constant.

Wall shear stress and yield stress can be defined as follows:

c

ccw L

RtP2

)( ⋅∆=τ , (4-30)

c

ycy L

trtP2

)()( ⋅∆=τ , (4-31)

Page 82: A Study of Non-Newtonian Viscosit y and Yield Stress of

67

where yr is a radial location below which the velocity profile is uniform as shown in

Fig. 4-6, i.e., plug flow, due to the yield stress. Now, for the Casson model, Eq. (4-8)

becomes [ ]stc hththgtP ∆−−=∆ )()()( 21ρ , indicating that the effect of the surface

tension is isolated from the pressure drop across the capillary tube. Using Eqs. (4-28)

and (4-29), one can obtain the expressions of shear rate and velocity profile at the

capillary tube as follows:

2

2)()(

2)(1

⋅∆−

⋅∆=−=

c

yc

c

cc

LtrtP

LrtP

kdrdV

γ& , (4-32)

−+−−−

∆⋅= ))((2))((

38)(

41),( 2

32

32

122 rRtrrRtrrRL

tPk

rtV cycycc

cc

for cy Rrtr ≤≤)( , (4-33)

))(31())((

)(41)( 3 trRtrR

LtP

ktV ycyc

c

cc +−

∆⋅= for rtry ≥)( . (4-34)

For the purpose of simplicity, one may define two new parameters,

cRrrC =)( and

c

yy R

trtC

)()( = , so that Eqs. (4-33) and (4-34) become as follows:

−+

−⋅

∆⋅=

cc

y

cc

y

cc

c

cc R

rRr

Rr

Rr

RrR

LtP

krtV 121

381

)(41),(

23

21

22

( )

−+

−−−

∆= )(1)(2)(1)(

38)(1

4)( 2

321

22

rCtCrCtCrCkL

tPRyy

c

cc

for cy Rrtr ≤≤)( , (4-35)

Page 83: A Study of Non-Newtonian Viscosit y and Yield Stress of

68

( )

( )

+⋅−

∆=

+⋅

∆=

)(311)(1

4)(

3111

4)(

)(

32

33

tCtCkL

tPR

Rr

RRr

RkL

tPtV

yyc

cc

c

yc

c

yc

c

cc

for rtry ≥)( , (4-36)

where )()(

)()(

)(tP

PtR

trtC

w

y

c

yy ∆

∞∆===

ττ

.

In order to determine the mean flow velocity at the riser tube, one has to find

the flow rate at the capillary tube first. The flow rate can be obtained by integrating

the velocity profile over the cross-sectional area of the capillary tube as follows:

−+−

∆=

⋅−⋅+⋅−

∆=

∆⋅−⋅+

∆⋅−

∆=

⋅=

421

4

4214

34

21

214

0

211

34

7161

8

)(211)(

34)(

7161

8

]))(

()2

(211)

2(

34

))(

()2

(7

16))(

[(8

2)(

yyyc

cc

w

y

w

y

w

y

c

cc

c

c

c

y

c

y

c

c

c

y

c

cc

R

c

CCCkL

PR

kLPR

LtP

RR

LtP

RLtP

kR

drrVtQ c

π

ττ

ττ

ττπ

ττ

τπ

π

(4-37)

Since )()( 2 tVRtQ rrπ= , the mean flow velocity at the riser tube can be determined by

the following equation:

Page 84: A Study of Non-Newtonian Viscosit y and Yield Stress of

69

]211

34

7161[

8

]))(

()2

(211)

2(

34

))(

()2

(7

16))(

[(8

)(

421

2

4

34

21

21

2

4

yyycr

cc

c

c

c

y

c

y

c

c

c

y

c

c

r

cr

CCCLkRPR

LtP

RR

LtP

RLtP

kRR

tV

−+−∆

=

∆⋅−⋅+

∆⋅−

∆=

−ττ

τ

(4-38)

where rR is the radius of the riser tube.

For the purpose of simplicity, Eq. (4-38) can be rewritten to clearly display

the unknowns and the observed variables as:

]))()((211

34

)))()(((7

16))()([(8

)(

321

4

21

21212

4

−∆−−∆−∆+

∆−−∆−∆−−=

styy

stystcr

cr

hththhh

hththhhththLkRgR

tVρ

(4-39)

where st

yy hthth

htP

PtC∆−−

∆=

∆∞∆

=)()()(

)()(21

. Note that Eq. (4-39) contains two

independent variables, i.e., )(1 th and )(2 th , and one dependent variable, i.e., )(tVr .

There are three unknown parameters to be determined through the curve fitting in Eq.

(4-39), namely sth∆ , k , and yh∆ . sth∆ is h∆ due to the surface tension, k is the

Casson constant, and yh∆ is h∆ due to the yield stress.

Once the equation for the mean flow velocity, )(tVr , was derived, one could

determined the unknown parameters using the experimental values of )(1 th and )(2 th .

A least-square method was used for the curve fitting. For the Casson model, there

were three unknown values, which were k , sth∆ , and yτ . Note that the unknown

values were assumed to be constant for the curve-fitting method. Since )(1 th and

Page 85: A Study of Non-Newtonian Viscosit y and Yield Stress of

70

)(2 th are strongly dependent on each other by the conservation of mass for

incompressible fluids, dt

tdh )(1 must be equal to dt

tdh )(2− . Therefore, it was more

convenient and accurate to use the difference between the velocities at the two riser

tubes, i.e., dt

ththd ))()(( 21 − , than to use dt

tdh )(1 and dt

tdh )(2 directly. In order to get

the difference between the two velocities from the experimental values, one could use

the central differential method as follows:

( ) [ ] [ ]t

tthtthtthtthdt

ththd∆

∆−−∆−−∆+−∆+=

−2

)()()()()()( 212121 . (4-40)

Using Eq. (4-39), the derivative of the velocity difference can be determined

theoretically as follows:

( )

]))()((211

34

)))()(((7

16))()([(4

)(2)()(

321

4

21

21212

4

21

−∆−−∆−∆+

∆−−∆−∆−−=

=−

styy

stystcr

c

r

hththhh

hththhhththLkRgR

tVdt

ththd

ρ

(4-41)

where )(tVr is the mean flow velocity at the riser tube.

In order to execute the curve-fitting procedure, one needs to have a

mathematical equation of rV for the Casson model. Eq. (4-40) and (4-41) were used

for the curve fitting of the experimental data to determine the unknown constants, i.e.,

k , sth∆ , and yh∆ . Note that Eq. (4-41) could be applicable for both Casson-model

Page 86: A Study of Non-Newtonian Viscosit y and Yield Stress of

71

fluids and Newtonian fluids regardless of the existence of the yield stress. The data

reduction procedure adopted is as follows:

1. Conduct a test and acquire all data, )(1 th and )(2 th .

2. Guess values for the unknowns, k , sth∆ , and yh∆ .

3. Calculate the following error values for all data points.

[ ]2 )(2)(2 valueslTheoreticavaluesalExperiment tVtVError −= (4-42)

4. Sum the error values for all data points.

5. Iterate to determine the unknowns that minimize the sum of the error.

6. Calculate wall shear rate and viscosity for all data points as follows:

( ) 2

21 )()(2

)( ystc

cw hhthth

kLgR

t ∆−∆−−=ρ

γ& , (4-43)

[ ]cw

stcw Lt

hththgRt

)( 2)()(

)( 21

γρ

η&

∆−−= . (4-44)

Note that when yh∆ becomes approximately zero (i.e., ≤ resolution of 5103.8 −× ), the

non-Newtonian viscosity, η , is reduced to k , a Newtonian viscosity. Furthermore,

the relation between wall viscosity and shear-rate can be obtained from Eqs. (4-43)

and (4-44) as follow:

)(

2

)(2

)(

4

)()(

tL

hgRk

tL

hgR

k

t

k

tkt

w

c

yc

w

c

yc

w

y

w

yw

γ

ρ

γ

ρ

γ

τ

γτ

η

&&

&&

+

+=

++=

(4-45)

where k and yh∆ are the fluid properties to be determined using the Casson model.

Page 87: A Study of Non-Newtonian Viscosit y and Yield Stress of

72

Yield stress could be also determined through the curve-fitting method from

the experimental data of )(1 th and )(2 th by using the Casson model. Since the

pressure drop across the capillary tube, )(tPc∆ , could be determined using Eq. (4-8),

)(∞∆ cP represents the effect of the yield stress on the pressure drop. The relationship

between the yield stress, yτ , and )(∞∆ cP can be written by the following equation:

c

cy

c

ccy L

RhgL

RP22

)( ⋅∆=

⋅∞∆=

ρτ . (4-46)

Therefore, once yh∆ is obtained using a curve-fitting method, the yield stress can be

automatically determined.

4.2.3 Herschel-Bulkley (H-B) Model

For a Herschel-Bulkley (H-B) model, the shear stress at the capillary tube can

be described as follows [Tanner, 1985; Ferguson and Kemblowski, 1991; Macosko,

1994]:

ynm τγτ += & when yττ ≥ , (4-47)

0=γ& when yττ ≤ , (4-48)

where

τ and γ& = shear stress and shear rate, respectively

yτ = a constant that is interpreted as yield stress

m and n = model constants.

Page 88: A Study of Non-Newtonian Viscosit y and Yield Stress of

73

Since the H-B model reduces to the power-law model when a fluid does not have a

yield stress, the H-B model is more general than the power-law model.

For the H-B model, wall shear stress and yield stress can also be defined as

follows:

c

ccw L

RtP2

)( ⋅∆=τ , (4-49)

c

cy

c

cc

c

ycy L

RhgL

RPL

trtP22

)(2

)()( ⋅∆=

⋅∞∆=

⋅∆=

ρτ , (4-50)

where yr is a radial location below which the velocity profile is uniform due to the

yield stress (see Fig. 4-7). Using Eqs. (4-47)-(4-50), one can obtain the expressions

of shear-rate outside of the core region as:

( ) ny

n

c

cc rrmLP

drdV 1

1

2−

∆=−=γ& for cy Rrtr ≤≤)( . (4-51)

The velocity profile outside of core region can be obtained by integrating Eq. (4-51)

as:

( ) ( )

−−−

∆⋅

+=

++n

n

yn

n

yc

n

c

cc trrtrR

mLtP

nnrtV

111

)()(2

)(1

),(

for cy Rrtr ≤≤)( . (4-52)

Since the velocity profile inside of the core region is a function of time, t , only, the

profile can be obtained using a boundary condition, )(),( tVrtV cc = at yrr = .

( ) nn

yc

n

c

cc trR

mLtP

nntV

11

)(2

)(1

)(+

∆⋅

+= for rtry ≥)( . (4-53)

Page 89: A Study of Non-Newtonian Viscosit y and Yield Stress of

74

Again, for the purpose of simplicity, one may define two new parameters, cR

rrC =)(

and c

yy R

trtC

)()( = , so that Eqs. (4-52) and (4-53) become as follows:

( ) ( )

−−−

∆⋅

+=

−−

∆⋅

+=

+++

+++

nn

yn

n

y

n

c

cnc

nn

c

y

c

nn

c

yn

c

cnc

c

tCrCtCmL

tPRn

n

Rtr

Rr

Rtr

mLtPR

nnrtV

111

1

1111

)()()(12

)(1

)()(1

2)(

1),(

for cy Rrtr ≤≤)( , (4-54)

( ) nn

y

n

c

cnc

nn

c

yn

c

cnc

c

tCmL

tPRn

n

Rtr

mLtPR

nntV

11

1

111

)(12

)(1

)(1

2)(

1)(

++

++

−⋅

∆⋅

+=

−⋅

∆⋅

+=

for rtry ≥)( . (4-55)

In order to determine the mean flow velocity at the riser tube, one has to find

the flow rate at the capillary tube first. The flow rate can be obtained by integrating

the velocity profile over the cross-sectional area of the capillary tube as follows:

( ) ( ) ( )

( ) ( ) ]13

212

2

[12

2)(

1312

1212

1

0

nn

ycnn

ycy

nn

ycycn

n

ycy

n

c

c

R

c

rRnnrRr

nn

rRrRrRrn

nmLP

drrVtQ c

++

++

+−−

+−

−⋅++−⋅

+

∆=

⋅= ∫

π

π

(4-56)

Page 90: A Study of Non-Newtonian Viscosit y and Yield Stress of

75

( ) ( ) ( )

( ) ( ) ]113

2112

2

111[21

]113

2112

2

111[12

1312

1212

113

1313

1213

1213

12131

nn

ynn

yy

nn

yyn

n

yy

n

c

cnn

c

nn

c

ynn

c

nn

c

y

c

ynn

c

nn

c

y

c

ynn

c

nn

c

y

c

ynn

c

n

c

c

CnnCC

nn

CCCCmLP

nnR

Rr

nnR

Rr

Rr

nnR

Rr

Rr

RRr

Rr

Rn

nmLP

++

+++

++

++

++

++

+−−

+−

−⋅++−

∆⋅

+=

+−

+−

−⋅

++

+

∆=

π

π

where st

y

w

y

c

yy hthth

htP

PtR

trtC

∆−−

∆=

∆∞∆

===)()()(

)()(

)()(

21ττ

.

Since )()( 2 tVRtQ rrπ= , the mean flow velocity at the riser tube can be

determined by the following equation:

( ) ( ) ( )

( ) ( ) ]113

2112

2

111[21

)(

1312

1212

1

2

13

nn

ynn

yy

nn

yyn

n

yy

n

c

c

r

nn

cr

CnnCC

nn

CCCCmLP

nn

RR

tV

++

+++

+−−

+−

−⋅++−

∆⋅

+=

(4-57)

Equation (4-57) can be rewritten to clearly display the unknowns and the observed

variables as follows:

Page 91: A Study of Non-Newtonian Viscosit y and Yield Stress of

76

[ ]

])()(

113

2

)()(1

)()(122

)()(1

)()(1

)()(1

)()( [

2)()(

1)(

13

21

12

2121

12

2121

1

21

2

21

1

212

13

nn

st

y

nn

st

y

st

y

nn

st

y

st

y

nn

st

y

st

y

n

c

st

r

nn

cr

hththh

nn

hththh

hththh

nn

hththh

hththh

hththh

hththh

mLhththg

nn

RR

tV

+

+

+

+

+

∆−−

∆−⋅

+−

∆−−

∆−⋅

∆−−

∆⋅

+−

∆−−

∆−⋅

∆−−

∆++

∆−−

∆−⋅

∆−−

∆×

∆−−⋅

+=

ρ

(4-58)

Note that Eq. (4-58) contains two independent variables, i.e., )(1 th and )(2 th , and

one dependent variable, i.e., )(tVr . There are four unknown parameters to be

determined through the curve fitting in Eq. (4-58), namely m , n , sth∆ , and yh∆ .

Once the equation for the mean flow velocity, )(tVr , was derived, one could

determined the unknown parameters using the experimental values of )(1 th and )(2 th

by using the same curve-fitting method of determining unknowns as in the case of the

Casson model. In the case of the H-B model, there were four unknown values, which

were m , n , sth∆ , and yh∆ . Note that the unknown values were assumed to be

constant for the curve-fitting method.

Using Eq. (4-58), the derivative of the velocity difference can be determined

theoretically as follows:

Page 92: A Study of Non-Newtonian Viscosit y and Yield Stress of

77

( )

[ ]

])()(

113

2

)()(1

)()(122

)()(1

)()(1

)()(1

)()( [

2)()(

12

)(2)()(

13

21

12

2121

12

2121

1

21

2

21

1

212

13

21

nn

st

y

nn

st

y

st

y

nn

st

y

st

y

nn

st

y

st

y

n

c

st

r

nn

c

r

hththh

nn

hththh

hththh

nn

hththh

hththh

hththh

hththh

mLhththg

nn

RR

tVdt

ththd

+

+

+

+

+

∆−−

∆−⋅

+−

∆−−

∆−⋅

∆−−

∆⋅

+−

∆−−

∆−⋅

∆−−

∆++

∆−−

∆−⋅

∆−−

∆×

∆−−⋅

+=

=−

ρ

(4-59)

where )(tVr is the mean flow velocity at the riser tube. In order to execute the curve-

fitting procedure, one needs to have a mathematical equation of rV for the H-B model.

Eq. (4-58) and (4-59) were used for the curve fitting of the experimental data to

determine the unknown constants, i.e., n , m , sth∆ , and yh∆ . Note that Eq. (4-59)

could be applicable for H-B fluids, Shear-thinning fluids, and Newtonian fluids

regardless of the existence of the yield stress.

After iterations for the determination of the unknowns that minimize the sum

of the error, wall shear rate and viscosity for all data points can be calculated as

follows:

( )[ ] nyst

n

c

cw hhthth

mLgR

t1

21

1

)()(2

)( ∆−∆−−⋅

=

ργ& , (4-60)

Page 93: A Study of Non-Newtonian Viscosit y and Yield Stress of

78

[ ]cw

stcw Lt

hththgRt

)( 2)()(

)( 21

γρ

η&

∆−−= . (4-61)

Note that when yh∆ becomes approximately zero (i.e., ≤ resolution of 5103.8 −× ), the

H-B model is reduced to power-law model. In addition, when n becomes 1, the

mathematical form of the H-B model yields to Bingham plastic [Tanner, 1985], which

can be described as follows:

yBm τγτ += & when yττ ≥ , (4-62)

0=γ& when yττ ≤ , (4-63)

where

τ and γ& = shear stress and shear-rate, respectively

yτ = a constant that is interpreted as the yield stress

Bm = a model constant that is interpreted as the plastic viscosity.

Similar to the Casson model, the relationship between wall viscosity and

shear-rate using the H-B can be expressed as follows:

)(2

)(

)()()(

1

1

tL

hgR

tm

ttmt

w

c

yc

nw

w

ynww

γ

ρ

γ

γτ

γη

&&

&&

+=

+=

(4-64)

where m , n , and yh∆ are the fluid properties to be determined using the H-B model.

Page 94: A Study of Non-Newtonian Viscosit y and Yield Stress of

79

(a) Motion of a cylindrical fluid element within a capillary tube.

(b) Free-body diagram of a cylinder of fluid.

Fig. 4-5. Fluid element in a capillary tube at time t .

Capillary tube

r cR

l

Flow direction

l

rlπτ 2

2rPπ 2)( rPP π∆−

Page 95: A Study of Non-Newtonian Viscosit y and Yield Stress of

80

Fig. 4-6. Velocity profile of plug flow of blood in a capillary tube.

Capillary tube

yr cR

Page 96: A Study of Non-Newtonian Viscosit y and Yield Stress of

81

CHAPTER 5. CONSIDERATIONS FOR EXPERIMENTAL STUDY

Chapter 5 presents the issues and considerations in the experimental study

with a scanning capillary-tube rheometer (SCTR). Theoretical and experimental

issues involved in the viscosity and yield stress measurements of fluids, such as

distilled water, bovine blood, and human blood, are examined.

Sections 5.1, 5.2, and 5.3 address the major assumptions in the study that may

affect the rheological measurements in the SCTR: unsteady effect, end effect, and

wall effect, respectively. In addition, section 5.4 reports other possible factors that

include pressure drop at riser tubes, effect of density variation of blood, and

thixotropic effect.

In section 5.5, the temperature consideration for the viscosity measurement of

human blood in the SCTR during a test is discussed. The temperature of human

blood was checked to see if it could be maintained at a body temperature of 37

during a viscosity measurement.

In section 5.6, the study on the effect of dye concentration on the viscosity of

distilled water is presented. The objective of the study was to see whether or not the

viscosity of distilled water could be altered by the addition of dye.

Page 97: A Study of Non-Newtonian Viscosit y and Yield Stress of

82

5.1. Unsteady Effect

In order to make the assumption of a quasi-steady flow behavior during a test

with the SCTR, one needs to make sure that the pressure drop due to the unsteady

state is negligibly small compared to that due to the friction through a capillary tube.

Distilled water and bovine blood were analyzed for the unsteady effects on the

viscosity measurements of the fluids.

Figure 5-1 shows the pressure drops due to both unsteady flow and friction at

the capillary tube obtained by using Eqs. (4-4), (4-6), and (4-8) in the case of distilled

water. Usually, the pressure drop due to the unsteady flow was less than 3 Pa at the

beginning of a test while pressure drop due to the friction at the capillary tube was

greater than 250 Pa (see Table 5-1).

As shown in Fig. 5-2, in the case of bovine blood, the pressure drop due to the

unsteady flow was also much smaller than that at the capillary tube. Typically, the

pressure drop due to the unsteady flow was less than 1.2 Pa at the beginning of a test

while pressure drop due to the friction at the capillary tube was greater than 700 Pa.

Furthermore, as shown in Table 5-2, the magnitude of the pressure drop due to the

unsteady flow, unsteadyP∆ , was always less than 1% of that of the pressure drop at the

capillary tube, cP∆ , over the entire shear-rate range. This confirms that the

assumption of a quasi-steady state could be used for the present data reduction

procedure.

Page 98: A Study of Non-Newtonian Viscosit y and Yield Stress of

83

(a)

(b)

Fig. 5-1. Pressure drop estimation for distilled water. (a) Pressure drop due to an unsteady flow in a test with distilled water. (b) Pressure drop at a capillary tube in a test with distilled water.

0

1

2

3

4

0 10 20 30

Time (s)

Pre

ssur

e dr

op (P

a)unsteadyP∆

0

100

200

300

400

0 10 20 30

Time (s)

Pre

ssur

e dr

op (P

a)

cP∆

Page 99: A Study of Non-Newtonian Viscosit y and Yield Stress of

84

Table. 5-1. Comparison of unsteadyP∆ and cP∆ for distilled water.

Time (s) unsteadyP∆ (Pa) cP∆ (Pa) 100×∆

c

unsteady

PP

(%)

0.5 2.89 245.46 1.18

1 2.61 222.08 1.18

3 1.56 132.60 1.18

5 0.93 79.40 1.17

10 0.26 22.98 1.13

15 0.07 6.85 1.02

20 0.02 2.02 0.99

Page 100: A Study of Non-Newtonian Viscosit y and Yield Stress of

85

(a)

(b)

Fig. 5-2. Pressure drop estimation for bovine blood. (a) Pressure drop due to an unsteady flow in a test with bovine blood. (b) Pressure drop at a capillary tube in a test with bovine blood.

0

0.4

0.8

0 30 60 90 120 150

Time (s)

Pre

ssur

e dr

op (P

a)

unsteadyP∆

0

200

400

600

800

0 30 60 90 120 150

Time (s)

Pre

ssur

e dr

op (P

a)

cP∆

Page 101: A Study of Non-Newtonian Viscosit y and Yield Stress of

86

Table. 5-2. Comparison of unsteadyP∆ and cP∆ for bovine blood.

Time (s) unsteadyP∆ (Pa) cP∆ (Pa) 100×∆

c

unsteady

PP

(%)

0.5 0.54 705.21 0.08

1 0.52 688.11 0.08

5 0.39 535.93 0.07

10 0.28 399.14 0.07

30 0.067 139.24 0.05

60 0.008 42.63 0.02

120 ≈0 15.27 ≈0

Page 102: A Study of Non-Newtonian Viscosit y and Yield Stress of

87

5.2. End Effect

Figure 5-3 shows the flow-pattern changes due to end effects at both (a)

entrance and (b) exit of a capillary tube. Due to the sudden contraction and expansion,

additional pressure drops can occur at the both ends. The most common method used

to estimate these minor pressure drops is to use the loss coefficient, LK , which is

defined as [Munson et al., 1998]:

2

21

c

EndL

V

PKρ

∆= (5-1)

so that

2

21

cLEnd VKP ρ=∆ (5-2)

where EndP∆ is the pressure drop due to the end effects and cV is the mean velocity at

the capillary tube.

With the present experimental set-up, the velocity in the capillary tube was

approximately 16 times greater than that in the riser tube. Therefore, the energy loss

by secondary flow patterns or eddies in the entrance and exit of the capillary tube

may appear to be significant in a high shear zone. In the case of a laminar flow, the

loss coefficient was reported to be approximately 2.24 [Ferguson and Kemblowski,

1991]. Using the value of the loss coefficient, the pressure loss due to the sudden

changes in geometry, EndP∆ , became only 1.79 Pa (for distilled water) and 1.88 Pa

(for bovine blood) for the maximum shear rate of 400 s-1 at a corresponding velocity

of 0.04 m/s. In contrast, the pressure drops across the capillary tube at the maximum

Page 103: A Study of Non-Newtonian Viscosit y and Yield Stress of

88

shear rate were 245 Pa (for distilled water) and 705 Pa (for bovine blood), indicating

that the loss due to the secondary flow patterns or eddies at both entrance and exit

could be neglected.

In these end regions (see Fig. 5-3), the flow is changing from (or to) its

previous (or future) distribution outside the capillary tube. The length of an end

region is generally a function of tube geometry and some dynamic parameters. The

entrance length, the length of tube required to achieve the fully developed simple

shear flow, can be estimated by using the following equation [Middleman, 1968]:

Re035.0 ⋅≈DLe (5-3)

where eL is the entrance region, mm) 0.8( ≈D is the inner diameter of a capillary

tube, and Re is the Reynolds number. The maximum Reynolds number from a

typical run in the present study was approximately 28.6. The entrance length in the

capillary tube used for the present study was estimated to be 0.0008 m using the

above equation. Generally, the ratio of the entrance length to the capillary-tube

length, c

e

LL

, should be the order of 0.01 in order to assume the effect of entrance

length to be negligible [Middleman, 1968]. Since the ratio was 0.008 in the present

study, it is reasonable to assume that the entrance length effect is negligibly small.

Page 104: A Study of Non-Newtonian Viscosit y and Yield Stress of

89

(a) Sudden contraction at entrance of a capillary tube

(b) Sudden expansion at exit of a capillary tube

Fig. 5-3. Flow-pattern changes due to end effects [Munson et al., 1998].

Page 105: A Study of Non-Newtonian Viscosit y and Yield Stress of

90

5.3. Wall Effect (Fahraeus-Lindqvist Effect)

Apart from end effects, other sources of error in a capillary-tube rheometry

should be considered. The wall effect is one of the most important error sources

[Barnes, 1995; Missirlis et al., 2001]. For example, during the flow of a suspension, a

thin layer of the solvent whose viscosity is lower than the viscosity of the suspension

solution may be formed near the capillary wall, and this wall effect becomes more

significant with the decrease in the capillary diameter [Dinnar, 1981; Ferguson and

Kemblowski, 1991; Fung, 1993].

In the case of blood flow, as shown in Fig. 5-4, the wall effect can be

described as a tendency for RBCs to move toward the center of the tube or blood

vessel [Thomas, 1962; Picart et al., 1998]. The plasma-rich zone next to the solid

wall, although very thin, has an important effect on blood rheology. In other words,

the plasma-rich layer near the wall must affect the measurement of blood viscosity by

any instrument with a solid wall. The reduction in the RBC concentration in this

layer near the wall decreases the measured value of blood viscosity, resulting in

erroneous viscosity results.

Thus, the apparent viscosity of whole blood decreases with the decrease in

tube diameter. However, as shown in Fig. 5-5, the wall effect is reported to be

negligibly small when the tube diameter is greater than approximately 0.4 mm

[Fahraeus R and Lindquist, 1931; Dintenfass, 1971; Dinnar, 1981; Stadler et al.,

1990; Pries et al., 1992] or 0.8 mm [Haynes, 1960; Barbee, 1971]. Note that the

Page 106: A Study of Non-Newtonian Viscosit y and Yield Stress of

91

apparent viscosity of blood decreases to a value close to plasma viscosity if the

diameter of the capillary tube decreases below 0.1 mm [Benis et al., 1970].

In order to check whether or not the present capillary tube diameter (with

0.797 mm ID) was large enough to prevent the wall effect, two additional capillary

tubes, whose diameters were 1.0 mm (with length = 130 mm) and 1.2 mm (with

length = 156 mm), were used for the viscosity measurements of bovine blood with

7.5% EDTA at a room temperature of 25. As shown in Fig. 5-6, the experiments

performed with three different capillary tubes with ID of 0.797 mm (the standard size

of the SCTR), 1.0 mm, and 1.2 mm provided almost identical viscosity results,

confirming that the wall effect was negligibly small for the present capillary tube with

ID equal to 0.797 mm.

Page 107: A Study of Non-Newtonian Viscosit y and Yield Stress of

92

Fig. 5-4. Migration of cells toward to the center of lumen (wall effect).

Flow

Cell-free region RBCs

Arterial wall

Page 108: A Study of Non-Newtonian Viscosit y and Yield Stress of

93

Fig. 5-5. Fahraeus-Lindquist effect due to the reduction in hematocrit in a tube with a small diameter and the tendency of erythrocytes to migrate toward the center of the tube [Fahraeus and Lindquist, 1931; Dintenfass, 1971; Dinnar, 1981; Stadler et al., 1990; Pries et al., 1992]

Tube diameter (microns)

10

4

1

Blood viscosity (cP)

400100 800

Page 109: A Study of Non-Newtonian Viscosit y and Yield Stress of

94

1

10

100

1 10 100 1000Shear rate (s-1)

Vis

cosi

ty (c

P)

0.797 mm1.0 mm1.2 mm

Fig. 5-6. Viscosity measurements for bovine blood with three different capillary tubes with ID of 0.797 mm (with length = 100 mm), 1.0 mm (with length = 130 mm), and 1.2 mm (with length = 156 mm).

Page 110: A Study of Non-Newtonian Viscosit y and Yield Stress of

95

5.4. Other Effects

5.4.1. Pressure Drop at Riser Tube

Since the small diameter of a capillary tube was selected to make sure that the

pressure drop at the capillary tube could be dominant, the pressure drops at riser tubes

should be negligibly small. It has been suggested that the pressure drop in the

reservoir should be estimated by using a power-law model as [Marshall and Riley,

1962; Metzger and Knox, 1965; Macosko, 1994]:

nn

c

rc

rcr

RRL

LPP 3

)(+

∆=∆ (5-4)

where

rP∆ = pressure drop in reservoir

rL = wetted length in reservoir

rR = radius of reservoir.

The power-law model is one of the simplest models, which can be used to

show non-Newtonian behavior of blood. Furthermore, the power-law model

generally provides almost identical viscosity results with both Casson and Herschel-

Bulkley models at the shear rates between approximately 300 and 30 s-1. The

viscosity results of blood obtained with those models will be discussed in Chapter 6

in detail. Typically, the power-law index, n , for healthy human blood is 0.75-0.85 at

a body temperature of 37.

Page 111: A Study of Non-Newtonian Viscosit y and Yield Stress of

96

Considering the reservoir in Eq. (5-4) as the riser tubes in the present system

and n = 0.8 for human blood, one can obtain the following relation between pressure

drops at capillary and riser tubes using Eq. (5-4).

cr PP ∆≈∆500

1 (5-5)

Therefore, in the case of human blood, the sum of the pressure drops at riser tubes is

approximately 0.2 Pa at a shear rate of 30 s-1 while the pressure drop at the capillary

tube is approximately 93 Pa.

It could be argued that Casson or Herschel-Bulkley model would have a larger

pressure drop than the power-law model at a lower shear rate. Therefore, we want to

examine whether or not the pressure drop at the riser tube is still negligibly small for

Casson or Herschel-Bulkley model compared to that at the capillary tube at a very

low shear rate by looking at the upper bound of the error. It is rather obvious that the

pressure drop at the riser tube at a low shear rate (i.e., below 30 s-1) should be smaller

than 0.2 Pa. Let’s consider a shear rate of 1 s-1. The pressure drop at the capillary

tube at γ& = 1 s-1 is approximately 15 Pa. Therefore, the pressure drop at the riser tube

is less than 1.33% (i.e., 0133.015

2.0= ). Hence, it is reasonable to assume that the

pressure drops at the riser tubes can be ignored compared to the pressure drop at the

capillary tube.

Page 112: A Study of Non-Newtonian Viscosit y and Yield Stress of

97

5.4.2. Effect of Density Variation

In order to measure the viscosity of blood by using the SCTR, one needs to

know the density of blood. However, in the case of human blood, it is not very

convenient to measure the density of blood for each viscosity measurement.

Therefore, the following relation [Chien et al., 1987] between hematocrit (Hct as a

dimensionless fraction) and blood density ( ρ in kg/m3) was used for the estimation

of the density of human blood.

Hct671026 +=ρ (5-6)

Table 5-3 shows the density of blood corresponding to hematocrit. In normal

hematocrit concentrations, i.e., 35-45% [Guyton and Hall, 1996], the density variation

is less than 1%, which barely affects the viscosity results of blood.

5.4.3. Aggregation Rate of RBCs – Thixotropy

As discussed in Chapter 2, the thixotropic effect on blood viscosity may be

more significant at low shear rates than high shear rates. In the SCTR, the shear rate

varies from high (approximately 400 s-1) to low (1 s-1) values. At least 50 seconds is

required during a test to have the fully aggregated quiescent state at a shear rate near

1 s-1 [Gaspar-Rosas and Thurston, 1998].

In the viscosity measurement of human blood with the SCTR, a typical test

duration in which the shear rate decreased from 10 to 1 s-1 was longer than 60 seconds.

Page 113: A Study of Non-Newtonian Viscosit y and Yield Stress of

98

It is reasonable to assume that the 60-second period is long enough to cause

aggregations if the aggregations were going to take place. To further validate the

above assumption, a longer capillary tube (125-mm length) was used. Since the test

duration increased in the longer capillary tube, an anticoagulant (7.5% EDTA) was

added to avoid the blood clotting. As shown in Fig. 5-7, the viscosity results obtained

by using the longer capillary tube showed excellent agreements with those obtained

by using the capillary tube with 100-mm length (also with 7.5% EDTA). Therefore,

it is concluded that, in the present system, the thixotropic effect of blood on the

viscosity measurement is negligibly small.

Page 114: A Study of Non-Newtonian Viscosit y and Yield Stress of

99

Table. 5-3. Density estimation

Hematocrit (%) Density

(kg/m3)

35 1049.5

40 1052.8

45 1056.2

50 1059.5

Page 115: A Study of Non-Newtonian Viscosit y and Yield Stress of

100

1

10

100

1 10 100 1000Shear rate (s-1)

Vis

cosi

ty (c

P)

100 mm125 mm

Fig. 5-7. Viscosity results for human blood with two different capillary tubes with length of 100 mm (with ID = 0.797 mm) and 125 mm (ID = 0.797 mm).

Page 116: A Study of Non-Newtonian Viscosit y and Yield Stress of

101

5.5. Temperature Considerations for Viscosity Measurement of Human Blood

For unadulterated human blood, the temperature of a SCTR was controlled

during the test at a body temperature of 37°C by using preheated disposable tube sets

and a heating pad installed inside the SCTR. In order to check whether or not the

temperature of blood was maintained at a body temperature, a special U-shaped tube

set was prepared for the experiment. Figure 5-8 shows the special U-shaped tube set

which is basically the same as a standard U-shaped tube set except three additional

thermocouples placed at both ends and on the outside surface of the capillary tube.

The temperatures of blood at three predetermined points were measured

during a viscosity test. Figure 5-9 provides the temperature measurement results for

human blood during the test. The temperature of blood at the exit of the capillary

tube was maintained at approximately 38, whereas that at the entrance was

gradually increased from about 36 at the beginning of the test and reached 37.5

at the end of the test. Therefore, it is reasonable to say that the temperature of human

blood flowing through the capillary tube during the test was maintained at a body

temperature of 37 with ± 1.

Page 117: A Study of Non-Newtonian Viscosit y and Yield Stress of

102

Fig. 5-8. Schematic diagram of a U-shaped tube set for temperature measurement.

Thermometer

Entrance Thermocouple

Capillary Surface Thermocouple

Exit Thermocouple

Page 118: A Study of Non-Newtonian Viscosit y and Yield Stress of

103

30

32

34

36

38

40

0 50 100 150 200

Time (s)

Tem

pera

ture

( )

EntranceExitCapillary Surface

Fig. 5-9. Temperature measurement at a capillary tube during a viscosity test.

Page 119: A Study of Non-Newtonian Viscosit y and Yield Stress of

104

5.6. Effect of Dye Concentration on the Viscosity of Water

5.6.1. Introduction

Figure 5-10 shows a schematic diagram of a SCTR, which consists of two

charge-coupled devices (CCDs) that are positioned vertically, two light-emitting

diodes (LEDs), two riser tubes and a capillary tube, a stopcock, and a data-acquisition

system. The essential feature in the SCTR is to use two riser tubes, where initial fluid

levels are different: one riser tube has a higher fluid level than the other one. Thus, at

t = 0, the fluid begins to fall from the riser tube with the high level to the riser tube of

low level by gravity. Since the flow rate depends on the pressure head between the

two fluid levels, the flow rate gradually decreases with time as the difference between

the two fluid level decreases with time. Since the flow rate can be estimated from the

time rate of change of the fluid level, one can estimate both flow rate and pressure

drop from the measurement of two fluid levels. Then, one can calculate shear rate

from the flow rate data and shear stress from the pressure drop data, respectively.

From the shear rate and shear stress, one can determine the viscosity of the liquid.

Thus, the most important experimental variable in the operation of the SCTR

is the measurement of two fluid levels in the riser tubes. As shown in Fig. 5-10, the

present SCTR uses an optical detector (i.e. CCD sensors and LED array) to measure

the fluid-level variations in the riser tubes. The optical detector works as follows: as

an opaque fluid level rises in the riser tube, the opaque fluid blocks the passage of the

light emitted by the LED. Accordingly, the number of the CCD sensors that receive

Page 120: A Study of Non-Newtonian Viscosit y and Yield Stress of

105

the light from the LED becomes smaller. Computer software records the changes in

the number of CCD sensors that receive the light from the LED. Since the number of

the CCD sensors that don’t receive the light from the LED is directly proportional to

the fluid level, one can determine the fluid level. In other words, the instantaneous

fluid levels are recorded in the form of pixel numbers (i.e., CCD sensors) versus time

in a computer data file through an analog-to-digital data–acquisition system. The

fluid level data from the two riser tubes were analyzed to determine the viscosity of

the fluid.

Therefore, it is essential to have an opaque fluid for the present SCTR

operation so that the light from the LED can be blocked by the opaque fluid as the

fluid level increases, and vice versa. Of course, one can use a laser light so that a

transparent fluid can be used as demonstrated by Kim et. al. (2000b). However, the

cost of a SCTR using such a laser-based system became prohibitively expensive,

making such a system economically unattractive.

In order to use the SCTR using CCD-LED arrangements for the viscosity

measurement of a transparent fluid, one may add dye to the fluid in order to make the

fluid opaque. However, the addition of a dye to a transparent fluid may alter the

viscosity of the fluid. Furthermore, the addition of the dye may make a transparent

Newtonian fluid such as water a non-Newtonian fluid if the concentration of the dye

is sufficiently large [Kim and Cho, 2002].

Therefore, the objective of the study is to investigate the effect of dye

concentration on the viscosity of distilled water in the SCTR. More specifically, the

Page 121: A Study of Non-Newtonian Viscosit y and Yield Stress of

106

present study plans to determine the maximum concentration of dye below which the

viscosity of the dye-water solution is not altered.

5.6.2. Experimental Method

Although distilled water is a Newtonian fluid, the aqueous solution of dye-

water may exhibit the non-Newtonian characteristics for a sufficiently large dye

concentration. Thus, in order to investigate the viscosity characteristics of a dye-

water solution, the present study used a non-Newtonian model to reduce experimental

data. In the previous chapters, various non-Newtonian models have been introduced

for the determination of blood viscosity with the SCTR, which include power-law

model, Casson model, and Herschel-Bulkley (H-B) model.

However, it is not very convenient to use the Casson model when a fluid

shows only shear-thinning characteristics without yield stress. The H-B model is

reduced to a power-law model for the case of fluids with no yield stress. Thus, in the

present study, a power-law model was used for the viscosity analysis of the dye-water

solution. The procedure of data reduction with power-law model will be discussed in

Chapter 6. Therefore, only the experimental results will be provided and discussed in

the next section.

Page 122: A Study of Non-Newtonian Viscosit y and Yield Stress of

107

5.6.3. Results and Discussion

In this study, six different concentrations (0.5, 1, 2, 3, 4, and 7% by volume)

of dye were used for the viscosity measurement of dye-water solution at 25. The

dye in a liquid form was purchased at a grocery store, which was a vegetable dye

produced from brand name, McCormick. For the validation of the method to reduce

data for the SCTR, the viscosities of dye-water solution with different dye

concentrations were compared with well-accepted reference data for water at 25

[Munson et al., 1998].

Figure 5-11 shows the variations of both power-law and consistency indices

of the dye-water solution for six different dye concentrations. Both indices were

determined through a curve-fitting method. Rectangular symbols indicate the power-

law index whereas triangular symbols indicate the consistency index. As shown in

the figure, both indices started to vary when the amount of dye used became greater

than 2% by volume. When the dye concentration was less than 2%, the power-law

index was exactly one. The values of the consistency index for 0.5%, 1%, and 2% of

dye concentration cases were 0.890, 0.878, and 0.888, respectively. The distilled

water viscosity, which is given in the literature as a function of temperature, was

estimated to be 0.892 cP at 25 [Munson et al., 1998; Kim et al., 2002]. When the

dye concentration was greater than 2%, the power-law index decreased from n = 1 at

2% to n = 0.913 at 7%, whereas the consistency index increased from k = 0.89 at 2%

to k = 1.68 at 7%. The present results indicated that the effect of dye concentration

Page 123: A Study of Non-Newtonian Viscosit y and Yield Stress of

108

on the viscosity of the dye-water solution was negligibly small when the amount of

dye used was less than 2% by volume.

Figure 5-12 shows the viscosity data for the dye-water solution with six

different dye concentrations. At a high shear-rate of 500 s-1, even with high

concentrations (i.e., 3, 4, 7%) of dye, the results showed that the effect of dye

concentration on the viscosity of water was very small. However, at low shear-rates

such as 1 and 10 s-1, the viscosity of the dye-water solution dramatically increased

with increasing dye concentration.

In the present experiment, the maximum concentration of dye, under which

the viscosity of the dye-water solution did not change, was approximately 2% by

volume. Compared with the reference data for water at 25 [Munson et al., 1998],

the test results obtained with 0.5%, 1%, and 2% of dye concentrations gave less than

2% error in the entire shear-rate range.

Page 124: A Study of Non-Newtonian Viscosit y and Yield Stress of

109

Fig. 5-10. Schematic diagram of a scanning capillary-tube rheometer (SCTR) system.

Riser tube 2

LED array

CCD 2

CCD 1

Riser tube 1

Computer system for data

collection Test fluid

Capillary tube Three-way stopcock

Page 125: A Study of Non-Newtonian Viscosit y and Yield Stress of

110

0.84

0.88

0.92

0.96

1

1.04

0 1 2 3 4 5 6 7 8Dye concentration (%)

Pow

er-la

w in

dex,

n

0.6

0.8

1

1.2

1.4

1.6

1.8

Con

sist

ency

inde

x, k

(c

Ps

n-1 )

Fig. 5-11. Variations of both power-law index and consistency index of dye-water solution due to the effects of dye concentrations.

Page 126: A Study of Non-Newtonian Viscosit y and Yield Stress of

111

Fig. 5-12. Viscosity data for dye-water solution with 6 different dye concentrations at 25.

0

0.5

1

1.5

2

2.5

0 1 2 3 4 5 6 7 8

Dye concentration (%)

Vis

cosi

ty (c

P)

at 500 1/sat 100 1/sat 10 1/sat 1 1/s

Page 127: A Study of Non-Newtonian Viscosit y and Yield Stress of

112

CHAPTER 6. EXPERIMENTAL STUDY WITH SCTR

Chapter 6 presents the results of viscosity and yield stress measurements with

the scanning capillary-tube rheometer (SCTR). Experimental tests were performed

with mineral oil, distilled water, bovine blood with 7.5% EDTA, and unadulterated

human blood.

Section 6.1 provides the viscosity results of both mineral oil and human blood

produced with the SCTR (with precision glass riser tubes) using the power-law model

for data reduction.

Section 6.2 gives the test results of distilled water, bovine blood, and human

blood obtained with the SCTR (with plastic riser tubes). Casson and Herschel-

Bulkley models were used for data reduction to handle the yield stress of blood.

Section 6.3 reports the effect of the three models on the viscosity and yield-

stress measurements of blood with the SCTR as well as on the flow patterns of blood

such as velocity profile and wall shear stress in a capillary tube.

6.1. Experiments with SCTR (with Precision Glass Riser Tubes)

The present study measured the viscosity of unadulterated blood at body

temperature, 37. Blood is a fluid consisting primarily of plasma and cells such as

erythrocytes, leukocytes and platelets. Erythrocytes (i.e., red blood cells, RBC)

constitute the majority of the cellular content and account for almost one half of the

Page 128: A Study of Non-Newtonian Viscosit y and Yield Stress of

113

blood volume. The presence of such a high volume of red blood cells makes blood a

non-Newtonian fluid whose viscosity varies with shear rate. Whole blood viscosity

decreases as shear rate increases, a phenomenon called “shear-thinning

characteristics”. In other words, whole blood behavior may be described using a

power-law model, a Casson model, or a Herschel-Bulkley model. In the present

study, the power-law model was chosen for simplicity.

In order to demonstrate the validity of the scanning capillary-tube rheometer,

the viscosity data were compared with data obtained using a cone-and-plate rotating

viscometer. Since the rotating viscometer produces only 7 data points at relatively

high shear rates due to the torque requirement, the accuracy of the new scanning

rheometer at a low shear rate range was demonstrated by comparison with the

viscosity of a standard-viscosity oil (a Newtonian fluid) from Cannon Instrument

Company (State Park, PA).

6.1.1. Description of Instrument

Figure 6-1 shows a schematic diagram of the scanning capillary-tube

rheometer, which consists of two charge-coupled devices (CCDs) that are positioned

vertically, two light emitting diodes (LEDs), two riser tubes and a capillary tube both

made of precision glass, two stopcocks, a transfer tube made of tygon, and a

computer acquisition system. The inside diameter of both the transfer and riser tubes

Page 129: A Study of Non-Newtonian Viscosit y and Yield Stress of

114

used in the present tests were 3 mm. The inside diameter and length of the capillary

tube were 0.797 mm and 100 mm, respectively.

The essential feature in the scanning capillary-tube rheometer is the use of an

optical detector (i.e., CCD sensors and LED array) to measure the fluid level

variations in the riser tubes, )(1 th and )(2 th , every 0.02 s. The instantaneous fluid

levels were recorded in a computer data file through an analog-to-digital data

acquisition system in the form of pixel numbers vs. time. Since 12 pixels are equal to

1 mm, one could determine the actual height changes in the riser tube with an

accuracy of 0.083 mm.

6.1.2. Testing Procedure

Typical tests are conducted as follows: The system was turned on and

connected to a computer. The software on the computer was executed, and

communication with the viscosity measurement system was properly established. At

that point, the computer was ready to acquire data from the CCD sensors in the

system. The experimental test run was initiated with a venipuncture on the patient

using a 19-gauge stainless steel needle. Fresh blood was first directed from the first

stopcock to the second stopcock to collect blood into the syringe. About 5 ml of

blood was collected in the syringe for tests with a cone-and-plate viscometer

(Brookfield DV-III) and hematocrit measurements, and the syringe was then removed

from the system. Approximately 0.5 ml of this fresh blood from the syringe was

Page 130: A Study of Non-Newtonian Viscosit y and Yield Stress of

115

immediately transferred to the sample cup of the Brookfield rotating viscometer that

was maintained at a constant temperature of 37 by a water bath connected to the

cup.

The viscosity measurements with the rotating viscometer were completed

within approximately 1 minute from the time when the blood left the human body.

Blood clotting rapidly developed inside the cone-and-plate test section. The rate of

blood clotting with time critically depends on the thrombotic tendency of a particular

individual’s blood. As soon as blood began to clot, the rotating viscometer flashed an

“EEEE” sign indicating an overloaded torque, and thus tests were stopped. This

usually happened within 2 minutes of the test. During the viscosity measurement

with the Brookfield rotating viscometer, hematocrit values were determined with a

microhematocrit centrifuge (International Clinical Centrifuge).

Immediately following the removal of the syringe, the experiment with the

scanning capillary-tube rheometer was continued with the second stopcock turned to

a position to allow blood flow to both the capillary tube and riser tube 2. When blood

reached a predetermined height of 300 pixels in the riser tube 2, the second stopcock

was shut to stop further blood flow into the riser tube 2, and the first stopcock was

then turned to direct blood flow into the riser tube 1 up to a height of 1000 pixels. At

t = 0, the data acquisition system was enabled, and both stopcocks were adjusted to

allow blood to flow from the riser tube 1 to tube 2 as driven by the gravity head. Of

note is that the initial pixel difference of 700 was chosen to produce the maximum

shear rate of approximately 400 s-1. If a higher shear rate is desired, an initial pixel

difference greater than 700 can easily be selected.

Page 131: A Study of Non-Newtonian Viscosit y and Yield Stress of

116

For the purpose of calibration, the present study used the scanning capillary-

tube rheometer to measure the viscosity of mineral oil, which had a standard

Newtonian viscosity of 9.9 cP at 25. In the tests with human blood and mineral oil,

the capillary tube and major portions of the transfer tube in which test fluids were

actually flowing through the capillary tube were placed in a water bath maintained at

37 and 25, respectively.

6.1.3. Data Reduction with Power-law Model

The mathematical procedure for data reduction using a power-law model was

discussed in Chapter 4. The least-square method was used for curve fitting of the

experimental data and Eq. (4-18) in order to determine the power-law index, n , and

the consistency index, m . A standard software package (Excel-Solver, Microsoft;

see Appendix E), which has a formula known as a Newton’s method (see Appendix

F) [Microsoft Corporation, 57926-0694; Harris, 1998; John, 1998; Brown, 2001], was

used for iterations to determine the values of n and m that minimize the sum of error

(see Eq. (4-19)).

The analysis of data reduction for a mineral oil is introduced in Fig. 6-2.

Figure 6-2(a) shows both experimental values of )(tθ and theoretical values of )(tθ

that were obtained with initial guesses of the two unknowns, whereas Figure 6-2(b)

shows the curve-fitting result after iterations to minimize the sum of error. The initial

guesses for n and m were 0.8 and 8 (cP·sn-1), respectively, in the case of the mineral

Page 132: A Study of Non-Newtonian Viscosit y and Yield Stress of

117

oil (see Fig. 6-2(a)). The resulting values of n and m were determined to be 1 and

9.91 (cP·sn-1), respectively, by the iterations using the Excel-Solver.

The analysis of data reduction for human blood is shown in Fig. 6-3. For the

human blood case, initial guesses of the two unknowns of n and m were 0.8 and 6

(cP·sn-1), respectively (see Fig. 6-3(a)). After iterations to minimize the sum of error,

the unknowns, n and m , were determined to be 0.83 and 9.27 (cP·sn-1), respectively

(see Fig. 6-3(b)). The initial guesses and resulting values of n , m , and ∞=∆ th for

both mineral oil and human blood were reported in Table 6-1.

6.1.4. Results and Discussion

Both mineral oil and unadulterated human blood were used in the present

study. The former was specially ordered as a dyed viscosity-standard fluid (i.e., 9.9

cP at 25) from Cannon Instrument Company (State Park, PA), and the latter was

obtained from donors. For comparison purpose, the viscosity of the human blood was

also measured by using a cone-and-plate rotating viscometer (Brookfield model DV-

III) at 37. The rotating viscometer used in the present study had an LV-type spring

torque with a CP-40 spindle. In order to maintain the preset temperatures, a water

bath (PolyScience model 2LS-M) was used, which controlled the temperature with an

accuracy of 0.1.

Figures 6-4 and 6-5 show test results obtained with the mineral oil at 25.

Figure 6-4 shows the fluid level variations in the riser tubes, )(1 th and )(2 th . Both

Page 133: A Study of Non-Newtonian Viscosit y and Yield Stress of

118

fluid levels converge gradually from the initial fluid level difference and eventually

reach an equilibrium fluid level. In the case of mineral oil, 14.5 mm of an initial fluid

level difference was used to ensure that viscosity measurements at a low shear rate

range were accurate. It is of note that, for mineral oil, ∞=∆ th was found to be zero.

Figure 6-5 shows viscosity results for the mineral oil at 25 obtained with

the SCTR. The power-law index of the mineral oil was determined to be 1.0 by a

computer program (Excel-solver), confirming that it was a Newtonian fluid. Based

on the present viscosity measurement method, the viscosity of the mineral oil was

found to be between 9.86 and 9.91 cP at 25, which was a 0.5% difference in the

whole range of shear rates from the standard viscosity of 9.9 cP at the same

temperature.

Figure 6-6 shows height variations in each riser tube as a function of time for

fresh human blood at 37. In the case of human blood, about 58 mm of initial fluid

level difference was used for the viscosity measurement so that one could obtain the

accurate viscosity of human blood over a wide shear rate range as low as 1 s-1. In

order to finish a test without using anticoagulants, the test should be completed within

3-4 minutes. Otherwise, blood may begin to clot. In the present study, one test run

took less than 2 minutes. For human blood, the trends of fluid level variations were

very similar to those for mineral oil. However, ∞=∆ th for human blood was not zero

but a finite value, which depended on the individual donor. The minimum and

maximum values of ∞=∆ th were found to be 3.86 mm and 6.22 mm, respectively,

Page 134: A Study of Non-Newtonian Viscosit y and Yield Stress of

119

among 8 donors. These values of ∞=∆ th represent the thixotropic characteristics of

blood that result in the yield stress.

Figure 6-7 shows the viscosity of unadulterated human blood at 37, which

was measured with both the SCTR and the cone-and-plate rotating viscometer (RV).

Closed circle symbols indicate viscosity data measured with the SCTR while triangle

symbols indicate those measured with the RV. The viscosity of the unadulterated

human blood measured with the present SCTR was based on a calculation method

that determined the power-law index, n , and consistency index, m . In the case

shown in Fig. 5, the values of n and m were 0.828 (dimensionless) and 9.267

)s(cP 1-n⋅ , respectively.

Compared with the measured data using the RV, the present test results from

the SCTR gave excellent agreement with those measured by the RV (i.e., less than

5% difference) in a shear rate range between 30 and 375 s-1. However, as the shear

rate decreased below 30 s-1, the RV was not recommended by Brookfield for the

measurement of blood viscosity. More specifically, the shear stress should vary from

a minimum of 10% to 100% of the full range of the torque sensor used in the

rotational type viscometer at a given shear rate for reasonably accurate viscosity

measurements [Brookfield, 1999]. Therefore, the minimum shear rate at which the

RV could be used for the viscosity measurement of human blood was 30 s-1.

Blood clotting in the RV was the other reason that one could not obtain more

than 7 data points. One could see the effects of blood clotting on viscosity as testing

time passed beyond 1 minute with the RV. Since only 0.5 ml was used for the RV

test, the blood contact area with the surface of the cone-and-plate was much bigger

Page 135: A Study of Non-Newtonian Viscosit y and Yield Stress of

120

than that in the case of the SCTR, a condition that might have caused rapid blood

clotting.

Figure 6-8 shows the viscosity of unadulterated human blood for two different

donors at 37, whose hematocrits were Hct = 41 and 46.5. Furthermore, human

blood from 8 donors was tested for viscosity measurements in the present study.

Every result using SCTR gave good agreement with that from the RV at high shear

rates but had a different slope with respect to shear rate individually. The viscosity

for the case with Hct = 46.5 was consistently greater than that for the case with Hct =

41. The difference between the two viscosity data was very small at high shear rates

greater than 300 s-1 whereas the difference was significant (i.e., greater than 200%) at

a low shear range, indicating the significance of low shear viscosity data.

In fact, it is well known that slip at the wall occurs in the flow of two-phase

systems because of the displacement of the disperse phase away from solid surfaces

[Barnes, 1995; Picart et al., 1998a, 1998b]. In the case of blood, a significant amount

of slip appears at low shear rates when the size of RBC (red blood cells) is relatively

large compared to wall roughness. For a smooth geometry like a glass tube, however,

the slip effect begins to be considerable from as low as 0.5 s-1 [Picart et al., 1998a].

Therefore, whole blood that was used in the present study did not show large slip

effects since the lowest shear rate data used was 1 s-1. In order to get reliable

viscosity data below 0.5 s-1, it may be necessary to use a rough surface capillary.

Page 136: A Study of Non-Newtonian Viscosit y and Yield Stress of

121

Fig. 6-1. Schematic diagram of the scanning capillary-tube rheometer with precision glass riser tubes.

Blood Blood from vein

LED Array

CCD CCD

Riser tube 2 Riser tube 1

Computer system

First stopcock

Second stopcock

Water bath

Capillary tube

Transfer tube

Page 137: A Study of Non-Newtonian Viscosit y and Yield Stress of

122

(a) With initial guess values

(b) With final resulting values

Fig. 6-2. Curve-fitting procedure with power-law model for mineral oil.

0

4

8

12

16

0 20 40 60 80 100 120

Time (s)

Experimental data

Theoretical data)(

mmθ

n = 0.8 m = 8 (cP·sn-1)

0

4

8

12

16

0 20 40 60 80 100 120

Time (s)

Experimental data

Theoretical data)(

mmθ

n = 1 m = 9.91 (cP·sn-1)

Page 138: A Study of Non-Newtonian Viscosit y and Yield Stress of

123

(a) With initial guess values

(b) With final resulting values

Fig. 6-3. Curve-fitting procedure with power-law model for human blood.

0

20

40

60

0 20 40 60 80 100 120

Time (s)

Experimental data

Theoretical data

)(

mmθ

n = 0.8 m = 6 (cP·sn-1)

0

20

40

60

0 20 40 60 80 100 120

Time (s)

Experimental data

Theoretical data

)(

mmθ

n = 0.83 m = 9.27 (cP·sn-1)

Page 139: A Study of Non-Newtonian Viscosit y and Yield Stress of

124

Table. 6-1. Comparison of initial guess and resulting value using power-law model.

Distilled Water Human Blood

Initial Guess

n = 0.8 m = 8 (cP·sn-1)

∞=∆ th = 0.03 mm

n = 0.8 m = 6 (cP·sn-1)

∞=∆ th = 3 mm

Resulting Value

n = 1 m = 9.91 (cP·sn-1)

∞=∆ th = 0.0271 mm

n = 0.83 m = 9.27 (cP·sn-1)

∞=∆ th = 3.86 mm

Page 140: A Study of Non-Newtonian Viscosit y and Yield Stress of

125

Fig. 6-4. Height variation in each riser tube vs. time for mineral oil (9.9 cP viscosity-standard oil).

0

20

40

60

0 50 100 150

Time (s)

Hei

ght (

mm

)

Riser tube 1

Riser tube 2

Page 141: A Study of Non-Newtonian Viscosit y and Yield Stress of

126

6

8

10

12

0 20 40 60

Shear rate (s-1)

Vis

cosi

ty (c

P)

SCTR

Fig. 6-5. Viscosity measurement for mineral oil at 25 with a scanning capillary-tube rheometer (SCTR).

Page 142: A Study of Non-Newtonian Viscosit y and Yield Stress of

127

Fig. 6-6. Height variation in each riser tube vs. time for human blood at 37.

0102030405060708090

100

0 50 100 150

Time (s)

Hei

ght (

mm

) Riser tube 1

Riser tube 2

Page 143: A Study of Non-Newtonian Viscosit y and Yield Stress of

128

1

10

100

1 10 100 1000

Shear rate (s-1)

Vis

cosi

ty (c

P)

SCTVRV

Hematocrit : 40.5

Fig. 6-7. Viscosity measurement (log-log scale) for human blood at 37 with rotating viscometer (RV) and scanning capillary-tube rheometer (SCTR).

Page 144: A Study of Non-Newtonian Viscosit y and Yield Stress of

129

1

10

100

1 10 100 1000

Shear rate (s-1)

Vis

cosi

ty (c

P)

SCTVRVHematocrit : 46.5

Hematocrit : 41

Fig. 6-8. Viscosity measurement (log-log scale) of unadulterated human blood at 37, measured with scanning capillary-tube rheometer (SCTR) and cone-and-plate rotating viscometer (RV), for two different donors.

Page 145: A Study of Non-Newtonian Viscosit y and Yield Stress of

130

6.2. Experiments with SCTR (with Plastic Riser Tubes)

A new U-shaped scanning capillary-tube rheometer (SCTR) has been

developed from the concept of a conventional capillary-tube viscometer. In general,

the capillary-tube viscometer is an attractive technique for several reasons. The basic

instrument is relatively inexpensive, easy to construct and simple to use

experimentally. Temperature control is relatively easy. The flow in a capillary tube

most closely simulates the blood flow in physiological conditions compared with the

flows in rotating viscometers and falling-object viscometers. Most of all, since the

capillary-tube viscometer uses the gravity as the driving force, it does not require

periodic calibration of any components. However, significance of end effect, wall

effect, and surface-tension effect should be carefully considered.

The end and wall effects can be made negligibly small by selecting

appropriate dimensions for the capillary and riser tubes. However, unlike the case of

the SCTR with precision glass riser tubes, the effect of the surface tension is a unique

and critical factor in using U-shaped disposable tube sets. Since inexpensive

disposable capillary-riser tube sets should be used for clinical applications, it may not

be easy to strictly control the surface quality of riser tubes within a certain limit.

Therefore, the effects of the surface tension at the riser tubes as well as the properties

of a testing fluid had to be considered in the viscosity and yield stress measurements

with the SCTR.

The resistance associated with air-liquid interfaces in plastic riser tubes of a

small diameter in the SCTR can be a significant part of the pressure head applied to

Page 146: A Study of Non-Newtonian Viscosit y and Yield Stress of

131

the SCTR, particularly at low shear rates [Jacobs, 1966; Mardles, 1969; Einfeldt and

Schmelzer, 1982]. The meniscus resistance depends on the surface tension of the

fluid-air interface and on the reciprocal of the internal radius of the riser tube.

Throughout the development of a U-shaped scanning capillary-tube rheometer

concept, the focus has been on how to isolate the effects of surface tension and yield

stress in obtaining low-shear-rate viscosity for non-Newtonian fluids like blood. This

study attempted to measure the viscosity of unadulterated blood at a body temperature

of 37. In the present study, both Casson and Herschel-Bulkley models were

selected for the viscosity and yield stress measurements of blood since both models

have a yield stress term.

6.2.1. Description of Instrument

Figure 6-9 shows a photograph of the SCTR with plastic riser tubes, which

consists of two charge-coupled devices (CCD 1 and CCD 2) that are positioned

vertically, two light-emitting diodes (LEDs), two riser tubes made of acrylic plastic

and a capillary tube made of glass, a stopcock, and a data-acquisition system. The

inside diameter of the riser tubes used in the study was 3.2 mm. The inside diameter

and length of the capillary tube were 0.797 mm and 100 mm, respectively. The small

diameter of the capillary tube, compared with that of the riser tubes, was chosen to

ensure that the pressure drop at the capillary tube was significantly greater than those

at the riser tubes and connecting fittings.

Page 147: A Study of Non-Newtonian Viscosit y and Yield Stress of

132

6.2.2. Testing Procedure

Tests with distilled water and bovine blood were performed at the room

temperature of 25. Riser tube 1 was first filled with the test fluid to the

predetermined height of approximately 550 pixels by using a syringe. Once the

desired level was reached, the stopcock was turned to a position to allow test fluid to

flow to the capillary tube. When the fluid reached the predetermined height of 100

pixels in riser tube 2, the stopcock was shut to stop further fluid movement into riser

tube 2. Next, the syringe was removed from the SCTR system, and then the stopcock

was turned to allow fluid to move from riser tube 1 to riser tube 2 by gravity. As

shown in Fig. 6-9, two CCD sensors were used: one located at the lower (left) side

and the other located at the upper (right) side. The difference between the locations

of the two CCD sensors was 500 pixels, a number that was taken into account in the

height measurements.

For the purpose of comparison, approximately 0.5 ml of bovine blood from

the syringe was immediately transferred to a cone-and-plate test cup of a rotating

viscometer (Brookfield model DV-III with LV-type spring torque with CP-40

spindle) that was maintained at a constant temperature of 25 by a water bath

connected to the cup. For unadulterated human blood, the temperature of the SCTR

was controlled during the test at a body temperature of 37 by using preheated

disposable tube sets and a heating pad (shown in Fig. 6-10) installed inside the SCTR.

After the temperature in the SCTR was stabilized, the viscosity measurement was

initiated with a venipuncture using a 19-gauge stainless steel needle. About 5 ml of

Page 148: A Study of Non-Newtonian Viscosit y and Yield Stress of

133

extra blood was collected for hematocrit measurements, which were determined with

a microhematocrit centrifuge (International Clinical Centrifuge) for both bovine

blood and human blood.

Viscosities of distilled water (a Newtonian fluid), bovine blood containing

7.5% EDTA, and unadulterated human blood were measured over a range of shear

rates. Because the CCD sensor requires an opaque fluid, dye was added to the

distilled water. The amount of dye used for distilled water was less than 1%

concentration by volume, and the dye-effect on the viscosity of distilled water was

negligibly small at this concentration. Bovine blood was purchased from Lampire

Biological Laboratories, Inc., and human blood was obtained from two healthy male

donors who were 29 and 51 years of age. For comparison purposes, a reference value

was used for the distilled water while the viscosity of the bovine blood was

independently checked by using the cone-and-plate rotating viscometer.

6.2.3. Data Reduction with Casson Model

The fluid level data from the two riser tubes were analyzed to determine the

viscosities of distilled water (a Newtonian fluid) and blood (a non-Newtonian fluid).

In order to measure blood viscosity using a U-shaped SCTR, one needs to isolate the

effects of both surface tension and yield stress on the viscosity of blood. The details

of mathematical procedure for curve-fitting using the Casson model have already

Page 149: A Study of Non-Newtonian Viscosit y and Yield Stress of

134

been introduced in Chapter 4. Thus, in this section, the procedure to determine the

unknown values for the Casson model is discussed.

6.2.3.1. Curve Fitting

As discussed in Chapter 4, there are three unknown values, i.e., k , sth∆ , and

yh∆ , to be determined through the iterations using the same software package (Excel

Solver; see Appendix E) used in the power-law model case. The least-square method

was used for curve fitting of the experimental data and Eq. (4-41) to obtain the three

unknowns involved in the Casson model.

The procedure of data reduction for distilled water is shown in Fig. 6-11.

Figure 6-11(a) shows mean-velocity variations at a riser tube which were obtained

experimentally and theoretically. In the case of theoretical values, the initial guesses

for the three unknowns were used to estimate the values. In Fig. 6-11(b), the curve-

fitting results after iterations to minimize the sum of error that were calculated by Eq.

(4-42) are shown. The initial guesses and final values of k , sth∆ , and yh∆ for

distilled water are shown in Table 6-2.

Figures 6-12 and 6-13 show the curve-fitting procedures for human blood

obtained from two donors. As shown in Table 6-2, the same initial guesses of the

three unknowns were used for the two different bloods. However, the resulting

values of the three unknowns in the Casson model for the two donors were very

different, validating the present curve-fitting method.

Page 150: A Study of Non-Newtonian Viscosit y and Yield Stress of

135

6.2.3.2. Results and Discussion

Figures 6-14 and 6-15 show test results obtained with distilled water at 25.

Figure 6-14 shows the fluid-level variations in the two riser tubes, )(1 th and )(2 th .

Both fluid levels converged gradually from the initial difference to an equilibrium

state. Even for the distilled water, ∞=∆ th was not zero due to the difference in the

surface tension between the two riser tubes. Unless the wetting conditions of the

liquid-solid interface at the two riser tubes are exactly same, the surface-tension

difference always exists.

Figure 6-15 shows viscosity results from two tests for distilled water at 25

obtained with the SCTR, together with the reference data for comparison. The values

of sth∆ were determined to be approximately 4 mm for both tests whereas the values

of yh∆ were determined to be zero by a computer program (Microsoft Excel-solver),

validating the data reduction procedure involving the yield stress. Based on this

viscosity measurement method, the viscosity of the distilled water was found to be

between 0.876 and 0.878 cP at 25. The solid line indicates the viscosity of the

distilled water calculated by using the so-called Andrade’s equation. Since the water

viscosity data are given in the literature as a function of temperature, the exact

viscosity of water at 25 was calculated using the Andrade’s equation [Munson et al.,

1998]:

TB

eD ⋅=µ (6-1)

Page 151: A Study of Non-Newtonian Viscosit y and Yield Stress of

136

where D and B are given as 41093.9 −× mPa·s and 2026.57 K, respectively, for water

in a temperature range between 20 and 30.

Based on Eq. (6-1), the water viscosity was estimated to be 0.892 cP at 25.

The test results obtained with the SCTR gave less than 2% error in the entire shear

rate range, validating the test methods and data reduction procedure. Thus, it is

concluded that it is extremely important to consider the effect of the surface tension

on the viscosity measurement using a gravity driven, U-shaped capillary-tube system.

Figure 6-16 shows height variations at the two riser tubes as a function of time

for bovine blood with 7.5 % EDTA at a room temperature of 25. The trends of

fluid-level variations for the bovine blood were very similar to those for the distilled

water. As expected, ∞=∆ th for the bovine blood was not zero but a finite value that is

slightly greater than that for the distilled water. The height difference due to surface

tension, sth∆ , and the height difference due to yield stress, yh∆ , were determined to

be in the range of 5.7-6.1 mm and 0.52-0.59 mm, respectively. The hematocrit of the

bovine blood was measured to be 35 percent.

Figure 6-17 shows the viscosity of the bovine blood with 7.5% EDTA at 25,

which was measured with both the SCTR (indicated by circles and triangles) and the

rotating viscometer (RV; indicated by diamonds). Compared with the measured data

using the RV, the test results from the SCTR gave excellent agreement within 3% in a

shear rate range between 15 and 300 s-1. However, as the shear rate decreased below

15 s-1, the viscosity data measured from the RV seemed to be incorrect. More

specifically, the torque for viscosity measurements with the RV should be greater

than 10% of the full scale (as suggested by Brookfield) at a given shear rate for

Page 152: A Study of Non-Newtonian Viscosit y and Yield Stress of

137

reasonably accurate viscosity measurements. In the case of the bovine blood, the

minimum shear rate that Brookfield recommended, based on the 10% criterion, was

approximately 30 s-1. In contrast, the SCTR gave a consistent viscosity measurement

over a range of shear rates as low as a shear rate of 1 s-1. Related to the uncertainty in

measuring blood viscosity at low shear rates, wall slip is an issue. Since the

minimum shear rate in this study was 1 s-1, as discussed earlier, one could assume that

the slip effect was negligibly small [Picart et al., 1999a].

Figures 6-18 and 6-19 show test results of fresh, unadulterated human blood at

a body temperature of 37. The test was completed within 2-3 min to avoid blood

clotting, which might have altered the viscosity of the blood. Figure 6-18 shows

height variations in the two riser tubes as a function of time for the fresh human blood

at 37. The value of ∞=∆ th for the fresh human blood had a finite value, which

depended on individual donors. The values of sth∆ for donors 1 and 2 were found to

be approximately 8.5 and 9.6 mm, respectively.

In the use of the SCTR, two phenomena of particular importance should be

pointed out in regards to clinical hemorheology: one is the carry-over effect and the

other is the surface tension effect. First, at the completion of a measurement, a thin

layer of the test blood sample was always retained on the tube wall, unless the tube

was very carefully washed and dried. This residual layer can be called carry-over.

Hence, for unadulterated blood-viscosity measurements, it was necessary to use

disposable capillary-riser tube sets to avoid the carry-over phenomenon. Second, the

difference between surface tensions at the two riser tubes may vary from one

Page 153: A Study of Non-Newtonian Viscosit y and Yield Stress of

138

disposable set to another. Although the riser tubes were made of the same material

(acrylic plastic), the surface tensions were slightly different from set to set.

The values of yh∆ representing the effect of the yield stress for donors 1 and 2

were found to be 0.7 and 0.2 mm, respectively. These results were consistent with

hematocrit data for donors 1 and 2, which were 42 and 35 percent, respectively.

Donor 2 (a physician) practices therapeutic bloodletting periodically, which explains

why the hematocrit of donor 2 was unusually low. It is of note that the value of yh∆

represents the thixotropic characteristics of fresh human blood that is closely related

to the yield stress.

A thixotropic blood exhibits a high viscosity when first sheared from rest.

The viscosity continues to decrease as shearing continues. Thixotropy is usually a

result of the partial destruction, by shearing, of the internal liquid structure. While at

rest, the internal structure made of suspended cells and plasma may form to create

aggregations of RBCs, for example. This phenomenon is generally referred to as

‘structure viscosity’.

In fact, the phenomenon of RBCs aggregations at low shear rates is well

known but not well understood so far. The forces leading to aggregations are weak,

so if a sample of normal blood is subjected to increasing shear rate, the aggregates

progressively break up and are generally monodispersed at a shear rate greater than

10 s-1. However, it is important to note that there could be two kinds of yield stresses:

a start-up yield stress and a stopping yield stress [Cho and Choi, 1993]. The yield

stress determined in this study was the stopping yield stress, an important

phenomenon in clinical hemorheology and studies of cardiovascular disease. In the

Page 154: A Study of Non-Newtonian Viscosit y and Yield Stress of

139

viscosity measurement with whole blood, a typical test duration in which the shear

rate decreased from 10 to 1 s-1 was longer than 60 seconds. It is reasonable to assume

that the 60-second period is long enough to cause aggregations if the aggregations

were going to take place.

Figure 6-19 shows the viscosities of the two donors at 37 measured with the

SCTR. The viscosities for donor 2 were significantly lower than those for donor 1

due to the difference in hematocrit. In addition, the viscosity curve for donor 2 was

flatter than that for donor 1 since donor 2 had a smaller value of yh∆ and a lower

hematocrit. Figure 6-20 shows shear stress variations against shear rate for both

donors. Like the case of viscosity, shear stress for donor 2 was significantly lower

than that for donor 1 due to differences in hematocrit and yield stress. The values of

the yield stress determined from Casson model were approximately 14 mPa and 5

mPa for donor 1 and 2, respectively. The difference between the viscosity data

increased as the shear rate decreased indicating that the viscosity was more influenced

by hemorheological parameters such as hematocrit and yield stress at low shear rates

than at high shear rates.

6.2.4. Data Reduction with Herschel-Bulkley (H-B) Model

The detailed mathematical procedure for curve-fitting using a Herschel-

Bulkley (H-B) model was provided in Chapter 4. As in the case of the Casson model,

the H-B model can also handle the yield stress of blood. However, in the data

Page 155: A Study of Non-Newtonian Viscosit y and Yield Stress of

140

reduction with the H-B model, there are four unknown values to be determined

through the curve-fitting technique, whereas the Casson model has only three

unknowns.

The four unknown values, i.e., m , n , sth∆ , and yh∆ , are to be determined

through a least-square method using the same software package (Excel-Solver,

Microsoft; see Appendix E) that uses the formula of Newton’s method. In the

process of curve fitting, Eq. (4-59) was used for theoretical values. In the case of the

H-B model, the experimental data of bovine blood with 7.5% EDTA were used to

validate the method of data reduction.

Figure 6-21(a) shows mean-velocity values at a riser tube which were

obtained both experimentally and theoretically. In the case of theoretical values, the

initial guesses for the four unknowns, which are shown in Table 6-3, were used. In

Fig. 6-21(b), the curve-fitting results after the iterations to determine the four

unknowns are shown. The initial guesses and final values of m , n , sth∆ , and yh∆

for bovine blood with 7.5% EDTA using the H-B model are shown in Table 6-3.

Figure 6-22 shows the viscosity of the bovine blood with 7.5% EDTA at 25,

which was measured with the SCTR. Three consecutive tests were performed with

bovine blood. As expected, the H-B model also produced very accurate and

repeatable results. The final values of four unknowns for each test are reported in

Table 6-4. The effects of constitutive models on the viscosity and yield stress

measurement of blood will be further discussed in the next section.

Page 156: A Study of Non-Newtonian Viscosit y and Yield Stress of

141

Fig. 6-9. Picture of a SCTR with plastic riser tubes.

CCD-LED arrays

Riser tube 1

Capillary tube Stopcock Test Fluid

Riser tube 2

Page 157: A Study of Non-Newtonian Viscosit y and Yield Stress of

142

Fig. 6-10. Heating pad for a test with unadulterated human blood.

Rheometer System

Heating Pad

Page 158: A Study of Non-Newtonian Viscosit y and Yield Stress of

143

(a) With initial guess values

(b) With final resulting values

Fig. 6-11. Curve-fitting procedure with Casson model for distilled water.

0

0.005

0.01

0.015

0.02

0.025

0 10 20 30 40

Time (s)

Experimental dataTheoretical data

rV2 (m/s)

0

0.005

0.01

0.015

0.02

0.025

0 10 20 30 40

Time (s)

Experimental dataTheoretical data

rV2 (m/s)

Page 159: A Study of Non-Newtonian Viscosit y and Yield Stress of

144

Table. 6-2. Comparison of initial guess and resulting value using Casson model.

Distilled Water Human Blood

Initial Guess

k = 1 (cP·s) yh∆ = 0.5 mm

sth∆ = 3 mm

k = 1 (cP·s) yh∆ = 0.5 mm

sth∆ = 5 mm

Resulting Value

k = 0.878 (cP·s) yh∆ = 0.00

sth∆ = 4.02 mm

Donor 1

k = 2.743 (cP·s) yh∆ = 0.74 mm

sth∆ = 8.47 mm

Donor 2

k = 2.121 (cP·s) yh∆ = 0.16 mm

sth∆ = 9.58 mm

Page 160: A Study of Non-Newtonian Viscosit y and Yield Stress of

145

(a) With initial guess values

(b) With final resulting values

Fig. 6-12. Curve-fitting procedure with Casson model for donor 1.

0

0.002

0.004

0.006

0.008

0 20 40 60 80 100 120

Time (s)

Experimental dataTheoretical data

rV2 (m/s)

0

0.002

0.004

0.006

0.008

0 20 40 60 80 100 120

Time (s)

Experimental dataTheoretical data

rV2 (m/s)

Page 161: A Study of Non-Newtonian Viscosit y and Yield Stress of

146

(a) With initial guess values

(b) With final resulting values

Fig. 6-13. Curve-fitting procedure with Casson model for donor 2.

0

0.002

0.004

0.006

0.008

0 20 40 60 80 100 120

Time (s)

Experimental dataTheoretical data

rV2 (m/s)

0

0.002

0.004

0.006

0.008

0 20 40 60 80 100 120

Time (s)

Experimental dataTheoretical data

rV2 (m/s)

Page 162: A Study of Non-Newtonian Viscosit y and Yield Stress of

147

Fig. 6-14. Height variation in each riser tube vs. time for distilled water at 25.

0

0.02

0.04

0.06

0.08

0.1

0 50 100 150Time (s)

Hei

ght (

m) Riser tube 1

Riser tube 2

Page 163: A Study of Non-Newtonian Viscosit y and Yield Stress of

148

Fig. 6-15. Viscosity measurement for distilled water at 25. 1 cP = 1 mPa·s.

0.5

0.7

0.9

1.1

1.3

1.5

0 100 200 300 400Shear rate (s-1)

Vis

cosi

ty (c

P)

Test 1Test 2Reference

Page 164: A Study of Non-Newtonian Viscosit y and Yield Stress of

149

Fig. 6-16. Height variation in each riser tube vs. time for bovine blood with 7.5% EDTA at 25.

0

0.02

0.04

0.06

0.08

0.1

0 30 60 90 120 150Time (s)

Hei

ght (

m)

Riser tube 1

Riser tube 2

Page 165: A Study of Non-Newtonian Viscosit y and Yield Stress of

150

Fig. 6-17. Viscosity measurement for bovine blood with 7.5% EDTA at 25 using both rotating viscometer (RV) and scanning capillary-tube rheometer (SCTR). Hematocrit was 35. 1 cP = 1 mPa·s.

1

10

100

1 10 100 1000Shear rate (s-1)

Vis

cosi

ty (c

P)

Test 1Test 2RV

Page 166: A Study of Non-Newtonian Viscosit y and Yield Stress of

151

Fig. 6-18. Height variation in each riser tube vs. time for human blood at 37.

00.010.020.030.040.050.060.070.080.09

0.1

0 30 60 90 120 150Time (s)

Hei

ght (

m)

B

A

A : Donor 1B : Donor 2

Page 167: A Study of Non-Newtonian Viscosit y and Yield Stress of

152

Fig. 6-19. Viscosity measurement for human blood (2 different donors) at 37. Hematocrits for donors 1 and 2 were 42 and 35, respectively. 1 cP = 1 mPa·s.

1

10

100

1 10 100 1000Shear rate (s-1)

Visc

osity

(cP

)

Donor 1Donor 2

Page 168: A Study of Non-Newtonian Viscosit y and Yield Stress of

153

Fig. 6-20. Shear-stress variation vs. shear rate for human blood (from 2 different donors) at 37.

1

10

100

1000

10000

1 10 100 1000

Shear rate (s-1)

She

ar s

tress

(mP

a)

Donor 1Donor 2

Page 169: A Study of Non-Newtonian Viscosit y and Yield Stress of

154

(a) With initial guess values

(b) With final resulting values

Fig. 6-21. Curve-fitting procedure with Herschel-Bulkley model for bovine blood.

0

0.001

0.002

0.003

0.004

0.005

0 40 80 120

Time (s)

Experimental dataTheoretical data

rV2 (m/s)

0

0.001

0.002

0.003

0.004

0.005

0 40 80 120

Time (s)

Experimental dataTheoretical datarV2

(m/s)

Page 170: A Study of Non-Newtonian Viscosit y and Yield Stress of

155

Table. 6-3. Comparison of initial guess and resulting value using Herschel-Bulkley model.

Bovine Blood

Initial Guess

n = 0.8 m = 5 (cP·sn-1)

yh∆ = 1 mm

sth∆ = 6 mm

Resulting Value

n = 0.875 m = 8.6 (cP·sn-1)

yh∆ = 1.2 mm

sth∆ = 5.8 mm

Page 171: A Study of Non-Newtonian Viscosit y and Yield Stress of

156

Fig. 6-22. Viscosity measurements of the bovine blood with 7.5% EDTA at 25, which were analyzed with Herschel-Bulkley model.

1

10

100

1 10 100 1000

Shear rate (s-1)

Vis

cosi

ty (c

P)

Test #1Test #2Test #3

Page 172: A Study of Non-Newtonian Viscosit y and Yield Stress of

157

Table. 6-4. Comparison of four unknowns determined with Herschel-Bulkley model for three consecutive tests.

Test #1 Test #2 Test #3

n = 0.875 m = 8.6 (cP·sn-1)

yh∆ = 1.2 mm

sth∆ = 5.8 mm

n = 0.876 m = 8.7 (cP·sn-1)

yh∆ = 1.2 mm

sth∆ = 5.8 mm

n = 0.872 m = 8.76 (cP·sn-1)

yh∆ = 1.24 mm

sth∆ = 5.8 mm

Page 173: A Study of Non-Newtonian Viscosit y and Yield Stress of

158

6.3. Comparison of Non-Newtonian Constitutive Models

It is well known that blood has both shear-thinning (pseudoplastic)

characteristics and yield stress. The present study examined the capability of three

constitutive models in handling data from the scanning capillary-tube rheometer

(SCTR) for the viscosity and yield-stress measurements of blood. Power-law model

was chosen for the shear-thinning behavior of the blood. Casson and Herschel-

Bulkley (H-B) models were selected to measure both the shear-thinning viscosity and

yield stress of the blood.

As shown in Table 6-5, many researchers used non-Newtonian constitutive

models for the investigations on blood rheology and flows which include power-law,

Casson, and H-B models. In addition, it has been pointed out that the selection of a

constitutive model could be very significant in analyzing blood flows [Tu and Deville,

1996; Siauw et al., 2000]. Siauw et al. (2000) performed a comparative study of non-

Newtonian models for the prediction of unsteady stenosis flows by using power-law

and Casson models, whereas Tu and Deville (1996) used H-B, Bingham, and power-

law models for their study on stenosis flows.

The objective of the present study was to investigate whether or not the results

of blood rheology and flow in the SCTR could be significantly altered by constitutive

models. Hence, the study investigated the effect of the three models on the viscosity

and yield-stress measurements of blood using the SCTR as well as on the flow

patterns of blood such as velocity profile and wall shear stress in a blood vessel.

Page 174: A Study of Non-Newtonian Viscosit y and Yield Stress of

159

6.3.1. Comparison of Viscosity Results

Figures 6-23(a) and 6-23(b) show the calibration results obtained in the SCTR

with distilled water at 25. Figure 6-23(a) shows the fluid-level variations in the

two riser tubes, )(1 th and )(2 th . Both fluid levels converged gradually from the

initial difference to an equilibrium state. Even for the distilled water, the difference

in the two fluid-levels at ∞=t , ∞=∆ th , was not zero due to the difference in the

surface tension between the two riser tubes.

Figure 6-23(b) shows viscosity results for distilled water at 25 obtained

with the SCTR using the three different non-Newtonian constitutive models, together

with the reference data for comparison. The values of sth∆ were determined to be

approximately 3.3-3.4 mm for the three models, whereas the values of yh∆ were

determined to be zero for all models, validating the data reduction procedure

involving the yield stress. The viscosity of distilled water was found to be between

0.884 and 0.905 cP at 25 for all three models. As shown in Table 6-6, the test

results obtained with the SCTR gave less than 2% error over the entire shear rate

range, validating the experimental procedure and data reduction method using the

three non-Newtonian constitutive models. Furthermore, the viscosity measurement of

water confirmed that one needs to consider the effect of the surface tension on the

viscosity measurement in the SCTR.

Figures 6-24(a) and 6-24(b) show test results obtained with bovine blood at

25. Figure 6-24(a) shows height variations at two riser tubes as a function of time

Page 175: A Study of Non-Newtonian Viscosit y and Yield Stress of

160

for the bovine blood with 7.5% EDTA as an anticoagulant. The trends of fluid-level

variations for the bovine blood were very similar to those for the distilled water. As

expected, ∞=∆ th for the bovine blood was not zero, but a finite value that is slightly

greater than that for the distilled water. The hematocrit of the bovine blood was

measured to be 35%.

Figure 6-24(b) shows the viscosity of the bovine blood at 25, which was

measured with both the SCTR and a rotating viscometer (RV). The SCTR results

obtained by using the three constitutive models showed very good agreement among

themselves at shear rates higher than approximately 30 s-1. However, the power-law

model maintained the rate of the viscosity change (the slope of a viscosity curve)

constant whereas both the Casson and H-B models increased the rate of viscosity

change as shear rate decreased, indicating the existence of yield stress in the bovine

blood. Moreover, the viscosity results using the Casson model showed very good

agreement with those using the H-B model.

The test results from the SCTR using the three non-Newtonian models gave

excellent agreement with the measured data using the RV within 5% in a shear-rate

range between 30 and 300 s-1. However, as the shear rate decreased below 30 s-1, the

viscosity data measured from the RV seemed to be incorrect. More specifically, the

torque for viscosity measurements with the RV should be greater than 10% of the full

scale (as suggested by Brookfield) for reasonably accurate viscosity measurements.

In the case of the bovine blood, the minimum shear rate that Brookfield

recommended, based on the 10% criterion, was approximately 30 s-1. In contrast, the

SCTR gave a consistent viscosity measurement over a range of shear rate as low as a

Page 176: A Study of Non-Newtonian Viscosit y and Yield Stress of

161

shear-rate of 1 s-1. In addition, the viscosities measured using both the Casson and H-

B models in the SCTR seemed to be more accurate than the power-law viscosity at

low shear rates. Table 6-7 provides the viscosity results at the several shear rates,

which were produced by using the three non-Newtonian constitutive models.

Figures 6-25(a) and 6-25(b) show test results of fresh, unadulterated human

blood at a body temperature of 37. The test was completed within 2-3 minutes to

avoid blood clotting, which might have altered the viscosity of the blood. Figure 6-

25(a) shows height variations in the two riser tubes as a function of time for the fresh

human blood at 37. The value of ∞=∆ th for the fresh human blood was measured to

be approximately 8.5 mm. The hematocrit of this donor’s blood was 42%.

Figure 6-25(b) shows the viscosity of the human blood at a body temperature

of 37, which was measured with the SCTR. The SCTR results obtained by using

the three constitutive models showed very good agreement among themselves at

shear rates higher than approximately 20 s-1. However, like the bovine blood case,

the power-law model maintained the rate of the viscosity change constant whereas

both the Casson and H-B models increased the rate of viscosity change as shear rate

decreased, indicating the existence of yield stress in the human blood. Table 6-8

shows the viscosity results of the unadulterated human blood for the three constitutive

models.

For cases of the bovine blood and fresh human blood, the power-law model

maintained constant slopes in the viscosity curve, while both the Casson and H-B

models showed rapid increases in the viscosity as the shear rate decreased. However,

Page 177: A Study of Non-Newtonian Viscosit y and Yield Stress of

162

the viscosities obtained from the three different constitutive models gave good

agreement at a high shear rate zone.

6.3.2. Comparison of Yield Stress Results

Over the years, many researchers have reported on the yield stress of blood.

The measurements of the rheological properties of fluids having yield stress are

summarized by Nguyen et al. (1983). According to their classifications, the yield-

stress measurement with the SCTR can be classified as indirect methods rather than

direct methods since the yield stress of a fluid can be obtained in the SCTR by using

constitutive models. Bingham plastic, Casson, and H-B models were used in their

study to describe the rheological behavior of yield-stress fluids.

Figures 6-26(a), 6-26(b), and 6-26(c) show the results of shear stress versus

shear rate for distilled water, bovine blood, and human blood, respectively. In the

case of the distilled water, the shear stresses obtained using the three different non-

Newtonian models were almost identical. The yield stress can be graphically

described as the intersecting point where the shear stress-shear rate curve meets with

the y-axis (i.e., at zero shear rate). Table 6-9 shows the yield-stress results together

with the model constants of the three constitutive models. As expected, the distilled

water shows no yield stress, while both the bovine and human bloods show finite

values of the yield stress for the cases of Casson and H-B models. Note that the yield

Page 178: A Study of Non-Newtonian Viscosit y and Yield Stress of

163

stress values for the Casson and H-B models were calculated using the following

equation:

c

cyy L

Rhg2

⋅∆=ρ

τ (6-2)

The yield stress of the human blood was consistently greater than that of the

bovine blood although the human blood was tested at a high temperature of 37. It

might be due to both the difference in hematocrit and the RBC aggregations of the

human blood at low shear rates. The yield stress values of the human blood with

hematocrit of 42% were measured to be 13.8 and 17.5 mPa for the Casson and H-B

models, respectively. Note that the yield stress measured in the present study was the

stopping yield stress, an important phenomenon in clinical hemorheology and

treatments of cardiovascular disease. The yield stress values vary from 1 mPa to 30

mPa for normal human blood with hematocrit of 40% [Chen et al., 1991; Picart et al.,

1999a], supporting the validity of the method of the present yield-stress measurement

using the SCTR.

The yield stress obtained with the H-B model was consistently greater than

that obtained with the Casson model in the cases of both bovine and human bloods as

shown in Table 6-9. In order to evaluate which model produces more accurate yield

stress results, the experimentally measured values of ∞=∆ th were compared with those

of yst hh ∆+∆ determined analytically through the curve-fitting procedure for the

bovine and human bloods (see Table 6-10). The value of the fluid-level difference in

riser tubes 1 and 2 at a time of 180 seconds was taken as a measure of ∞=∆ th . Hence,

Page 179: A Study of Non-Newtonian Viscosit y and Yield Stress of

164

the experimental values (i.e., ∞=∆ th ) should be bigger than those (i.e., yst hh ∆+∆ ) to

be obtained analytically.

As shown in Table 6-10, the values of yst hh ∆+∆ obtained with the Casson

model were consistently smaller than the experimentally measured values while the

values obtained with H-B model were bigger. Based on the comparison, one may

conclude that the Casson model does a better job in determining the yield stress of

blood than the H-B model. However, it is of note that both models produced almost

identical values of sth∆ , thus almost identical surface tensions, for the bovine and

human blood.

6.3.3. Effect of Yield Stress on Flow Patterns

Figure 6-27 shows the variations of )(tC y for the bovine blood with 7.5%

EDTA, indicating that the plug-flow region grows at the capillary tube with

increasing time. Due to the difference in yield stress values for the Casson and H-B

models, the size of the plug-flow region estimated from the two models start to differ

after approximately 30 seconds. Note that the H-B model predicts a much larger

plug-flow region than the Casson model.

Figures 6-28(a), 6-28(b), and 6-28(c) show velocity profiles at the capillary

tube for the bovine blood at room temperature of 25, which were plotted at three

mean velocities of 3, 0.3, and 0.03 cm/s, respectively. At a relatively high velocity of

3 cm/s, the three constitutive models predicted identical velocity profiles. However,

Page 180: A Study of Non-Newtonian Viscosit y and Yield Stress of

165

as the mean velocity (i.e., shear rate) decreased, the clear deviation among the three

models started to appear near the center of the tube at 0.3 cm/s and became bigger at

0.03 cm/s, a phenomenon which can be attributed to the difference in yield stress

values. Therefore, it can be concluded that the yield stress plays an important role in

the determination of both the blood viscosity and velocity profiles in a blood flow.

Figure 6-28(c) also shows that the size of the plug-flow region at the center of the

tube for the H-B model is much larger than that for the Casson model at

approximately 0.03 cm/s.

The shear rate, viscosity, and shear stress obtained with both the Casson and

H-B models for the bovine blood were plotted as a function of the mean velocity at

the capillary tube in Figs. 6-29(a), 6-29(b), and 6-29(c), respectively. Wall shear

stress represents the friction exerted on the vessel wall by moving blood. It has been

shown in a number of studies that wall shear stress may play an important role in

endothelial cell morphology and functions influencing the production of substances

such as nitric oxide, prostacyclin, and endothelin [Baldwin and Thurston, 1995;

Usami et al., 1995; Fung, 1996; Mitsumata et al., 1996; Samijo et al., 1998; Frame et

al., 1998; Cotran et al., 1999; Kensey and Cho, 2001].

As shown in Fig. 6-29(c), the value of the wall shear stress is almost

independent of the selection of a constitutive model. Due to the difference in the size

of the plug-flow regions, the wall shear rates for the Casson model at low mean

velocities were much lower than those for the H-B model, resulting in consistently

higher wall viscosity for the Casson model as shown in Fig. 6-29(b). Since the wall

shear stress can be calculated from the product of viscosity (shown in Fig. 6-29(b))

Page 181: A Study of Non-Newtonian Viscosit y and Yield Stress of

166

and wall shear rate (shown in Fig. 6-29(a)) at a given mean velocity, the difference in

the wall shear stress between the two models is very small.

Page 182: A Study of Non-Newtonian Viscosit y and Yield Stress of

167

Table. 6-5. Various physiological studies with non-Newtonian constitutive models.

Researchers Power-law Casson Herschel-Bulkley

Chakravarty and Datta (1992) x x

Siauw et al. (2000) x x

Tu and Deville (1996) x x

Liepsch and Moravec (1984) x

Walburn and Schneck (1976) x

Rohlf and Tenti (2001) x

Misra et al. (2000, 2002) x

Das and Batra (1995) x

Dash et al. (1996) x

Walawender et al. (1975) x

Rodkiewicz et al. (1990) x

Misra and Kar (1991) x

Chakravarty and Datta (1989, 1992)

x

Page 183: A Study of Non-Newtonian Viscosit y and Yield Stress of

168

(a)

(b)

Fig. 6-23. Test with distilled water at 25. (a) Fluid-level variations in two riser tubes. (b) Viscosity results.

0

20

40

60

80

100

0 30 60 90 120 150

Time (s)

Hei

ght (

mm

)

Riser tube 2

Riser tube1

0.5

0.7

0.9

1.1

1.3

1.5

0 100 200 300 400Shear rate (s-1)

Vis

cosi

ty (c

P)

power-lawH-BCassonReference

Page 184: A Study of Non-Newtonian Viscosit y and Yield Stress of

169

Table. 6-6. Measurements of water viscosity.

Model Viscosity (cP) *Error

Power-law 0.905 1.46%

Herschel-Bulkley (H-B) 0.884 0.90%

Casson 0.886 0.68%

*Error is based on the comparison with the reference value (0.892 cP at 25)

Page 185: A Study of Non-Newtonian Viscosit y and Yield Stress of

170

(a)

(b)

Fig. 6-24. Test with bovine blood at 25. (a) Height variations in riser tube vs. time for bovine blood with 7.5% EDTA. (b) Viscosity measurement for bovine blood with 7.5% EDTA using a rotating viscometer (RV) and a scanning capillary-tube rheometer (SCTR).

0

20

40

60

80

100

0 30 60 90 120 150

Time (s)

Hei

ght (

mm

) Riser tube 1

Riser tube 2

1

10

100

1 10 100 1000Shear rate (s-1)

Vis

cosi

ty (c

P)

power-lawH-BCassonRV

Page 186: A Study of Non-Newtonian Viscosit y and Yield Stress of

171

Table. 6-7. Measurements of bovine blood viscosity.

Viscosity (cP) from SCTR Shear rate

(s-1) Viscosity

(cP) from RV

Power-law Casson H-B

300 4.43 4.39 4.49 4.28

150 4.78 4.75 4.84 4.71

90 5.11 5.03 5.18 5.09

45 5.75 5.44 5.85 5.71

30 6.25 5.7 6.38 6.2

15 8.81 6.16 7.7 7.21

7.5 17 6.67 9.7 8.9

3 7.4 14.5 12.8

Lower than 3

8.38 (at 1 s-1) 22.5

(at 1.35 s-1) 18.55 (at 1.55 s-1)

Page 187: A Study of Non-Newtonian Viscosit y and Yield Stress of

172

(a)

(b)

Fig. 6-25. Test with unadulterated human blood at 37. (a) Height variations in riser tubes vs. time. (b) Viscosity results.

0

20

40

60

80

100

0 30 60 90 120 150

Time (s)

Hei

ght (

mm

) Riser tube 1

Riser tube 2

1

10

100

1 10 100 1000Shear rate (s-1)

Vis

cosi

ty (c

P)

power-lawH-BCasson

Page 188: A Study of Non-Newtonian Viscosit y and Yield Stress of

173

Table. 6-8. Measurements of human blood viscosity.

Viscosity (cP) from SCTR Shear rate (s-1)

Power-law Casson H-B

300 3.89 4.11 4.09

150 4.45 4.47 4.57

90 4.93 4.85 4.97

45 5.67 5.59 5.63

30 6.15 6.16 6.15

15 7.06 7.65 7.3

7.5 8.12 9.95 9.1

Lower than 5 9.76 (at 3 s-1) 14.73

(at 3.33 s-1) 12.9 (at 3.3 s-1)

Lower than 3 12.17 (at 1 s-1) 27.26

(at 1.18 s-1) 21.9 (at 1.32 s-1)

Page 189: A Study of Non-Newtonian Viscosit y and Yield Stress of

174

(a)

(b) (c)

Fig. 6-26. Wall shear stress at a capillary tube vs. shear rate. (a) for distilled water at 25. (b) for bovine blood at 25. (c) for human blood at 37.

1

10

100

1000

1 10 100

She

ar s

tress

(mP

a)

power-lawH-BCasson

Bovine Blood

1

10

100

1000

1 10 100Shear rate (s-1)

She

ar s

tress

(mP

a)

power-lawH-BCasson

Human Blood

1

10

100

1000

1 10 100

She

ar s

tress

(mP

a) power-law

H-BCasson

Water

Page 190: A Study of Non-Newtonian Viscosit y and Yield Stress of

175

Table. 6-9. Comparison of model constants, yh∆ , and yτ . (Note that =][m cP·sn-1)

Power-law H-B Casson

n = 1 m = 0.905

n = 1 m = 0.884 k = 0.886 cP

Distilled water (25) yh∆ = 0

yτ = 0 yh∆ = 0

yτ = 0 yh∆ = 0

yτ = 0 n = 0.8866

m = 8.3771 n = 0.8753 m = 8.599 k = 3.7302 cP

Bovine blood (25) yh∆ = 0

yτ = 0 yh∆ = 0.8 mm

yτ = 16.4 mPa yh∆ = 0.52 mm

yτ = 10.7 mPa n = 0.7991

m = 12.171n = 0.8601 m = 8.9721 k = 3.2896 cP

Human blood (37) yh∆ = 0

yτ = 0 yh∆ = 0.85 mm

yτ = 17.5 mPa yh∆ = 0.67 mm

yτ = 13.8 mPa

Page 191: A Study of Non-Newtonian Viscosit y and Yield Stress of

176

Table. 6-10. Comparison of ∞=∆ th and yst hh ∆+∆ .

H-B Casson

∞=∆ th

(experimental)

6.5 mm 6.5 mm

Bovine blood (25)

yst hh ∆+∆ (analytical)

6.6 mm ( sth∆ = 5.8 mm

yh∆ = 0.8 mm)

6.26 mm ( sth∆ = 5.74 mm

yh∆ = 0.52 mm)

∞=∆ th

(experimental)

9.13 mm 9.13 mm

Human blood (37)

yst hh ∆+∆ (analytical)

9.25 mm ( sth∆ = 8.4 mm

yh∆ = 0.85 mm)

9.07 mm ( sth∆ = 8.4 mm

yh∆ = 0.67 mm)

Page 192: A Study of Non-Newtonian Viscosit y and Yield Stress of

177

0

0.2

0.4

0.6

0.8

1

0 50 100 150 200Time (s)

Cy

CassonH-B

Fig. 6-27. Variations of a plug-flow region at a capillary tube as a function of time for bovine blood with 7.5% EDTA at 25.

Page 193: A Study of Non-Newtonian Viscosit y and Yield Stress of

178

(a)

(b)

(c)

Fig. 6-28. Velocity profiles at a capillary tube for the bovine blood with 7.5% EDTA at 25: (a) at a mean velocity of 3 cm/s. (b) approximately 0.3 cm/s. (c) approximately 0.03 cm/s.

0

0.02

0.04

0.06

0 0.01 0.02 0.03 0.04

Radius (cm)

Vel

ocity

(cm

/s)

power-lawCassonH-B

0.03 cm/s

0

0.2

0.4

0.6

0 0.01 0.02 0.03 0.04

Vel

ocity

(cm

/s)

power-lawCassonH-B

0.3 cm/s

0

2

4

6

0 0.01 0.02 0.03 0.04

Vel

ocity

(cm

/s)

power-lawCassonH-B

3 cm/s

Page 194: A Study of Non-Newtonian Viscosit y and Yield Stress of

179

(a)

(b)

(c) Fig. 6-29. (a) Wall shear rate, (b) Viscosity, and (c) Wall shear stress. Plotted as a function of mean velocity at a capillary tube using three non-Newtonian models for bovine blood with 7.5% EDTA. Note that 1 dyne/cm2 = 102 mPa.

0.1

1

10

100

1000

0.01 0.1 1 10

Wal

l she

ar ra

te (s

-1)

CassonH-B

1

10

100

1000

0.01 0.1 1 10

Vis

cosi

ty (c

P) Casson

H-B

0.01

0.1

1

10

100

0.01 0.1 1 10Mean velocity (cm/s)

Wal

l she

ar s

tress

(d

yne/

cm2 )

CassonH-B

Page 195: A Study of Non-Newtonian Viscosit y and Yield Stress of

180

CHAPTER 7. CONCLUSIONS AND RECOMMENDATIONS

The present study introduces a scanning capillary-tube rheometer to measure

liquid viscosity over a range of shear rates continuously from high to low shear rates

(as low as 1 s-1). The feasibility and accuracy of the new viscosity measurement

technique has been demonstrated for a standard-viscosity oil and unadulterated

human bloods by comparing the results obtained with a power-law model against an

established viscosity measurement technique. One of the advantages of this new

rheometer is that one can measure the viscosity of whole blood without using

anticoagulants. In addition, the viscosity measurement of whole blood can be

completed within 2 minutes in a clinical setting, rendering viscosity results over a

wide range of shear rates. The viscosity data from the new rheometer gave excellent

agreement with those measured within 1 minute by the rotating viscometer. The

rotating viscometer could not be used after 1 minute of use with an unadulterated

blood sample due to blood clotting.

The present study introduces a mathematical method to isolate the surface-

tension and yield-stress effects on the viscosity measurement in using a SCTR. The

feasibility and validity of the method to reduce data for the SCTR have been

demonstrated for distilled water and bovine blood by comparing with reference data

and the results from a Brookfield cone-and-plate rotating viscometer. The viscosity

of unadulterated human blood has also been measured using the SCTR.

Page 196: A Study of Non-Newtonian Viscosit y and Yield Stress of

181

The effect of the surface tension was taken into account by using an additional

term, sth∆ , while the effect of the yield stress was considered as a model constant in

either Casson or Herschel-Bulkley model. For the SCTR using gravity as a driving

force, it was necessary to consider the effect of the surface tension even for

Newtonian fluids. Furthermore, in the case of a thixotropic liquid like whole blood,

the surface-tension effect should be isolated from the yield-stress effect to obtain

accurate viscosity data over a range of shear rates using the SCTR. In order to avoid

the influence of the carry-over phenomenon on viscosity measurements, disposable

tube sets were used for tests with fresh human blood.

The present study also investigated the effect of dye concentration on the

viscosity of a dye-water solution using a SCTR. In the experiment, six different

concentrations (0.5, 1, 2, 3, 4, and 7% by volume) of dye were used at 25ºC. When

the dye concentration was greater than 2%, the viscosity of the dye-water solution

could be significantly altered particularly at low shear rates. Based on the experiment

with the SCTR, one can conclude that the maximum 2% concentration of dye by

volume can be used to make a transparent aqueous solution opaque for the operation

of the SCTR.

The present study investigated the effects of three non-Newtonian constitutive

models on the viscosity and yield stress measurements in a scanning capillary-tube

rheometer: power-law, Casson, and Herschel-Bulkley models.

For a Newtonian fluid (i.e., distilled water), all three models produced

excellent viscosity results. For non-Newtonian fluids (i.e., bovine and human

bloods), both Casson and H-B models gave viscosity results which are in good

Page 197: A Study of Non-Newtonian Viscosit y and Yield Stress of

182

agreement with each other as well as with the results obtained by a conventional

rotating viscometer, whereas the power-law model seemed to produce inaccurate

viscosities at low shear rates due to its inability to handle the yield stress of blood.

The yield stress values obtained from the Casson and H-B models for the

human blood were measured to be 13.8 and 17.5 mPa, respectively. The two models

showed some discrepancies in the yield-stress values. The results from the Casson

model seemed to be more accurate than those from the H-B model.

The ability to estimate the wall shear stress in various arterial vessels could be

a significant step in clinical hemorheology. In the present study, the wall shear stress

was found to be almost independent of a constitutive model, whereas the size of the

plug flow region varies substantially with the selected model, altering the values of

the wall shear rate at a given mean velocity. The model constants and the method of

the shear stress calculation given in the study can be useful in the diagnostics and

treatment of cardiovascular diseases.

Recommendations for Future Research

- The present study developed a new rheometer that was specially designed for

measuring unadulterated human blood. However, the measurement was not

strictly in vivo. It would be very useful to develop a method to measure the

viscosity of human blood in vivo.

Page 198: A Study of Non-Newtonian Viscosit y and Yield Stress of

183

- The present study focused on the method of measuring the viscosity of

unadulterated human blood. As discussed in Chapter 2, whole blood could be

affected by several factors such as RBC deformability and aggregation. The

effects of RBC deformability and aggregation on the blood viscosity should

be studied.

- The present study measured the viscosity and yield stress of human blood

without adding any anticoagulants. The study on the effect of thrombotic

tendency of each individual person on both viscosity and yield stress of blood

should be conducted.

- The two yield stress models, Casson and Herschel-Bulkley models, gave

different yield stresses for blood in the present study. It is not very clear

which model is more accurate. An experimental method of measuring

velocity profiles should be developed to determine the more accurate model

for characterizing blood sample.

Page 199: A Study of Non-Newtonian Viscosit y and Yield Stress of

184

LIST OF REFERENCES

Ajmani, R.S., Rifkind, J.M., Hemorheological changes during human aging, Gerontology, 44 (1998) 111-120. Baldwin, A.L., Thurston, G., Endothelial cell morphology in relation to blood flow and stretch in arterioles and veins, Biorheology, 32 (1995) 112. Barbee, J.H., The effect of temperature on the relative viscosity of human blood, Biorheology, 10 (1973) 1. Barbee, J.H., Cokelet, G.R., The Fahraeus effect, Microvasc. Res. 34 (1971) 6. Barnes, H.A., Thixotropy-a Review, J. Non-Newtonian Fluid Mech., 70 (1997) 1-33. Barnes, H.A., The yield stress-a review- everything flows?, J. Non-Newtonian Fluid Mech. 81 (1999) 133. Barnes, H.A., A review of the slip (wall depletion) of polymer solutions, emulsions and particle suspensions in viscometers: its cause, character, and cure, J. Non-Newtonian Fluid Mech., 56 (1995) 221-251. Barnes, H.A., Hutton, J.F., Walters, K., An Introduction to Rheology, Elsevier, Amsterdam, 1989. Barnes, H.A., Walters, K., The yield stress myth? Rheol. Acta, 24 (1985) 323-326. Benis, A.M., Usami, S., Chien, S., Effect of hematocrit and inertial losses on pressure-flow relations in the isolated hind paw of the dog, Circulation Res. 27 (1970) 1047. Bird, R.B., Armstrong, R.C., Hassager, O., Dynamics of Polymeric Liquids, Vol. 1, Wiley, New York, 1987. Bowdler, A.J., Foster, A.M., The effect of donor age on the flow properties of blood. Part I: Plasma and whole blood viscosity in adult males, Exp Gerontol 22 (1987) 155-64. Briley, D.P., Giraud, G.D., Beamer, N.B. et al., Spontaneous echo contrast and hemorheologic abnormalities in cerebrovascular disease, Stroke, 25 (1994) 1564.

Page 200: A Study of Non-Newtonian Viscosit y and Yield Stress of

185

Brookfield Engineering Laboratories, Inc., Rheology/Methodology, Seminar Reference, Manual No. SB 197-600-B398, 1999. Brown, A.M., A step-by-step guide to non-linear regression analysis of experimental data using a Microsoft Excel spreadsheet, Comp. Prog. Meth. Biomed., 65 (2001) 191-200. Casson, N., Rheology of Disperse System, Pergamon, London, 1959. Chakravarty, S., Datta, A., Dynamic Response of stenotic blood flow in vivo, Mathl. Comput. Modelling, 16 (1992) 3-20. Chakravarty, S., Datta, A., Effects of stenosis on arterial rheology through a mathematical model, Mathl. Comput. Modelling, 12 (1989) 1601-1612. Chakravarty, S., Datta, A., Dynamic Response of arterial blood flow in the presence of multi-stenoses, Mathl. Comput. Modelling, 13 (1990) 37-55. Chen, H.Q., Zhong, L.L., Wang, X.Y., Zhou, T., Chen, Z.Y., Effects of gender and age on thixotropic properties of whole blood from healthy adult subjects, Biorheology, 28 (1991) 177-183. Cheng, D.C.H. and Evans, F., Phenomenological characteristization of the rheological behavior of inelastic reversible thixotropic and antithixotropic fluids, British J Applied Physics, 16 (1965)1599-1617. Chien, S., Dormandy, J., E. Ernst, A. Matrai, Clinical Hemorheology, Martinus Nijhoff Publishers, Dordrecht, 1987. Cinar, Y., Demir, G., Pac, M., Cinar, A.B., Effect of hematocrit on blood pressure via hyperviscoesity, American Journal of Hypertension, 12 (1999) 739-743. Ciuffetti, G., Mercuri, M., Mannarino, E., et al., Peripheral Vascular disease: Rheologic variables during controlled ischemia, Circulation, 80 (1989) 348. Cho, Y.I., Choi, E., The rheology and hydrodynamic analysis of grease flows in a circular pipe, Tribology Trans. 36 (1993) 545. Cho, Y.I., Hartnett, J.P., Viscoelastic effects in the falling ball viscometer, J. Rheol. 24 (1980) 891. Cho, Y.I., Kensey, K.R., Effects of the Non-Newtonian viscosity of blood on hemodynamics of diseased arterial flows: Part 1, Steady flow, Biorheol. 28 (1991) 241.

Page 201: A Study of Non-Newtonian Viscosit y and Yield Stress of

186

Cho, Y.I., Kim, W.T., Kensey, K.R., A new scanning capillary tube viscometer, Rev. Sci. Instrum. 70 (1999) 2421. Cotran, R.S., Kumar, V., Collins, T., Robbins Pathologic Basis of Disease, 6th ed., WB Saunders, Philadelphia, 1999. Coull, B.M., Beamer, N., Garmo, P., et al., Chronic blood hyperviscosity in subjects with acute stroke, transient ischemic attack, and risk factors for stroke, Stroke, 22 (1991) 162. Craveri A., Tornaghi G., Paganardi L., Hemorrheologic disorders in obese patients: study of the viscosity of the blood, erythrocytes, plasma, fibrinogen and the erythrocyte filtration index. Minerva Med. 78 (1987) 899–906. Cross, M.M., Rheology of Non Newtonian Fluids: A New Flow Equation for Pseudoplastic Systems, J. Colloid Sci., 20 (1965) 417. Crouch, J.D., Keagy, B.A., Schwartz, J.A., Wilcox, J.R., Johnson, G., Current Surgery, 43 (1986) 395 Das, B., Batra, R.L., Non-Newtonian flow of blood in an arteriosclerotic blood vessel with rigid permeable walls, J. Theor. Biol. 175 (1995) 1-11. Dash, R.K., Jayaraman, G., Mehta, K.N., Estimation of increased flow resistance in a narrow catheterized artery-A theoretical model, J. Biomechanics, 29 (1996) 917-930. de Kee, D., Mohan, P., Soong, D.S., Yield stress determination of styrenebutadiene-styrene triblock copolymer solutions, J. Macromol. Sci., B25 (1986) 153. Dinnar U., Cardiovascular Fluid mechanics, CRC Press, Boca Raton, FL, 1981. Dintenfass, L., Blood microrheology-Viscosity factors in blood flow, ischaemia and thrombosis, Appleton-Century-Crofts, New York, 1971 Dintenfass, L., Blood viscosity factors in severe non-diabetic and diabetic retinopathy, Biorheology, 14 (1977) 151-157. Dintenfass, L., Modifications of blood rheology during aging and age-related pathological conditions, Aging, 1 (1989) 99-125. Einfeldt, J., Schmelzer, N., The theory of capillary viscometers taking into account surface tension effects, Rheol. Acta. 21 (1982) 95. Ernst, E.. Haemorheological consequences of chronic cigarette smoking. J Cardiovasc Risk 2 (1995) 435–439.

Page 202: A Study of Non-Newtonian Viscosit y and Yield Stress of

187

Ernst, E., Matrai, A., Marshall, M., Blood rheology in patients with transient ischemic attacks, Stroke, 19 (1988) 634. Fahraeus, R., Lingqvist, T., Viscosity of blood in narrow capillary tubes, Am. J. Physiol. 96 (1931) 562. Ferguson, J., Kemblowski, Z., Applied Fluid Rheology, Elsevier Science, London and New York, 1991. Fisher, M., Meiselman, H.J., Hemorheological factors in cerebral ischemia, Stroke, 22 (1991) 1164. Fowkes, F.G.R., Pell, J.P., Donnan, P.T., et al., Sex differences in susceptibility to etiologic factors for peripheral atherosclerosis: importance of plasma, fibrinogen, and blood viscosity, Arterioscler Thromb., 14 (1994) 862. Frame, M.D.S., Chapman, G.B., Makino, Y., Sarelius, I.H., Shear stress gradient over endothelial cells in a curved microchannel system, Biorheology, 35 (1998) 245. Fung, Y.C., Biomechanics: Motion, Flow, Stress, and Growth, Springer-Verlag, New York, 1990. Fung, Y.C., Biomechanics: Mechanical Properties of Living Tissues, Springer, New York, 1993. Fung Y.C., Biomechanics: Circulation, Springer, New York, 1996. Gaspar-Rosas, A., Thurston, G.B., Erythrocyte aggregate rheology by transmitted and reflected light, Biorheology, 25 (1988) 471. Goldsmith, H.L., The microrheology of human erythrocytes suspensions. In: Theoretical and Applied Mechanics, edited by Becker, E., Mikhailov, G.K., Proc. 13th IUTAM Congress, Springer, New York, 1972 Grotta, J., Ostrow, P., Fraifeld, E., Hartman, D., Gary, H., Fibrinogen, blood viscosity, and cerebral ischemia, Stroke, 16 (1985) 192. Guyton, A.C., Hall, J.E., Textbook of Medical Physiology, 9th edition, W.B. Sanders Company, Philadelphia, 1996. Harris, D., Nonlinear least-squares curve fitting with Microsoft excel Solver, J. Chem. Ed., 75 (1998) 119. Haynes, R.H., Physical basis of the dependence of blood viscosity on tube radius, Am. J. Physiol. 212 (1960) 1193.

Page 203: A Study of Non-Newtonian Viscosit y and Yield Stress of

188

Heywood, N.I., Cheng, D.C.H., Comparison of methods for predicting head loss in turbulent pipe flow of non-Newtonian fluids, Trans. Inst. Meas. Control, 6 (1984) 33. Herschel, W.H., Bulkley, R., Measurement of consistency as applied to rubber-benzene solution, Proc. Am. Soc. Test. Matls., 26 (1926) 621. Hildebrand, F.B., Introduction to Numerical Analysis, 2nd Ed., Dover, New York, 1987. Hill, M.A., Court J.M., Mitchell, G.M., Blood rheology and microalbuminuria in type 1 diabetes mellitus, Lancet 2 (1982) 985. Hoare, E.M., Barnes, A.J., Dormandy, J.A., Abnormal blood viscosity in diabetes and retinopathy, Biorheology, 13 (1976) 21-25. Holdsworth, S.D., Rheological models used for the prediction of the flow properties of food products: a Literature Review, Trans IChemE, 71 (1993) 139. How, T.V., Advanced in Hemodynamics and Hemorheology, Vol. 1, JAI Press, London, 1996. Huang, C.R., Siskovic, M., Robertson, R.W., Fabisiak, W., Smitherberg, E.J., Copley, A.L., Quantitative characterization of thixotropy of whole human blood, Biorheology, 12 (1975) 279. Isaacson, E., Keller, H.B., Analysis of Numerical Methods, Dover, New York, 1994. Jacobs, H.R., Meniscal resistance in tube viscometry, Biorheol. 3 (1966) 117. Jan, K., Chien, S., Bigger, J.T., Observations on blood viscosity changes after acute myocardial infarction, Circulation, 51 (1975) 1079. John, E.G., Simplified Curve Fitting using spreadsheet add-ins, Int. J. Engng. Ed., 14 (1998) 375-380. Kameneva, M.V., Antaki, J.F., Watachi, M.J., and Borovetz, H.S., Biorheology, 31 (1994) 297. Kameneva, M.V., Garrett, K.O., Watach, M.J., Borovetz, H.S., Red blood cell aging and risk of cardiovascular diseases, Clinical Hemorheology and Microcirculation, 18 (1998), 67-74. Kensey, K.R., Cho, Y.I., The Origin of Atherosclerosis, Vol. 1, EPP Medica, New Jersey, 2001.

Page 204: A Study of Non-Newtonian Viscosit y and Yield Stress of

189

Keentok, M., The measurement of the yield stress of liquids, Rhol. Acta, 21 (1982) 325-332. Kim, S., Cho, Y.I., The effect of dye concentration on the viscosity of water in a scanning capillary-tube viscometer, J. Non-Newtonian Fluid Mech. 2002 (submitted) Kim, S., Cho, Y.I., Hogenauer, W.N., Kensey, K.R., A method of isolating surface tension and yield stress effects in a U-shaped scanning capillary-tube viscometer using a Casson model, J. Non-Newtonian Fluid Mech. 103 (2002) 205-219. Kim, S., Cho, Y.I., Jeon, A.H., Hogenauer, B, Kensey, K.R., A new method for blood viscosity measurement, J. Non-Newtonian Fluid Mech. 1939 (2000a) 1-10. Kim, S., Cho, Y.I., Kensey, K.R., Pellizzari, R.O., Stark, P.R., A scanning dual-capillary-tube viscometer, Rev. Sci. Instrum. 71 (2000b) 3188. Lee A.J., Mowbray P.I., Lowe G.D., Blood viscosity and elevated carotid intima-media thickness in men and women: the Edinburgh Artery Study, Circulation 97 (1998) 1467–1473. Leiper, J.M., Lowe, G.D.O., Anderson, J. et al., Effects of diabetic control and biosynthetic human insulin on blood rheology in established diabetics, Diabetes res., 1 (1982) 27. Letcher, R.L., Chien, S., Pickering, T.G., et al., Direct relationship between blood pressure and blood viscosity in normal and hypertensive subjects: role of fibrinogen and concentration. Am J Med 70 (1981) 1198–1202. Letcher, R.L., Chien, S., Pickering, T.G., et al., Elevated blood viscosity in patients with borderline essential hypertension. Hypertension 5 (1983) 757–762. Levenson, J., Simon, A.C., Cambien, F.A., Beretti, C., Cigarette smoking and Hypertension, Arteriosclerosis, 7 (1987) 572. Liepsch, D., Moravec, S., Pulsatile flow of non-Newtonian fluids in distensible models of human arteries, Biorheology, 21 (1984) 571. Lowe, G.D.O., Blood rheology in arterial disease, Clinical Science 71 (1986) 137-146. Lowe, G.D.O., Drummond, M.M. Lorimer, A.R., et al., Relation between extent of coronary artery disease and blood viscosity, British Medical Journal, 8 (1980) 673. Lowe, G.D.O., Fowkes, F.G.R., Dawes, J., et al., Blood viscosity, fibrinogen, and activation of coagulation and leukocytes in peripheral arterial disease and the normal population in the Edinburgh artery study, Circulation, 87 (1993) 1915.

Page 205: A Study of Non-Newtonian Viscosit y and Yield Stress of

190

Macosko, C.W., Rheology; Principles, Measurements, and Applications, Wiley-VCH, New York, 1994. Magnin, A., Piau, J.M., Con-and-plate rheometry of yield stress fluids-Study of an aqueous gel, J. Non-Newtonian Fluid Mech., 36 (1990) 85-108. Mardles, E.W.J., The flow of liquids through fine capillaries and narrow channels: The meniscus resistance, Biorheol. 6 (1969) 1. Marshall, D.I., Riley, D.W., J. Appl. Polym. Sci. 6 (1962) 546. Metzger, A.P., Knox, J.P., Trans. Sci. Rheol., 9 (1965) 13. Mewis, J., Spaull, A.J.B., Rheology of concentrated dispersions, Advances in Cooloid and Interface Science, 6 (1976) 173-200. Microsoft Corporation, Microsoft Excel Users Guide, Version 5, Document Number XL 57926-0694. Middleman, S., The Flow of High Polymers, Interscience, New York, 1968. Misra, J.C., Ghosh, S.K., Flow of a Casson fluid in a narrow tube with a side branch, International Journal of Engineering Science, 38 (2000) 2045-2077. Misra, J.C., Kar, B.K., A mathematical analysis of blood flow from a feeding artery into a branch capillary, Mathl. Comput. Modelling, 15 (1991) 9-18. Misra, J.C., Pandey, S.K., Peristaltic transport of blood in small vessels: Study of a mathematical model, Computers and Mathematics with Applications 43 (2002) 1183-1193. Missirlis, K.A., Assimacopoulos, D., Mitsoulis, E., Chhabra, R.P., Wall effects for motion of spheres in power-law fluids, J. Non-Newtonian Fluid Mech., 96 (2001) 459-471. Mitsumata, M., Hagiwara, H., Yamane, T., Shu, K., Yoshida, Y., Morphology and function of endothelial cells under the blood flow, Biorheology, 33 (1996) 424. Most, A.S., Ruocco, N.A., Gewirtz, H., Effect of a reduction in blood viscosity on maximal myocardial oxygen delivery distal to a moderate coronary stenosis, Circulation, 74 (1986) 1085. Munson, B.R., Young, D.F., Okiishi, T.H., Fundamentals of Fluid Mechanics, Wiley, New York, 1998.

Page 206: A Study of Non-Newtonian Viscosit y and Yield Stress of

191

Nakamura, M., Sawada, T., Numerical study on the flow of a non-Newtonian fluid through an axisymmetric stenosis, ASME J. Biomech. Eng., 110 (1988) 137-143. Nguyen, Q.D., Boger, D.V., Yield stress measurement for concentrated suspensions, J. Rheol., 27 (1983) 321. Nguyen, Q.D., Boger, D.V., Measuring the flow properties of yield stress fluids, Annu. Rev. Fluid. Mech. 24 (1992) 47. Park, N.A., Irvine Jr., T.F., The falling needle viscometer. A new technique for viscosity measurements, Wärme-und Stoffübertragung 18 (1984) 201. Persson, S.U., Gustavsson, C.G., Larsson, H., Persson, S.. Studies on blood rheology in patients with primary pulmonary hypertension. Angiology, 42 (1991) 836–842. Picart, C., Piau, J.M., Galliard, H., Human blood shear yield stress and its hematocrit dependence, J. Rheol. 1 (1998a) 42. Picart, C., Piau, J.M., Galliard, H., Carpentier, P., Blood low shear rate rheometry: influence of fibrinogen level and hematocrit on slip and migrational effects, Biorheology, 35 (1998b) 335-353. Poon, P.Y.W., Dornan, T.L., Orde-Peckar, C. et al., Blood viscosity, glycaemic control and retinopathy in insulin dependent diabetes, Clin. Sci., 63 (1982) 211. Pries, A.R., Neuhaus, D. Gaehtgens, P., Blood viscosity in tube flow: dependence on diameter and hematocrit, American J. Physiology, 263 (1992) H1770-H1778. Reinhart, W.H., Haeberli, A., Stark, J., and Straub, P.W., J. Lab. Clinical Med., 115 (1990) 98 Resch K.L., Ernst E., Matrai A.,.Can rheologic variables be of prognostic relevance in arteriosclerotic diseases? Angiology 42 (1991) 963–970. Rosenblum, W.I., Blood, 31 (1968) 234 Rosenson, R.S., Viscosity and ischemic heart disease, J. Vascular Medicine and Biology, 4 (1993) 206. Rodkiewicz, C.M., Sinha, P., Kennedy, J.S., On the application of a constitutive equation for whole human blood, Transactions of the ASME, 112 (1990) 198. Rohlf, K., Tenti, G., The role of the Womersley number in pulsatile blood flow a theoretical study of the Casson model, J. Biomechanics, 34 (2001) 141-148.

Page 207: A Study of Non-Newtonian Viscosit y and Yield Stress of

192

Sajjadi, S.G., Nash, G.B., Rampling, M.W., Cardiovascualr flow modeling and measurement with application to clinical medicine, Clarendon Press, Oxford, 1999. Samijo, S.K., Willigers, J.M., Barkhuysen, R., Kitslaar, P.J.E.H.M., Reneman, R.S., Brands, P.J., Wall shear stress in the human common carotid artery as function of age and gender, Cardiovascular Research, 39 (1998) 515. Schmid- nbeinoSch && H., Wells, R.E., Rheological properties of human erythrocytes and their influence upon anormalous viscosity of blood, Physioly Rev, 63 (1971) 147-219. Sharp, D.S., Curb, J.D., Schatz, IJ., et al., Mean red cell volume as a correlate of blood pressure. Circulation 93 (1996) 1677–1684. Siauw, W.L., Ng, E.Y.K., Mazumdar, J., Unsteady stenosis flow prediction: a comparative study of non-Newtonian models with operator splitting scheme, Medical Engineering & Physics, 22 (2000) 265-277. Skinner, S.J., J. Appl. Polym. Sci. 5 (1961) 55. Smith, W.C.S., Lowe, G.D.O., Lee, A.J., Tunstall-Pedoe, H., Rheological determinants of blood pressure in a Scottish adult population, J. Hypertension, 10 (1992) 467. Stadler, A.A., Zilow, E.P. Linderkamp, O., Blood viscosity and optimal hematocrit in narrow tubes, Biorheology, 27 (1990) 779-788. Stoltz, J.F., Singh, M., Riha, P., Hemorheology in Practice. IOS Press, 1999. Tanner, R.I., Engineering rheology, Oxford University Press, New York, 1985. Thomas, H.W., The wall effect in capillary instruments: An improved analysis suitable for application to blood and other particulate suspensions, Biorheology, 1 (1962) 41-56. Thurston, G.B., Effect of hematocrit on blood viscoelasticity and in establishing normal values, Biorheology, 15 (1978) 239-249. Toth, K., Kesmarky, G., Vekasi, J., et al., Hemorheological and hemodynamic parameters in patients with essential hypertension and their modification by alpha-1 inhibitor drug treatment. Clin Hemorheol Microcirc 21 (1999) 209–216. Tsuda, Y., Satoh, K., Kitadai, M., et al., Chronic hemorheological effects of the calcium antagonist nilvadipine in essential hypertension. Arzneimittelforschung 47 (1997) 900–904.

Page 208: A Study of Non-Newtonian Viscosit y and Yield Stress of

193

Tu, C., Deville, M., Pulsatile flow of non-Newtonian fluids through arterial stenoses, J. Biomechanics, 29 (1996) 899-908. Uhlherr, P.H.T., A novel method for measuring yield stress in static fluids, In Proc. 4th Natl. Conf. Rheol., Aust. Sco. Rheol., (1986) 231. Usami, S., Wung, S.L., Chien, S., The response of endothelial cell to shear flow, Biorheology, 32 (1995) 231. Walawender, W.P., Chen, T.Y., Cala, D.F., An approximate Casson fluid model for tube flow of blood, Biorheol. 12 (1975) 111. Walburn, F.J., Schneck, D.J., A constitutive equation for whole human blood, Biorheology, 13 (1976) 201. Wildemuth, C.R., Williams, M.C., A new interpretation of viscosity and yield stress in dense slurries: coal and other irregular particles, Rheol. Acta 24 (1985) 75-91. Yarnell, J.W.G., Sweetnam, P.M., Rumley, A., Lowe, G.D.O., Lifestyle and hemostatic risk factors for ischemic heart disease: the Caerphilly Study. Arterioscler Thromb Vasc Biol 20 (2000) 271–279. Young, D.M., Gregory, R.T., A Survey of Numerical Mathematics, Vol. 1, Dover, New York, 1988.

Page 209: A Study of Non-Newtonian Viscosit y and Yield Stress of

194

APPENDIX A. Nomenclature

English Letters

C dimensionless radius at yττ = , Rr

yC dimensionless radius at yττ = , Rry

g gravitational acceleration [m·s-2]

H height [m]

h fluid height [m]

h∆ fluid height difference [m]

k constant for Casson model [cP]

LK loss coefficient

L length of tube [m]

l length of fluid element [m]

m consistency index in Power-law and Herschel-Bulkley models [cP·sn-1] M torque [N·m]

n power-law index

P pressure [Pa]

P∆ pressure drop [Pa]

R radius [m]

Page 210: A Study of Non-Newtonian Viscosit y and Yield Stress of

195

R mean radius [m]

r radial distance [m]

yr radial distance at yττ = [m]

Re Reynolds number

Q volumetric flow rate [m3·s-1]

s distance measured along stream line [m]

t time [s]

V flow velocity [m·s-1]

V mean flow velocity [m·s-1]

∞v terminal velocity [m·s-1]

Greek Letters

τ shear stress [Pa] or [cP·s-1]

γ& shear rate [s-1]

µ Newtonian viscosity [cP]

η non-Newtonian viscosity [cP]

0η constant viscosity near zero shear rate [cP]

∞η constant viscosity near infinite shear rate [cP]

ρ density [kg·m-3]

Ω angular velocity [rad·s-1]

Page 211: A Study of Non-Newtonian Viscosit y and Yield Stress of

196

β angle [rad]

Subscripts

atm atmosphere

B Bingham plastic model

c capillary tube

e entrance

End end effects

f fluid

i inner cylinder

L loss

n needle

o outer cylinder

r reservoir or riser tube

st surface tension

t time

unsteady unsteady state

w wall

y yield stress

∞ infinity

Page 212: A Study of Non-Newtonian Viscosit y and Yield Stress of

197

APPENDIX B. Falling Object Viscometer – Literature Review

Falling Cylinder/Needle

Figure B-1 shows a schematic diagram of a falling cylinder moving in a fluid

filled in another cylinder. In case that the gap between two cylinders is very small,

then simple shear can be obtained as indicated in Fig. B-1(a). As the cylinder is

dropped in a closed cylinder, the displaced fluid must flow back, which results in the

velocity profile shown in Fig. B-1(b). Typically, a small diameter needle is dropped

in a large cylinder of the test fluid. After the needle falls for a distance great enough

to allow the fluid to reach a steady state, the terminal velocity, ∞v , is determined by

timing between two marks. This generally limits the technique to transparent fluids.

Assuming a wide gap, i.e., κ << 1, the relations for shear stress, τ , and the

Newtonian viscosity, µ , are as follows [Macosko, 1994]:

κρρ

τ2

)( gRfn −= (B-1)

)ln1(2

)( 2

κρρ

µ +−

=∞v

gRfn (B-2)

where

nρ = density of needle

fρ = density of fluid

g = gravitational acceleration

Page 213: A Study of Non-Newtonian Viscosit y and Yield Stress of

198

R = inner radius of outer cylinder

∞v = terminal velocity of needle

Park and Irvine (1988) gave relations for a power-law fluid. They also demonstrate

that one can easily change the density of the needle and thus the shear stress, τ , by

using a hollow tube filled with various amounts of dead weight. In that way, one can

obtain non-Newtonian viscosity, η , as a function of shear rate, γ& .

Page 214: A Study of Non-Newtonian Viscosit y and Yield Stress of

199

Fig. B-1. Falling cylinder viscometers. (a) open ends for high viscous samples

(b) closed end, free falling

n

Page 215: A Study of Non-Newtonian Viscosit y and Yield Stress of

200

APPENDIX C. Specification of CCD and LED Array

Description

Syscan’s SV352A8-01 Contact Image Sensor (CIS) is a black/white linear

image sensor module, which is originally designed for scanning a document. Figure

C-1 illustrates the cross sectional view of the SV352A8-01 module. The module

consists of a LED light source to illuminate the document, a one-to-one erect graded

index micro lens array to focus the document image on the photo detector array, an

array of linear MOS image sensors to convert the image into an electronic signal, a

glass cover to protect the sensor array, micro lens array, and LED light source from

dust, 8-pin connector for input/output connections and a protective case to house all

of the components.

Key Features

• Compact size: 12 mm height × 15 mm width × 70 mm length

• Resolution: 12 dots/mm (304dpi)

• Scanning length: 2.2 inch or 3.2 inch

• Scanning speed: 2.5 ms/line

•Single power supply (+5V)

Page 216: A Study of Non-Newtonian Viscosit y and Yield Stress of

201

Fig. C-1. Cross sectional view of SV352A8-01 module.

Sensor Array PCB Substrate

Cover Glass

Rod Lens LED Light Source

Connector

Plastic Housing

Page 217: A Study of Non-Newtonian Viscosit y and Yield Stress of

202

APPENDIX D. Biocoating of Capillary Tube

For the experiment with unadulterated human blood, capillary tubes used in a

SCTR have been coated with biocompatible materials. The coating work was carried

out by a company called Biocompatibles in Farnham, Surrey (U.K.). The following

procedure was employed to coat the inner surface of the capillary tube:

1. Prepare 100 ml of a 10mg/ml PC1036 solution in 99% Hexane and 1%

Ethanol.

2. Clean the capillary tube lumen by using a 20ml syringe to pull and push the

Hexane through the lumen vertically.

3. Pass compressed air vertically through both ends of the capillary tube lumen

for 2 seconds at a flow rate of 30-35 liters per minute to remove any

remaining traces of Hexane.

4. Coat the capillary tube by using a 20 ml syringe to pull and push the PC1036

polymer solution through the lumen vertically.

5. Immediately after coating, pass compressed air vertically through both ends of

the capillary tube lumen for 2 seconds at a flow rate of 30-35 liters per minute

to remove any remaining traces of PC1036 polymer solution.

6. Place the capillary tube horizontally in an incubator preheated to 70 for 4

hours to allow the coating to cure.

Page 218: A Study of Non-Newtonian Viscosit y and Yield Stress of

203

The capillary tube lumen was reported to be free from blockage after the coating

procedure. In addition, the thickness of the polymer coating cured on the inner

surface of the capillary tube was reported approximately 1 µm.

Page 219: A Study of Non-Newtonian Viscosit y and Yield Stress of

204

APPENDIX E. Microsoft Excel Solver

A powerful tool that is widely available in a spreadsheet format provides a

simple means of fitting experimental data to both linear and non-linear functions

[Microsoft Corporation, 57926-0694]. The curve-fitting method by using Excel

Solver is well-known and widely used in scientific researches [Harris, 1998; John,

1998; Brown, 2001]. The procedure and its mode of operation are very easy and

obvious. Frequently in engineering, science and business, data is collected and

plotted as a graphical representation of the variables involved. The next step is to

create an association between the variables by connecting the points with a line.

Once drawn, the line is examined and a model which best fits the data points assumed

when the theoretical solution for the data points is not available. Then, this is fitted

and used to replace the existing set of data points as the appropriate model. However,

in case that the theoretical solution is available, this procedure can also be used to

determine the unknowns in the solution.

In order to fit a curve to a data series, using the excel solver is simplicity itself.

If a data series contains the x and y values, and an appropriate model has been

available. Fitting the chosen model is then as follows [Harris, 1998; John, 1998;

Brown, 2001]:

1. Enter the known x and y values as a data series onto the spreadsheet.

Page 220: A Study of Non-Newtonian Viscosit y and Yield Stress of

205

2. Add a further column containing a chosen model. The parameters

(unknowns) of the chosen model are estimated and located in any free cells.

These are the ‘Change cells’.

3. Add a further column which expresses the squared error between the known

y values and the assumed model values.

4. Sum the squared error column in an appropriate free cell.

5. Evoke Solver by selecting the Tolls menu and Solver to present the Solver

dialogue box.

6. In the dialogue box, enter the sum of the squared error cell as the target cell.

7. Set the Equal to option to Min.

8. Enter the selected ‘Change cells’ to the ‘By changing cells’ box.

9. Include any constraints and modify the options as necessary.

10. Select the Solver button to initiate the curve fitting.

The values of the assumed model parameters (unknowns) will then be adjusted in

each of the ‘Change cells’ until the Target cell value is a minimum. Excel Solver

uses Newton’s method of iteration to determine the best combination of unknowns

that fit into the model [Microsoft Corporation, 57926-0694].

Page 221: A Study of Non-Newtonian Viscosit y and Yield Stress of

206

APPENDIX F. Newton’s Method of Iteration

The Newton method is one of the most widely used methods for root finding.

It can be used for the problem to find solutions of a system of non-linear equations

[Young and Gregory, 1988; Isaacson and Keller, 1994; Hildebrand, 1987]. Consider

the general problem of fitting a function of the following type:

);( AXfy = (F-1)

where X (Variables) ),...,,( 21 nxxx= and A (Parameters) ),...,,( 21 maaa= .

Choosing the parameters, A ),...,,( 21 maaa= , which minimize the sum of

error, the least-square error function, )(AE , can be expressed by using the following

equation:

[ ]∑=

−=l

j

jj yAXfAE1

2)()( );()( (F-2)

A necessary condition that the parameters corresponding to a minimum is that they

are a stationary point. Therefore, the following system of equations should be

satisfied:

01

=∂∂aE , 0

2

=∂∂aE , …, and 0=

∂∂

maE . (F-3)

Note that some or all of the equations in Eq. (F-3) may be non-linear. Applying the

chain rule to the definition of the error function E , one may rewrite Eq. (F-3) in the

following forms:

Page 222: A Study of Non-Newtonian Viscosit y and Yield Stress of

207

[ ]

[ ]

[ ] 0);();(

0);();(

0);();(

)(

1

)()(

2

)(

1

)()(

1

)(

1

)()(

=∂

∂−

=∂

∂−

=∂

∂−

=

=

=

m

jl

j

jj

jl

j

jj

jl

j

jj

aAXfyAXf

aAXfyAXf

aAXfyAXf

M

(F-4)

Considering the non-linear cases, the standard form for these problems is Eq. (F-4).

0)(

0)(0)(

2

1

=

==

AF

AFAF

m

M i.e.,

0)( =AF (F-5)

where F and A are vectors.

Supposed that an initial approximation, ),...,,( )0()0(2

)0(10 maaaA = , to a solution

of the system is provided, the Newton’s method can be used. The Newton’s method

is based on the Taylor Expansion, which can be expressed in matrix form as follows:

termsorderhigherAAAJAFAF )()()()( 000 +−⋅+= (F-6)

where J is the Jacobian matrix whose elements are evaluated at 0A . Since )(AF

should vanish, and the higher order terms can be assumed to be negligible, Eq. (F-6)

can be reduced to:

)()()( 000 AFAAAJ −=−⋅ (F-7)

The above equation is a linear system of equations, so one can solve it for 01 AA − .

Page 223: A Study of Non-Newtonian Viscosit y and Yield Stress of

208

APPENDIX G. Repeatability Study with Distilled Water

0.80.820.840.860.88

0.90.920.940.960.98

1

0 100 200 300 400 500Shear rate (s-1)

Vis

cosi

ty (c

P)

Test #1Test #2Test #3Test #4Test #5Reference (0.892 cP)

Fig. G-1. Repeatability study #1.

Page 224: A Study of Non-Newtonian Viscosit y and Yield Stress of

209

0.80.820.840.860.88

0.90.920.940.960.98

1

0 100 200 300 400 500

Shear rate (s-1)

Vis

cosi

ty (c

P)

Test #1Test #2Test #3Test #4Test #5Reference (0.892 cP)

Fig. G-2. Repeatability study #2.

Page 225: A Study of Non-Newtonian Viscosit y and Yield Stress of

210

APPENDIX H. Experimental Data

Table H-1. A typical experimental data set of human blood obtained by a scanning capillary-tube rheometer with precision glass riser tubes. One out of 100 data points is selected from an original data set.

Time (s) Pixel Number at Riser Tube 1 Pixel Number at Riser Tube 2 0.00 1183 399 2.00 1040 460 4.00 1036 509 6.00 1036 550 8.00 1004 583

10.00 976 609 12.00 946 632 14.00 922 650 16.00 907 666 18.00 894 679 20.00 885 690 22.00 891 700 24.00 882 707 26.00 875 714 28.00 868 720 30.00 863 725 32.00 859 730 34.00 854 734 36.00 851 737 38.00 847 740 40.00 845 742 42.00 842 743 44.00 840 745 46.00 838 747 48.00 836 748 50.00 834 750 52.00 833 751 54.00 832 752 56.00 830 753 58.00 829 754 60.00 828 754

Page 226: A Study of Non-Newtonian Viscosit y and Yield Stress of

211

Table H-1. Continued.

Time (s) Pixel Number at Riser Tube 1 Pixel Number at Riser Tube 2 62.00 827 755 64.00 826 756 66.00 825 756 68.00 825 757 70.00 824 757 72.00 823 757 74.00 822 758 76.00 822 758 78.00 821 758 80.00 820 759 82.00 820 759 84.00 819 759 86.00 819 759 88.00 818 759 90.00 818 760 92.00 818 760 94.00 817 760 96.00 817 760 98.00 816 760 100.00 816 760 102.00 816 760 104.00 815 760 106.00 815 760 108.00 814 760 110.00 814 760 112.00 814 760 114.00 813 760 116.00 813 760 118.00 813 759 120.00 812 759 122.00 812 759 124.00 812 759 126.00 812 759 128.00 811 759 130.00 811 759

Page 227: A Study of Non-Newtonian Viscosit y and Yield Stress of

212

Table H-1. Continued.

Time (s) Pixel Number at Riser Tube 1 Pixel Number at Riser Tube 2 132.00 811 759 134.00 811 758 136.00 811 758 138.00 810 758 140.00 810 758 142.00 810 758 144.00 809 758 146.00 809 758 148.00 809 758 150.00 809 758 152.00 809 758 154.00 808 758 156.00 808 758 158.00 808 758 160.00 808 758 162.00 809 759 164.00 808 758 166.00 808 758 168.00 808 758 170.00 808 758 172.00 807 758 174.00 808 758 176.00 807 758 178.00 807 758 180.00 807 758

Page 228: A Study of Non-Newtonian Viscosit y and Yield Stress of

213

Table H-2. A typical experimental data set of distilled water obtained by a scanning capillary-tube rheometer with plastic riser tubes. One out of 100 data points is selected from an original data set.

Time (s) Pixel Number at Riser Tube 1 Pixel Number at Riser Tube 2 0.00 568 130 2.00 411 288 4.00 292 407 6.00 220 478 8.00 178 522

10.00 152 548 12.00 137 561 14.00 128 571 16.00 122 576 18.00 118 580 20.00 117 583 22.00 115 584 24.00 115 585 26.00 114 585 28.00 112 585 30.00 112 586 32.00 112 586 34.00 112 586 36.00 111 587 38.00 111 587 40.00 111 587 42.00 111 587 44.00 111 587 46.00 111 587 48.00 111 587 50.00 111 587 52.00 111 587 54.00 111 587 56.00 111 587 58.00 111 587 60.00 111 587 62.00 111 587 64.00 111 587 66.00 111 587 68.00 111 587 70.00 111 587

Page 229: A Study of Non-Newtonian Viscosit y and Yield Stress of

214

Table H-2. Continued.

Time (s) Pixel Number at Riser Tube 1 Pixel Number at Riser Tube 2 72.00 111 587 74.00 111 587 76.00 111 587 78.00 111 587 80.00 111 587 82.00 111 587 84.00 111 587 86.00 111 587 88.00 111 587 90.00 111 587 92.00 111 587 94.00 111 587 96.00 111 587 98.00 111 587 100.00 111 587 102.00 111 587 104.00 111 587 106.00 111 587 108.00 111 587 110.00 111 587 112.00 111 587 114.00 111 587 116.00 111 587 118.00 111 587 120.00 111 587 122.00 111 587 124.00 111 587 126.00 111 587 128.00 111 587 130.00 111 587 132.00 111 587 134.00 111 587 136.00 111 587 138.00 111 587 140.00 111 587 142.00 111 587 144.00 111 587 146.00 111 587 148.00 111 587 150.00 111 587

Page 230: A Study of Non-Newtonian Viscosit y and Yield Stress of

215

Table H-3. A typical experimental data set of bovine blood obtained by a scanning capillary-tube rheometer with plastic riser tubes. One out of 100 data points is selected from an original data set.

Time (s) Pixel Number at Riser Tube 1 Pixel Number at Riser Tube 2 0 550 115 2 520 152 4 481 182 6 455 210 8 430 235

10 405 259 12 384 280 14 365 299 16 349 316 18 332 331 20 319 345 22 304 359 24 288 370 26 278 381 28 268 390 30 259 399 32 251 408 34 244 415 36 237 422 38 231 428 40 224 435 42 218 440 44 213 445 46 208 450 48 203 455 50 198 458 52 193 462 54 190 465 56 187 469 58 184 472 60 181 475 62 179 477 64 176 480 66 174 482 68 172 485 70 170 487

Page 231: A Study of Non-Newtonian Viscosit y and Yield Stress of

216

Table H-3. Continued.

Time (s) Pixel Number at Riser Tube 1 Pixel Number at Riser Tube 2 72 168 489 74 166 491 76 164 492 78 162 494 80 161 496 82 159 497 84 158 498 86 156 500 88 155 501 90 154 502 92 153 503 94 152 504 96 150 505 98 150 506 100 149 507 102 148 508 104 147 509 106 146 510 108 146 511 110 145 511 112 144 512 114 144 512 116 143 513 118 143 513 120 142 514 122 142 515 124 141 515 126 141 516 128 140 516 130 140 517 132 139 517 134 139 517 136 139 518 138 138 518 140 138 518 142 138 518 144 138 519 146 137 519 148 137 519 150 137 519

Page 232: A Study of Non-Newtonian Viscosit y and Yield Stress of

217

Table H-3. Continued.

Time (s) Pixel Number at Riser Tube 1 Pixel Number at Riser Tube 2 152 137 520 154 136 520 156 136 520 158 136 520 160 136 521 162 135 521 164 135 521 166 135 522 168 135 522 170 134 522 172 134 522 174 134 523 176 133 523 178 133 523 180 159 555

Page 233: A Study of Non-Newtonian Viscosit y and Yield Stress of

218

Table H-4. A typical experimental data set of human blood obtained by a scanning capillary-tube rheometer with plastic riser tubes. One out of 100 data points is selected from an original data set.

Time (s) Pixel Number at Riser Tube 1 Pixel Number at Riser Tube 2 0.00 584 133 2.00 535 185 4.00 484 235 6.00 446 273 8.00 411 311

10.00 380 341 12.00 356 366 14.00 335 387 16.00 315 408 18.00 299 423 20.00 283 437 22.00 272 450 24.00 261 460 26.00 252 469 28.00 244 476 30.00 237 482 32.00 230 491 34.00 224 497 36.00 219 502 38.00 214 507 40.00 210 510 42.00 206 513 44.00 201 517 46.00 196 520 48.00 193 522 50.00 192 525 52.00 189 527 54.00 187 529 56.00 186 531 58.00 184 533 60.00 182 534 62.00 181 535 64.00 179 537 66.00 178 538 68.00 177 539 70.00 176 540

Page 234: A Study of Non-Newtonian Viscosit y and Yield Stress of

219

Table H-4. Continued.

Time (s) Pixel Number at Riser Tube 1 Pixel Number at Riser Tube 2 72.00 176 541 74.00 175 542 76.00 174 543 78.00 173 544 80.00 172 545 82.00 171 545 84.00 171 546 86.00 170 546 88.00 170 547 90.00 169 547 92.00 169 548 94.00 168 548 96.00 168 549 98.00 167 549 100.00 167 550 102.00 167 550 104.00 166 550 106.00 166 551 108.00 166 551 110.00 165 551 112.00 165 551 114.00 165 552 116.00 165 552 118.00 164 552 120.00 164 552 122.00 164 552 124.00 163 552 126.00 163 553 128.00 163 553 130.00 163 553 132.00 162 553 134.00 162 553 136.00 162 553 138.00 162 553 140.00 162 553 142.00 161 553 144.00 161 554 146.00 161 554 148.00 161 554 150.00 160 554

Page 235: A Study of Non-Newtonian Viscosit y and Yield Stress of

220

Table H-4. Continued.

Time (s) Pixel Number at Riser Tube 1 Pixel Number at Riser Tube 2 152.00 160 554 154.00 160 554 156.00 160 554 158.00 160 554 160.00 160 554 162.00 160 555 164.00 160 555 166.00 160 555 168.00 159 555 170.00 159 555 172.00 159 555 174.00 159 555 176.00 159 555 178.00 159 555 180.00 159 555

Page 236: A Study of Non-Newtonian Viscosit y and Yield Stress of

221

VITA

Sangho Kim Education Drexel University Philadelphia, PA Doctor of Philosophy in Mechanical Engineering 12/2002 Master of Science in Mechanical Engineering

Kyungpook National University Taegu, Korea Bachelor of Science in Mechanical Engineering 02/1997 Journal Publications • S. Kim, Y.I. Cho, K.R. Kensey, R.O. Pellizzari and P.R. Stark

“A scanning dual-capillary-tube viscometer” Review of Scientific Instruments, Vol. 71, No. 8, August 2000, 3188-3192 • S. Kim, Y.I. Cho, A.H. Jeon, B. Hogenaeur and K.R. Kensey “A new method for blood viscosity measurement” J. Non-Newtonian Fluid Mechanics, 94 (2000) 47-56 • S. Kim, Y.I. Cho, W.N. Hogenaeur and K.R. Kensey “A method of isolating surface tension and yield stress effects in a U-shaped scanning capillary-tube viscometer using a Casson model” J. Non-Newtonian Fluid Mechanics, 103 (2002) 205-219 S. Kim and Y.I. Cho “The effect of dye concentration on the viscosity of water in a scanning capillary-tube viscometer” J. Non-Newtonian Fluid Mechanics, 2002 (submitted) S. Kim, Y.I. Cho, and W.N. Hogenaeuer “Non-Newtonian constitutive models for the viscosity and yield-stress measurements of blood using a scanning capillary-tube rheometer” Biorheology, 2002 (submitted)

Conference Publications

• S. Kim and Y.I. Cho “A new method of measuring blood viscosity with a U-shaped scanning capillary-tube viscometer using a Casson model” Proceedings of the IEEE 28th Annual Northeast Bioengineering Conference, April 20-21, 2002, 253-254 • Y.I. Cho and S. Kim “A new scanning capillary tube viscometer for blood viscosity measurement” Proceedings of the First National Congress on Fluids Engineering, September 1-2, Korea, 2000, 5-8

US Patents • U.S. Patent No. 6,428,488 • U.S. Patent No. 6,412,336

• U.S. Patent No. 6,322,524 • U.S. Patent No. 6,402,703 • U.S. Patent No. 6,450,974