viscous effects on penetrating shafts in clays

157
VISCOUS EFFECTS ON PENETRATING SHAFTS IN CLAY by SANDEEP PRAKASH MAHAJAN ________________________ A Dissertation Submitted to the Faculty of the DEPARTMENT OF CIVIL ENGINEERING AND ENGINEERING MECHANICS In Partial Fulfillment of the Requirements For the Degree of DOCTOR OF PHILOSOPHY WITH A MAJOR IN CIVIL ENGINEERING In the Graduate College THE UNIVERISTY OF ARIZONA 2 0 0 6

Upload: independent

Post on 21-Nov-2023

1 views

Category:

Documents


0 download

TRANSCRIPT

VISCOUS EFFECTS ON PENETRATING SHAFTS IN CLAY

by

SANDEEP PRAKASH MAHAJAN

________________________

A Dissertation Submitted to the Faculty of the

DEPARTMENT OF CIVIL ENGINEERING AND ENGINEERING MECHANICS

In Partial Fulfillment of the Requirements For the Degree of

DOCTOR OF PHILOSOPHY

WITH A MAJOR IN CIVIL ENGINEERING

In the Graduate College

THE UNIVERISTY OF ARIZONA

2 0 0 6

2

THE UNIVERSITY OF ARIZONA GRADUATE COLLEGE

As members of the Dissertation Committee, we certify that we have read the dissertation prepared by Sandeep Prakash Mahajan entitled Viscous Effects on Penetrating Shafts in Clay and recommend that it be accepted as fulfilling the dissertation requirement for the Degree of Doctor of Philosophy _________________________________________________________________Date: June 30, 2006

Muniram Budhu _________________________________________________________________Date: June 30, 2006

Achintya Haldar _________________________________________________________________Date: June 30, 2006

Chandrakant .S. Desai _________________________________________________________________Date: June 30, 2006

Dinshaw N. Contractor Final approval and acceptance of this dissertation is contingent upon the candidate’s submission of the final copies of the dissertation to the Graduate College. I hereby certify that I have read this dissertation prepared under my direction and recommend that it be accepted as fulfilling the dissertation requirement. ________________________________________________ Date: June 30, 2006 Dissertation Director: Muniram Budhu

3

STATEMENT BY AUTHOR

This dissertation has been submitted in partial fulfillment of requirements for an advanced degree at the University of Arizona and is deposited in the University Library to be made available to borrowers under rules of the Library.

Brief quotations from this dissertation are allowable without special permission,

provided that accurate acknowledgment of source is made. Requests for permission for extended quotation from or reproduction of this manuscript in whole or in part may be granted by the head of the major department or the Dean of the Graduate College when in his or her judgment the proposed use of the material is in the interests of scholarship. In all other instances, however, permission must be obtained from the author. SIGNED: ______________________ Sandeep Prakash Mahajan

4

ACKNOWLEDGEMENTS I would like to acknowledge many people for helping me during my doctoral work. I have no words to describe my gratitude towards my advisor, Dr. Budhu, for his generous time and commitment. Throughout my doctoral work he encouraged me to develop independent thinking and research skills. He continually stimulated my analytical thinking and greatly assisted me with scientific writing. I am very much grateful to Dr. Budhu for his friendship, counsel and above all for his critical scrutiny and comments of my writing which has immensely improved my technical writing skills. He has modeled a lesson I will gladly carry forward with me in my future work. I am also very grateful for having an exceptional doctoral committee and wish to thank Dr. Haldar, Dr. Desai, Dr. Contractor and Dr. Merry for their support and encouragement. They have generously given their time and I thank them for their contribution and their good-natured support. I am grateful to Dr. Valdes for giving me opportunities to teach undergraduate courses which significantly enhanced my teaching abilities. I extend many thanks to my colleagues and friends, especially Melissa Cox, Pawan Baheti, Juan Lopez and Dustin Agnew for their assistance at various times. Additional thanks to the staff Alice Stilwell, Karen Van Winkle, Lajeana Hall, Olivia Hanson, Tom Demma and Steve Albanese. Finally, I'd like to thank my family, especially my parents for their support and care over the years. I'm grateful to my wife, Anjali, for her continual encouragement and enthusiasm.

5

I dedicate this dissertation to supreme Lord Shree Ganesh for his grace.

6

TABLE OF CONTENTS

LIST OF ILLUSTRATIONS.............................................................................................. 9

LIST OF TABLES............................................................................................................ 11

ABSTRACT...................................................................................................................... 12

CHAPTER 1 : INTRODUCTION................................................................................... 14

1.1 General .................................................................................................................. 14

1.2 Failure strength of soils.......................................................................................... 15

1.3 Post failure strength of soils ................................................................................... 18

1.4 Problem statement................................................................................................. 18

1.5 Hypotheses ............................................................................................................ 21

1.6 Goal and objectives of this research ..................................................................... 21

1.7 Scope of this research ........................................................................................... 22

1.8 Research outcomes................................................................................................ 22

1.9 Organization of dissertation.................................................................................. 22

1.10 Key terms .............................................................................................................. 24

1.11 Notations ............................................................................................................... 25

CHAPTER 2 : LITERATURE REVIEW......................................................................... 28

2.1 Introduction........................................................................................................... 28

2.2 Soil failure due to penetration............................................................................... 29 2.2.1 General.............................................................................................................. 29 2.2.2 Critical State Model (CSM).............................................................................. 29

2.3 Modeling soil penetration ..................................................................................... 34 2.3.1 General.............................................................................................................. 34 2.3.2 Analytical studies.............................................................................................. 35

7

TABLE OF CONTENTS (continued) 2.3.3 Experimental study ........................................................................................... 41 2.3.4 Numerical studies.............................................................................................. 44 2.3.5 Summary ........................................................................................................... 45

2.4 Soil resistance on a penetrating shaft.................................................................... 46 2.4.1 General.............................................................................................................. 46 2.4.2 Skin friction and end bearing resistance ........................................................... 47 2.4.3 Viscous (friction) resistance ............................................................................. 48 2.4.4 Summary ........................................................................................................... 48

2.5 Viscous behavior of soils ...................................................................................... 49 2.5.1 Theory of viscoelastic deformation .................................................................. 49 2.5.2 Creep deformation ............................................................................................ 51 2.5.3 Viscous behavior at critical state ...................................................................... 55 2.5.4 Summary ........................................................................................................... 56

2.6 Viscous shear force ............................................................................................... 57 2.6.1 Shear viscosity .................................................................................................. 57 2.6.2 Viscous drag on shafts in clays......................................................................... 58 2.6.3 Creeping flow.................................................................................................... 59 2.6.4 Viscous drag on bodies in creeping flow.......................................................... 59 2.6.5 Effect of boundaries.......................................................................................... 62

2.7 Soil state around the shaft ..................................................................................... 62

2.8 Summary ............................................................................................................... 66

CHAPTER 3 : MATHEMATICAL FORMULATION AND ANALYSIS ..................... 67

3.1 Introduction........................................................................................................... 67

3.2 Post-failure response of soil as yield stress fluid .................................................. 67

3.3 Analytical method................................................................................................. 70 3.3.1 General.............................................................................................................. 70 3.3.2 Assumptions...................................................................................................... 72 3.3.3 Viscous drag on a penetrating shaft in clay ...................................................... 72

3.4 Results and discussion of analysis ........................................................................ 80 3.4.1 Parameters influencing viscous drag ................................................................ 80 3.4.2 Effects of the size of CS zone ........................................................................... 80 3.4.3 Velocity profile within CS zone ....................................................................... 82 3.4.4 Shear viscosity of clay ...................................................................................... 83

8

TABLE OF CONTENTS (continued)

3.5 Conclusion ............................................................................................................ 86

CHAPTER 4 : SHEAR VISCOSITY OF CLAYS AT CS............................................... 87

4.1 Introduction........................................................................................................... 87

4.2 Current Investigations of soils viscosity ............................................................... 88 4.2.1 General.............................................................................................................. 88 4.2.2 Landslides and earth flows................................................................................ 88 4.2.3 Strain rate effects .............................................................................................. 91

4.3 Modeling of viscous behavior............................................................................... 93

4.4 Shear viscosity using the fall cone test ................................................................ 94 4.4.1 Penetration test.................................................................................................. 94 4.4.2 The fall cone test............................................................................................... 96 4.4.3 Theoretical approach....................................................................................... 101 4.4.4 Experiments .................................................................................................... 105

4.5 Results and discussion of the experimental study............................................... 108

4.6 Application to the CPT ....................................................................................... 118

4.7 Conclusion .......................................................................................................... 119

CHAPTER 5 : SUMMARY, CONCLUSIONS AND RECOMMENDATIONS ......... 121

5.1 Summary and conclusions .................................................................................. 121

5.2 Key Contributions............................................................................................... 123

5.3 Potential applications .......................................................................................... 123

5.4 Recommendations for future research ................................................................ 124

APPENDIX A : MATLAB CODES............................................................................... 125

APPENDIX B : FALL CONE TEST RESULTS .......................................................... 132

APPENDIX C : APPLICATION USING AN ILLUSTRATIVE EXAMPLE .............. 144

REFERENCES ............................................................................................................... 147

9

LIST OF ILLUSTRATIONS

Figure 1.1 Typical stress-strain response of soils at constant normal effective stress and

interpretation of peak and critical state friction angles. ............................................ 16

Figure 1.2 (a) Soil state at the tip of the penetrating shaft (b) soil flow around the penetrating shaft........................................................................................................ 20

Figure 2.1 Three dimensional plot of yield surface in CSM............................................. 30

Figure 2.2 Critical State Model (CSM)…………………………………………………..31

Figure 2.3 (a) Failure patterns under deep foundations (b) Expanding Cavity in an infinite soil mass (Vesic, 1972) ............................................................................................. 36

Figure 2.4 Deep penetration viewed as a steady flow problem (Baligh, 1985)............... 38

Figure 2.5 Plastic soil flow around a spherical penetrometer (Randolph et al. 2000)...... 42

Figure 2.6 Disturbance caused at the conical tip of a penetrometer (Allersma, 1987)..... 43

Figure 2.7 Mechanical models, E = elastic body, V = Viscous, Newtonian body and P = plastic body............................................................................................................ 50

Figure 2.8 Models of viscoelastic bodies (a) Maxwell body (b) Kelvin Body................. 52

Figure 2.9 Creep and viscous flow deformation............................................................... 54

Figure 2.10 Creeping Flow Past (a) a Sphere (b) a circular disk (Van Dyke, 1982)........ 61

Figure 2.11 Soil Disturbance around Penetrating Shafts in Soft Clay (Zeevart, 1948).... 63

Figure 3.1 Typical flow curves for yield stress fluids. ..................................................... 69

Figure 3.2 Analogy of the post-failure response of a soil to a yield stress fluid response.................................................................................................................................... 71

Figure 3.3 Forces acting on a soil element in the CS zone around the shaft .................. 773

Figure 3.4 Velocity profile of the viscous soil in the CS zone ......................................... 77

Figure 3.5 Effect of size of the CS zone on the viscous drag on a cylindrical shaft ........ 81

Figure 3.6 Velocity profiles for different sizes of the CS zone ........................................ 84

Figure 3.7 Upward movement of soil in the CS zone....................................................... 85

10

LIST OF ILLUSTRATIONS (continued)

Figure 4.1 Soil flow of sensitive clay due to heavy rainfall (Reference: Geological Survey of Canada)................................................................................................................. 90

Figure 4.2 The fall cone test apparatus ............................................................................. 95

Figure 4.3 Fall cone and the soil cup ................................................................................ 98

Figure 4.4 Liquid limit from typical fall cone test results ................................................ 99

Figure 4.5 Illustration of the soil state around a fall cone .............................................. 102

Figure 4.6 Modified experimental setup for the fall cone test ........................................ 106

Figure 4.7 LVDT connected to the top of fall cone shaft ............................................... 107

Figure 4.8 Penetration-time relationship of the cone (Test no. 15C1)............................ 110

Figure 4.9 (a) Velocity and (b) acceleration of the cone (Test no. 15C1) ...................... 112

Figure 4.10 Fall cone energy (Ec) ................................................................................... 114

Figure 4.11 PE, KE and Ec with penetration depth (Test no. 15C1) ............................. 115

Figure 4.12 Shear viscosity - LI relationship for kaolin used in this study ................... 117

Figure B1 (a) Penetration-time relationship (b) velocity of the cone for test 15C2 ....... 133

Figure B2 (a) Penetration-time relationship (b) velocity of the cone for test 15C3 ....... 134

Figure B3 (a) Penetration-time relationship (b) velocity of the cone for test 15C4 ....... 135

Figure B4 (a) Penetration-time relationship (b) velocity of the cone for test 15C5 ....... 136

Figure B5 (a) Penetration-time relationship (b) velocity of the cone for test 15C6 ....... 137

Figure B6 (a) Penetration-time relationship (b) velocity of the cone for test 5C1 ......... 138

Figure B7 (a) Penetration-time relationship (b) velocity of the cone for test 5C2 ......... 139

Figure B8 (a) Penetration-time relationship (b) velocity of the cone for test 5C3 ......... 140

Figure B9 (a) Penetration-time relationship (b) velocity of the cone for test C1 ........... 141

Figure B10 (a) Penetration-time relationship (b) velocity of the cone for test C2 ......... 142

Figure B11 (a) Penetration-time relationship (b) velocity of the cone for test C3 ......... 143

11

LIST OF TABLES Table 1 Experimental details .......................................................................................... 109 Table 2 Test data............................................................................................................. 113 Table 3 Estimated shear strength and shear viscosity..................................................... 116

12

ABSTRACT

When a rigid shaft such as a jacked pile or the sleeve of a cone penetrometer penetrates

soil, the soil mass at the shaft tip fails. This failed soil mass flows around the shaft

surface and creates a disturbed soil zone. The soil in this zone, which is at a failure or

critical state (CS), flows and behaves like a viscous fluid. During continuous penetration,

the shaft surface is subjected to an additional viscous shear stress above the static shear

stress (interfacial solid friction). The total resistance on the shaft in motion is due to the

static and viscous shear components. Current methods of calculating the penetration

resistance in soils are based on static interfacial friction, which determine the force

required to cause failure at the shaft-soil interface and not the viscous drag. The main

aim of this research is to understand the viscous soil resistance on penetrating shafts in

clays.

This research consists of two components. First, a theoretical analysis based on creeping

flow hydrodynamics is developed to study the viscous drag on the shaft. The results of

this analysis reveal that the size of the CS zone, the shear viscosity of the soil and

velocity of the shaft influence the viscous drag stress. Large increases in viscous drag

occur when the size of the CS zone is less than four times the shaft radius.

13

Second, a new experimental procedure to estimate the shear viscosity of clays with water

contents less than the liquid limit is developed. Shear viscosity is the desired soil

parameter to estimate viscous drag. However, there is no standard method to determine

shear viscosity of clays with low water contents (or Liquidity Index, LI). Soils can reach

CS for water contents in the plastic range (LI<1) and exhibit viscous behavior. The fall

cone test is widely used to interpret the index (liquid and plastic limit) and strength

properties of clays. In this study the existing analysis of the fall cone test is reexamined

to discern the viscous drag as the cone penetrates the soil. This reexamination shows that

the shear viscosity of clays with low water contents (LI<1.5) can be estimated from time-

penetration data of the fall cone. Fall cone test results on kaolin show that the shear

viscosity decreases exponentially with an increase in LI.

The results of this research can be used to understand practical problems such as jacked

piles in clays, cone penetrometer sleeve resistance and advancement of casings in soil for

drilling or tunneling operations.

14

CHAPTER 1

INTRODUCTION

1.1 General

Several studies exist in geotechnical engineering to predict failure and pre-failure

behavior of soils for calculating their load carrying capacities and settlements at working

loads. In these methods the pre-failure or the failure responses (Figure 1.1) of soil have been

investigated with the help of some ideal material behavior (e.g. elastic, rigid-plastic and

elasto-plastic). To determine failure, the undrained shear strength ( us ) and the effective

friction angle (φ′) are essential strength parameters used for total and effective stress

analyses respectively. The shear strains in these analyses are restricted to the pre-failure state

or until the soil reaches failure strength. In this chapter, the following are presented.

1. A brief understanding on failure and post-failure strength of soils.

2. The problem statement of this research.

3. The hypotheses, objectives and the scope of this research.

4. A guide to the reader on the organization of this dissertation, key terms and

notations.

15

1.2 Failure strength of soils

When a soil is subjected to shearing forces, the soil deforms with changes in stresses

and strains. Under a constant vertical (normal) effective stress, n′σ , all soils tend to reach an

approximately constant shear stress and constant void ratio for continued shearing. Consider

a soil in loose and dense states, sheared at the same vertical effective stress. Loose (non

dilative) soils show a gradual increase in shear stresses as the shear strain increases (strain

hardens) until an approximately constant shear stress, called critical state (CS) shear stress,

csτ , is attained (Figure 1.1 a). Loose soils compress and become denser until a constant void

ratio, called critical void ratio, cse , is reached (Figure 1.1 b). Dense soils show a rapid

increase in shear stress reaching a peak shear stress, pτ , at low shear strains and then show a

decrease in shear stress with increasing shear strain (strain softens), ultimately attaining a CS

shear stress (Figure 1.1 a). The CS shear stress increases with increasing n′σ . Dense soils

compress initially and then expand until a critical void ratio (equal to that of loose soil for

the same n′σ ) is attained. The final state of soil is the CS, a state in which the material

undergoes continuous soil deformation under constant volume and constant shear stress

ratio. Constant shear stress ratio is a ratio of deviatoric stress, q, divided by mean effective

pressure, p′ . Failure and CS are synonymous in this dissertation.

The shear strength of soils is due to friction and interlocking of soil particles.

Coulomb’s frictional law (1776) forms the basis of determining the failure stress in most

.

16

Figure 1.1 Typical stress-strain response of soils at constant normal effective stress and

interpretation of peak and critical state friction angles.

n′σ

τ

(b)

cs'φ p'φ

τ

(c)

Critical shear stress

τp

τcs

γ

Peak Dilative soil

Non-dilative soil

(a)

γ

Dilative soil

Non-dilative soil

e

Shear strain

Shear strain

Shea

r stre

ss

ecs

Normal effective stress

Critical void ratio

C

P

C

P

Critical state

Peak

C

Voi

d ra

tio

Shea

r stre

ss

Crit

ical

stat

e

17

geotechnical engineering applications. The frictional force on the slip plane according to

Coulomb’s law can be written as

sF N= µ [1.1]

where F is the interface frictional force, N is the normal force on the slip plane and sµ is

the coefficient of static friction between two rigid surfaces. The material frictional

behavior is often expressed as 's tanµ = φ , in which 'φ is the frictional property of the

material known as the internal friction angle.

Coulomb’s failure law to compute the failure shear stress for an effective stress

analysis (ESA) can be written as

n tan( )′ ′τ = σ φ [1.2]

where n′σ is the normal effective stress on the slip plane or the interface, τ is the shear

stress acting on the slip plane or the interface. The internal friction angle is the most

important parameter for an ESA in geotechnical engineering.

It is common to identify two values of friction angles for soils (Figure 1.1 c); the peak

friction angle, p'φ and the critical state friction angle, cs′φ . The peak friction angle is

substituted in Equation [1.2] to compute pτ , as observed in dense soils. CS friction angle is

used to compute csτ . The peak friction angle is not a fundamental soil parameter. It depends

on the capacity of the soil to dilate (expand), which is influenced by the arrangement

(packing) of soil particles and the normal effective stress on the failure plane. The CS

friction angle is constant for a given soil and is a fundamental soil parameter.

18

1.3 Post failure strength of soils

Coulomb’s equation gives the information on the soil strength when failure on a slip

plane is initiated. However, there is a certain class of problems in geoengineering which

comprise post-failure soil response. In such response there is no distinct failure plane but the

associated soil mass flows like a fluid after reaching CS (failure). Such problems include

mudslides and soil flow around penetrating rigid bodies such as the shaft surface of a jacked

pile, a sleeve of a cone penetrometer or the installation of spud-can footings for offshore

structures in soft clays. To investigate these problems, it is necessary to model the flow

behavior of soil at CS and beyond.

Soils at CS are similar to yield stress fluids with the yield stress equal to the CS shear

strength. If applied stress on a soil at CS exceeds the yield value the soil will flow like a

viscous fluid in its post-failure response. The viscous soil flow occurs at low Reynolds

number - a flow analogous to creeping flow in hydrodynamics. However, viscous flow has

received scant attention in geomechanics because of the overwhelming need to study the

pre-failure soil behavior.

1.4 Problem statement

Penetration problems such as a cone penetrometer test (CPT) or installation of jacked

piles in clays are common in onshore and offshore geotechnical engineering. When a rigid

shaft penetrates a fine grained soil, the soil mass at and near the tip is subjected to high

19

stresses and fails i.e. reaches CS (Fig. 1.4.1a). The clay at the tip of the shaft then flows

adjacent to the shaft surface during continuous penetration (Fig. 1.4.1b). A zone of

disturbance is created around the surface of the shaft. When the shaft is in motion, the soil

flow offers resistance to the moving shaft. The resistance of a shaft during penetration will

then depend on the stress-strain (solid) relationship and the flow (viscous) properties of the

soil.

The current methods of estimating the shaft resistance in clays utilize the static shear

strength properties of soil. These available methods are inadequate to determine the viscous

drag. A potential approach to analyze viscous drag is to use hydrodynamics principles of

creeping flow.

In a hydrodynamics method, shear viscosity of the soil will be an essential parameter

to compute viscous drag. Shear viscosity has been investigated for soils with higher water

contents (greater than the liquid limit) using viscometers. However, soils can reach CS at

water contents less than their liquid limits. Available tests to determine the viscosity of soils

are not suitable for soils with low water contents. A potential test to determine the shear

viscosity of soil at low water contents is the fall cone test, which is currently used to

measure the index and shear properties of fine grained soils (clays).

20

Figure 1.2 (a) Soil state at the tip of the penetrating shaft (b) soil flow around the

penetrating shaft

Failed soil mass

Undisturbed soil

τ

Critical shear stress τcs

γ Shear strain Sh

ear s

tress

C

(b)

(a)

21

The intention of this study is to apply viscous flow (hydrodynamics) principles to soils

at CS with the main purpose of understanding the effects of viscous soil resistance on

penetrating shafts or objects in clay.

1.5 Hypotheses

The hypotheses for this study are:

1. Viscous drag is a component of the total penetration resistance offered to rigid

bodies (e.g. shaft) during continuous penetration in clay and can be analyzed

using creeping flow hydrodynamics.

2. The fall cone test is a potential tool to determine shear viscosity of clays with

low water contents.

1.6 Goal and objectives of this research

The goal of the research is to understand the viscous drag on penetrating shafts in

clays. To meet this goal, the following objectives were established:

Objective 1: Develop an analytical (theoretical) method based on creeping flow

hydrodynamics to determine the viscous drag on penetrating shafts in clay.

22

Objective 2: Develop an experimental arrangement and procedure to estimate the

shear viscosity of clays.

1.7 Scope of this research

1. An analytical method will be developed for a cylindrical shaft penetrating

in soft clay at a constant rate.

2. Skin friction on the shaft surface will be analyzed. End bearing resistance

is not studied.

1.8 Research outcomes

The contribution of this study is an understanding of the post-failure soil flow in soil

penetration problems. A new analytical approach based on creeping flow principles in

hydrodynamics is developed to investigate such problems. Shear viscosity of soil at failure

(CS) is identified as a key soil parameter, which can be estimated by a new experimental

procedure using a fall cone test.

1.9 Organization of dissertation

This research is described and presented through different chapters. A brief summary

of the topics included in the chapters is as follows:

23

Chapter 1: The introduction, problem statement, hypothesis, objectives and scope of this

research are posed in this chapter to give an overall picture of this study.

Chapter 2: The studies and theories conducted in understanding soil behavior during

penetration process are reviewed. These include:

1. the critical state (CS) model used to predict soil response

2. current approach used to determine the penetrating shaft resistance

3. the modeling of viscous behavior of soil for creep deformation studies

4. the analogy of soil behavior at CS as viscous fluid

5. the differences in viscous behavior at CS and in creep deformation and

6. the theory of computing viscous drag on objects in creeping flow

Chapter 3: The assumptions made for this analytical study are listed. The mathematical

formulations to derive the equation for viscous drag on a cylindrical penetrating shaft are

explained. Based on the derived equation, the parameters influencing the viscous drag on

a shaft are discussed.

Chapter 4: The importance of shear viscosity of soils to determine the viscous drag is

presented and discussed. Viscosities of soil and experimental studies for its measurement

are reviewed. The fall cone test is reviewed and its theory is extended to measure shear

viscosity. The experimental procedure and setup proposed to measure the shear viscosity

by using the extended theory is explained. Experimental measurements, computations

24

and results are presented and discussed. An illustrative example is worked out to describe

the application of this study in measuring viscous drag on the friction sleeve in a CPT

test.

Chapter 5: This dissertation is concluded listing the key results and findings, key

contributions, potential applications of this study. Recommendations for future research

are also listed in this chapter.

1.10 Key terms

Unless otherwise stated, the following definitions apply to some common terms in

this dissertation:

CPT: Cone penetrometer test

CS: Critical state, a state at which continuous soil deformation occurs under constant

volume and constant shear stress ratio.

CS zone: Annular region around the shaft surface in which the soil is at CS

Sleeve: Cylindrical sleeve used to measure friction resistance in a CPT. The sleeve has an

outer diameter equal to the base diameter of the cone and a cross-sectional area of 10 cm2

Soils: Clays

Failure: When soils reach CS

CS shear stress (strength): Constant shear (stress) strength of the soil at CS.

25

Viscous soil: Clay which is at CS and behaves like a viscous fluid.

Creeping flow: Flow of viscous fluids at very low Reynolds number, that is, in general, a

very slow viscous flow.

Post-failure flow (response): Viscous flow of the soil that has attained the CS.

Viscosity: Shear viscosity which offers shear resistance to flow.

Liquid limit: Water content of the soil at which there is a transition of the soil phase from

plastic to liquid state.

Liquidity index (LI): Quantitative measure of the current soil state, value less than 0

signify the solid state and values greater than 1 signifies liquid state. Any value between

0 and 1 indicates that the soil is in plastic range.

Low LI: LI less than 1.5.

Units: The units in this dissertation follow the SI system. 1.11 Notations

a acceleration of cone

C viscous drag force constant

D outer diameter of the CPT friction sleeve

F non-dimensional cone resistance factor fv viscous drag force per unit of shaft

fz total resisting force per unit length of shaft in z direction

fs measured sleeve friction in CPT

26

fss static friction stress on friction sleeve

fsv viscous drag stress on friction sleeve

h penetration depth of cone heq dynamic equilibrium depth of cone

sh penetration depth for static equilibrium of cone

fh final penetration depth of cone L characteristic length of the object moving in viscous fluid

LI liquidity index K fall cone factor (constant) to determine soil shear strength m mass of cone

chN cone bearing capacity factor accounting for the heave around the cone p′ mean effective pressure

q deviatoric stress

Q volume of viscous soil flow at a horizontal cross-section in the CS zone

r radial distance from the center of the shaft

0r radius of the cylindrical shaft

R radius of the sphere

eR Reynolds number

Ro radius of cylindrical annulus of viscous soil (radius of CS zone)

su undrained shear strength

u vertical velocity of the soil flow in the CS zone

27

V velocity moving in viscous fluid

zV velocity of the shaft in z direction

W weight of cone

ρ mass density of viscous fluid

µ dynamic viscosity of a fluid

pµ shear viscosity of (clay) soil

λ cs /τ τ

oλ 0 0R / r

oβ dimensionless parameter representing the size of critical state zone

τ total (dynamic) shear resistance stress

csτ static shear resistance (critical state shear) stress

vτ viscous drag stress

yτ yield stress γ& shear strain rate

′φ internal friction angle of soil

cs′φ critical state friction angle ξ sv sf / f δ inclination angle (in radians) of the heaved soil surface θ half cone apex angle

28

CHAPTER 2

LITERATURE REVIEW

2.1 Introduction

Soil penetration problems are regarded amongst the most challenging problems in

geotechnical engineering. Penetration problems are frequently associated with the

evolution of large deformations and strains, which complicate the analysis procedures. As

a result, there exist disparate ideas for understanding these problems. Studies are

available to understand soil failure and the stress-strain behavior of the soil both at the

shaft tip and adjacent to the shaft surface. Literature on soil penetration analyses and

studies is summarized in this chapter.

The continuous penetration of a shaft is associated with flow of soil adjacent to

the shaft surface. The hypothesis advocated in this dissertation is that the mass of soil

adjacent to the shaft surface will flow like viscous fluid exhibiting creeping (slow

viscous) characteristics. The body in creeping viscous flow is subjected to viscous

resistance (drag). Basic background on creeping flow and computation of viscous drag

force is included this chapter.

29

2.2 Soil failure due to penetration

2.2.1 General

When a rigid shaft penetrates into a soil, the soil mass at the tip of the shaft is

subjected to large stresses and strains. This soil fails and reaches critical state (CS). A

region of soil failure is formed at the shaft tip (Figure 1.2 a). On further penetration of the

shaft, the failed soil mass at the tip flows adjacent to the shaft surface (Figure 1.2 b).

Critical State Model (CSM) is one of the models used to analyze pre-failure or failure

behavior of soils. CSM is described briefly in the following Section 2.2.2.

2.2.2 Critical State Model (CSM)

The CSM is expressed in terms of three variables: the mean effective pressure, 'p ,

the deviatoric (shear) stress, q, and the void ratio, e. A soil is assumed to fail on a unique

failure surface in ( q , 'p , and e) space as shown in Figure 2.1. The three dimensional

figure of CSM can be understood better in two dimensional space by plotting a critical

state line (CSL) separately in (q, 'p ) and ( 'p ,e ) as shown in Figure 2.2. CSM is based on

the concept that all soils under shearing will reach a CS (Figure 2.2), which is attained at

a constant shear stress ratio and constant volume. The constant shear stress

.

30

Yield surface

Critical state line

Yield surface

Normal consolidation line

Reconsolidation line

p′

e

q

Figure 2.1 Three dimensional plot of yield surface in CSM

31

Figure 2.2 Critical State Model (CSM)

cp'

e

'p1

CSL

fp′cse C

op'

'p

q

cp'

Slope,Mfq

Critical state line (CSL)

fp′

C

op'

Yield surface Stress

path

cp'

e

Γe

'p1

CSL

fp′

cse C

(ln scale)

λ

κ

Normal consolidation line (NCL)

The yield surface expands during strain hardening and contracts during strain softening. At critical state the soil undergoes continuous shearing at constant volume and constant stress

ratio Mpq

'f

f =

32

ratio is f'f

qp

⎛ ⎞⎜ ⎟⎝ ⎠

in which fq is the deviatoric stress at failure (CS) and 'fp is the mean

effective stress at failure.

CSM is popular to interpret and predict soil responses subject to various loadings.

Triaxial test data is commonly used in analyses using the CSM. For a triaxial test, mean

effective pressure, 'p , and deviatoric stress, q, are computed as:

' '' 1 32p

3σ + σ

= [2.1]

' '1 3q = σ −σ [2.2]

where '1σ and '

3σ are the major and minor principal effective stresses.

According to CSM, the failure stress state is insufficient to guarantee failure; the

soil structure must also be loose enough. Irrespective of the initial density, a given soil at

a given mean effective stress will reach the same constant density or critical void

ratio, cse , at failure (Figure 1.1 b). In the case of loose soil, the volume decreases

continuously until a critical void ratio is reached. For a dense state of the same soil, the

soil initially compresses and then dilates or increases in volume until it approaches the

same cse as the loose soil. For undrained conditions, the failure occurs under constant

volume and the initial void ratio, e0, is same as the failure (CS) void ratio.

According to CSM, the soil behaves like an elastic material up to certain load

combination and then yields (or behaves) plastically. The Cam-Clay model (Schofield

and Wroth, 1968), which is based on the CS concept, incorporates the volume changes,

elastic strains and plastic yielding to predict soil response. The Cam-Clay model is an

33

incremental hardening/softening elasto-plastic model that is widely used to predict the

stress-strain behavior of soft clays.

The elements of the Cam-Clay models are (1) a yield surface which separates the

elastic and plastic states of the soil (2) a flow rule that governs the hardening or softening

behavior during yielding and (3) a failure law. There is an initial yield surface for all soils

based on the preconsolidation mean effective stress, cp′ . The initial yield surface

represents the loading history of the soil. The yield surface is represented by an ellipse

passing through the origin. The yield surface of the popular Modified Cam-Clay model

(Roscoe and Burland, 1968) takes the form of an ellipsoid (Figure 2.2), which is defined

as follows:

( )22 2 2cM p M p p q 0′ ′ ′− + = [2.3]

where M is the frictional constant and the slope of the CSL in (q, 'p ) space, and cp′ is

preconsolidation mean effective stress. Hardening is modeled through the expansion of

the initial yield surface while softening is modeled by the contraction of the initial yield

surface.

Schofield and Wroth (1968) proposed that a given soil will fail at the CS such

that:

f fq Mp′= [2.4]

and

cs fe e ln pΓ ′= −λ [2.5]

34

where λ is the compression index and Γe is the void ratio corresponding to 'p = 1 kPa on

the failure line of a plot of void ratio versus mean effective stress (Figure 2.2) and

f 1 3 cs

f 1 3 cs

q 3( ) 6sinMp 2 3 sin

′ ′ ′σ −σ φ= = =

′ ′ ′ ′σ + σ − φ [2.6]

Equation [2.6] is an established mathematical relationship between M and cs′φ . However

they are conceptually different. The friction angle, φ′cs, obtained from Coulomb criterion

is sliding friction at the interface of two solid bodies whereas M is an internal friction

between the particles of a failed soil mass at CS. According to Schofield and Wroth

(1968), at CS, a soil “behaves as a frictional fluid rather than yielding as a solid; it is as

though the material had melted under stress.” At CS the soil deforms (or flows) at

constant volume similar to an incompressible viscous fluid.

2.3 Modeling soil penetration

2.3.1 General

Research on soil penetration problems is common in geotechnical engineering.

These studies are intended to understand the soil resistance and deformations linked with

the penetration process. Most of these studies focus on the analyses of cone penetration

test (CPT) and sampling disturbances caused by sampling tubes. Penetration resistance

measured from in-situ penetration tests (e.g. CPT) is utilized to interpret the soil

parameters used in stability analyses. CPT data is commonly used in the design of pile

35

foundations. The measured cone tip resistance and sleeve friction are used to estimate the

end bearing capacity and skin friction stress, respectively.

Large deformations result during soil penetration. Understanding soil deformations

and disturbances is important in order to have an initial perception to predict soil

behavior. Studies have been conducted to understand the soil behavior around penetrating

objects in soils. These studies are based on analytical methods (e.g. Meyerhof, 1951;

Vesic, 1972; Baligh, 1985; Randolph et al., 2000; Sagaseta and Whittle, 2001),

experimental methods (e.g. Allersma, 1987) and numerical methods (e.g. Teh and

Houlsby, 1991; Budhu and Wu, 1992; van den Berg, 1994; Yu et al., 2000). These

analyses are briefly discussed in subsequent sections.

2.3.2 Analytical studies

Initial investigations of penetration problems are founded on bearing capacity and

cavity expansion theories. In a cone penetration analysis, the cone base is analyzed

similar to a pile base assumed as a deep circular footing (e.g. Meyerhof, 1951). A failure

pattern is assumed at the deep footing base (Figure 2.3 a). Equilibrium equations are used

to determine the collapse load, which is the load required to cause an incipient failure

along the failure pattern. The soil is treated as a rigid plastic material. The resulting

vertical pressure is identified as the bearing capacity.

36

Figure 2.3 (a) Failure patterns under deep foundations (b) Expanding Cavity in an infinite

soil mass (Vesic, 1972)

(a)

(b)

37

Vesic (1972) extended the cavity expansion theory, originally used in metal

indentation problems (Hill, 1950), to analyze deep penetration in soils. In this method, a

spherical cavity of zero radius is assumed in the soil located near the cone tip. The

pressure around the tip of a cone to cause penetration is the limit pressure required to

expand the cavity (Figure 2.3 b) to a radius equal to the radius of the cone base. The

required pressure for the expansion of the cavity is a function of the shear strength and

compressibility of the soil.

Baligh (1985) developed an analytical technique called the strain path method

(SPM) in an attempt to understand and predict soil behavior during installation of various

rigid bodies (e.g. piles, cone penetrometers, samplers, etc.) into soils. According to

Baligh (1985), the penetration process resembles a steady flow of soil around the

penetrating object rather than an expansion of a cavity in soil. Hence, the soil

deformations due to penetration should be integrated in the analysis. In the SPM, the

penetration process (Figure 2.4) is viewed as a steady flow of soil around a penetrometer.

Soil flow is assumed to occur along streamlines around the rigid body.

In the first step of this method, an initial estimate of the flow field is made using

classical fluid mechanics approach by modeling the soil as an ideal, incompressible and

inviscid fluid. Approximate velocity fields, which satisfy the conservation of volume (or

.

38

Figure 2.4 Deep penetration viewed as a steady flow problem (Baligh, 1985)

39

mass) are then estimated. Velocity fields are then differentiated with respect to the

directional (spatial) coordinates to determine the strain rates. Integration of these strain

rates along the assumed streamlines defines the strain paths (deformation field) for soil

elements around the cone. Once the strain paths of individual soil elements are known,

the second step of this analysis is to use the material constitutive equations to derive the

effective soil stresses. This process is repeated along a number of stream lines to evaluate

the effective stress field around the cone.

The SPM provides a good framework for elucidating and solving penetration

problems that involve large deformations. Solutions using the SPM provide more realistic

predictions than the initially applied bearing capacity or cavity expansion methods to

estimate penetration effects. This method is valuable in geotechnical engineering

applications for predicting the performance of deep (pile) foundations. The estimated soil

strains in the SPM are approximate and independent of the material properties. However,

in realistic situations, material properties will influence the strain fields. As an effect, the

effective stresses computed in the SPM often result in equilibrium errors. These errors

will be small if the assumed strain field is close to the actual one. It is also not clear

whether this analysis can be applied to frictional materials (e.g. sand).

Teh and Houlsby (1991) presented a finite element analysis for undrained

penetration of clay based on the SPM. The strain field from SPM is introduced into the

finite element model as an initial strain condition. The clay is idealized as a

homogeneous, incompressible and elastic-perfectly plastic (or von Mises) material. The

40

corresponding stress field is computed using the finite element model. This analysis

corrects the equilibrium error encountered in the SPM.

Application of the SPM (Baligh, 1985) is restricted to the conditions of steady deep

penetration and cannot predict ground surface deformations. Sagaseta and Whittle (2001)

extended the SPM and called it the shallow strain path method (SSPM). Shallow

penetration causes heave at the ground surface (in the far field), while settlement occurs

in a thin layer adjacent to the shaft (in the near field). In order to treat this problem, the

SPM was modified to SSPM, which includes a stress free ground surface. The soil mass

in both SPM and SSPM is modeled as an ideal semi-infinite fluid that is laterally

unbounded and moves in a uniform flow field along streamlines around the rigid body.

SSPM is used to analyze the deformations and strains caused by shallow undrained

penetration of shafts in clays. SSPM results show a favorable agreement with the field

measurements of building movements caused by installation of large pile groups. The

comparisons show that the SSPM is capable of reliably predicting the deformations

within the soil mass but generally underestimates the vertical heave measured at the

ground surface.

Randolph et al. (2000) examined the resistance of a spherical penetrometer

penetrating in clay. A solution was developed using upper bound and lower bound

approaches, which was supported by a finite element analysis. The clay was assumed to

be a rigid-plastic material obeying either Tresca or von Mises yield criterion. The

assumed velocity field for soil flow provided for an axisymmetric flow condition around

the penetrometer (Figure 2.5). The soil deformation computations were based on small

41

strain formulation for the plastic flow of the soil. The solutions of this analytical

approach compared well with results from finite element analysis. The computed

penetration resistance is 6-10% lower than the upper bound solution, and within 1% of

the lower bound solution.

2.3.3 Experimental study

Allersma (1987) used photo-elastic measuring technique to visualize stresses that

occur during the penetration of a foundation element (e.g. penetrometer, jacked pile) in

granular material. Random-shaped crushed glass was used as a substitute for sand. The

transmission of forces through the crushed glass was determined using a polariscope - an

instrument for ascertaining, measuring, or exhibiting the properties of polarized light. An

automated device to quantify the optically measurable stress was used to determine stress

distribution at the tip of a penetrometer. The disturbance observed at the conical tip of a

penetrometer is depicted in Figure 2.6. The soil mass at the penetrometer tip is subjected

to high stresses and reaches a failure state.

Lead markers placed below the penetrometer tip were used to monitor the

deformation. It was observed that markers close to the tip initially move in the downward

and horizontal directions. However, at later stages of penetration, an upward motion was

observed. The failed soil at the tip of the penetrometer flows upward and adjacent to the

shaft surface during further advancement.

42

Figure 2.5 Plastic soil flow around a spherical penetrometer (Randolph et al. 2000)

43

Figure 2.6 Disturbance caused at the conical tip of a penetrometer (Allersma, 1987)

44

2.3.4 Numerical studies

The problem of soil penetration has been analyzed by using numerical techniques

such as finite element analysis. Small strain or large strain computations have been

employed to model the penetration process. Large strain analysis allows for the

simulation of large deformations that occur in soil penetration problems such as CPT.

Budhu and Wu (1992) presented a large strain analysis using an updated

Lagrangian finite element formulation for understanding the disturbance in soft clays due

to sampler penetration. Soil disturbances due to sampling operations are of major concern

to a geotechnical engineer attempting to estimate in-situ properties of soil by means of

laboratory tests. They studied the effects of stress increase around the samplers due to

penetration. The results of a parametric study to determine the influence on sampling

disturbances due to the rate of penetration, thickness and tip angle of the sample tube are

also presented. The penetration of the sampler is simulated by splitting a group of nodes

ahead of the penetration route and applying incremental displacements so as to match the

geometric configuration of the sampling tube. Thin-layer interface elements were

included to model the frictional interface of varying roughness between the sampler and

the soil. The degree of disturbance for a frictionless sampler was found to be constant

after a penetration depth of 75 % of the sample tube diameter. On the other hand the

degree of disturbance for a frictional sampler keeps increasing as the penetration

advances.

45

van den Berg (1994) presented a more comprehensive large strain analysis of the

CPT in clay and sand using an Eulerian formulation. In large strain analyses, it is

necessary to decide the new location of boundary nodes and redefine the mesh after each

calculation step, making this procedure more complicated (Budhu and Wu, 1992). To

avoid re-meshing in large strain finite element calculations, van den Berg (1994)

uncoupled the nodal displacements and velocities from the material displacements and

velocities. To validate the results of this analysis, laboratory penetration tests were

conducted in homogeneous clays. The effects observed during the tests were similar to

that reproduced by the numerical analysis.

Yu et al. (2000) presented a finite element procedure based on steady-state

deformation of clay to analyze cone penetration in soils. The proposed procedure can be

applied to both clays and sands. In their analysis they focused on an undrained condition

in clays. The total displacements experienced by soil particles at a particular instant in

time during CPT were computed. This method demands less computational time as

compared to the other large strain finite element methods previously described. The

application of this approach is limited to isotropic and homogeneous soil profiles, and is

not suitable for layered deposits.

2.3.5 Summary

The analyses of soil penetration problems range from simple cavity expansion

theory to numerical methods. The primary focus of existing studies has been in

46

computing the stress-strain behavior of soils during penetration. Finite element models

using small and large strain formulations were implemented to model the soil behavior

during rigid body penetration. Most of the solutions were derived from plasticity models

to predict the pre-failure or failure response of soils. Continuous soil penetration is a

steady flow process. The soil at the tip of the penetrating shaft fails and reaches CS. The

soil mass at CS flows near the shaft surface during continuous penetration of the shaft.

2.4 Soil resistance on a penetrating shaft

2.4.1 General

The total resistance on a penetrating shaft in clay is due skin friction on the shaft

surface and the end bearing resistance at the shaft base or tip. Penetration resistance is

usually computed as the failure (or collapse) load, which is the sudden decrease in soil

strength. The interfacial frictional stress on the soil-shaft interface is usually determined

assuming the soil-shaft interface as a failure plane. An effective stress or a total stress

analysis is used to determine the frictional stress. Interfacial frictional stress multiplied by

the shaft area is the total skin friction resistance.

47

2.4.2 Skin friction and end bearing resistance

Effective stress or total stress analyses are widely used in geotechnical

engineering problems. Effective stress analysis (ESA) is used for long-term

considerations where drained conditions prevail (Budhu, 2000). Internal friction angle

( cs′φ ) is the frictional strength parameter for an ESA. The critical failure shear stress ( csτ )

on the shaft surface for an ESA is

cs n cstan( )′ ′τ = σ φ [2.7]

where n′σ is the normal effective stress on the soil-shaft interface.

For short-term or undrained conditions, total stress analysis (TSA) is used

(Budhu, 2000). The undrained shear strength, su, is the strength parameter in TSA. For

shafts penetrating in soft clays, the interface friction stress for a TSA is calculated on the

basis of a reduced undrained shear strength, αsu,, where α is a skin friction factor

obtained from experiments (Tomlinson, 1957). The factor α is a ‘catch-all’ factor that

includes the disturbance region (zone) created around the due to shaft penetration.

However, the effects of the size of this disturbance zone on the penetration resistance of

the shaft have not been investigated.

The end bearing resistance is calculated using either an ESA or a TSA. The end

bearing capacity equations are usually derived using limit equilibrium approach. Limit

equilibrium methods assume a failure mechanism beneath the shaft base treated as a deep

footing (Figure 2.3 a). One or more equilibrium equations can be used to determine the

ultimate limit load required to initiate failure.

48

2.4.3 Viscous (friction) resistance

The soil mass adjacent to the penetrating shaft surface is at critical state. During

continuous penetration of a shaft, the clay adjacent to the shaft surface will flow like a

viscous fluid rather than sliding like a rigid body along the shaft-soil interface. Along

with interfacial friction, referred as static friction for this study, the shaft will be

subjected to an additional viscous resistance due to the post-failure soil flow. According

to Marsland and Quarterman (1982), the relationship between the resistance of the

continuous penetrating rigid body and the rate of penetration depends on the stress-strain

(solid) relationship and the flow (viscous) properties of the soil. Shear viscosity is a

parameter that resists the motion of material particles with respect to each other, and is

analogous to internal friction. For a post-failure soil flow, shear viscosity of soil is

required to determine the viscous resistance.

2.4.4 Summary

Penetration resistance of a shaft in clay is computed as the static collapse load

causing the failure. The soil mass sliding along the soil-shaft interface is treated as a rigid

body. The shear strength parameters such as 'csφ or su are used to estimate the failure

stress and determine the static frictional resistance. During continuous penetration the

shaft is subjected to an additional viscous resistance above the static resistance, which

can be determined by modeling the soil at CS as a viscous fluid.

49

2.5 Viscous behavior of soils

2.5.1 Theory of viscoelastic deformation

Viscous behavior of soils is considered an integral part of soil rheology. The

theory of linear viscoelastic deformation forms the fundamental basis of rheology and

uses a combination of elastic, plastic and viscous properties of the body. Rheological

equations of viscoelasticity connect stress, strain, strain-rate and time. Mechanical

models (e.g. Maxwell and Kelvin) of viscoelasticity are widely used to simulate the

rheological properties of soil. Elastic properties are simulated by a model in the form of

an elastic element, a spring, denoted by the symbol E (Figure 2.7). The shear behavior

complies with Hooke’s law as

Gτ = γ [2.8]

where τ is the shear stress, G is the shear modulus and γ is the (elastic) shear strain,

which is recoverable with the removal of stress.

The model used for viscous bodies, denoted by the symbol V (Figure 2.7), is a

fluid-filled dashpot with a perforated piston moving down the cylinder and obeying

Newton’s law as:

τ = µγ& [2.9]

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

50

Figure 2.7 Mechanical models, E = elastic body, V = Viscous, Newtonian body and P

= plastic body

P

τ

E V

τ τ

G µ

51

Plastic properties are simulated by a dry-friction element, denoted by the symbol,

P (Figure 2.7) and obeying Saint-Venant’s law as

ypτ = τ [2.10]

where ypτ is the stress at which the friction slider begins to slide, inducing plastic strains

in the body.

The combination of two or three elements (E, V and P) described above is used to

model the viscoelastic behavior. For example, a Maxwell body can be represented by

linking an elastic element in series with a viscous element (Figure 2.8 a). The Kelvin

body consists of an elastic element connected in parallel with a viscous element (Figure

2.8 b). These simple models are very popular and have the capacity to demonstrate the

properties of a material visually. These models are used to study material responses such

as creep.

2.5.2 Creep deformation

According to the classical theories of elasticity and plasticity, the magnitude of

stress is defined by the magnitude of applied load and how it is applied. If the applied

load remains unchanged, the resulting stresses and strain also remains unchanged. In real

bodies, the stress-strain behavior is observed to change with time. Creep is a long-term

deformation occurring under a constant external load, resulting from changes in the state

of stress and strain of a body as a function of time.

52

Figure 2.8 Models of viscoelastic bodies (a) Maxwell body (b) Kelvin Body

τ

G

µ

τ

G µ

(a) (b)

53

Total strain, γ , in a body is written as:

0 (t)γ = γ + γ [2.11]

where 0γ is the strain induced immediately (or in a very short interval of time) after the

application of load and (t)γ is the strain developing with time without change in the

magnitude of applied load. The rate of strain, ddtγ

γ =& , is observed to decrease (tends to

zero) with time as illustrated in Figure 2.9. The strain (t)γ attains a constant finite value

at large times.

The creep of soils below foundations has led to total and differential settlements

of structures, instability of slopes and tilting of retaining walls, causing considerable

economic losses. The creep behavior of soils has been extensively studied and is

elaborated in many articles and books (e.g. Whitman, 1957; Yong and Japp, 1967;

Mitchell, 1976; Vyalov, 1986; Desai, 2001). Still, there is no clear understanding on the

mechanics of creep in soils.

Whitman (1957) and Yong and Japp (1967) investigated the effects of the rate of

loading on compressive strength of sands and cohesive soils. They investigated the creep

behavior of soils by modeling it as a viscous deformation occurring at slow rate. They

concluded that soil behavior can be modeled like a viscous fluid.

Vyalov (1986) described the rhelogical behavior of soils with respect to the states

of stresses and strains. Viscoelastic models such as Maxwell and Kelvin are employed to

study time-dependent deformation. Vyalov (1986) stated that the qualitative aspect of

.

54

Figure 2.9 Creep and viscous flow deformation

t

γ

(t)γ

2

1 1- viscous flow 2- creep deformation

55

creep behavior can be represented by these models.

2.5.3 Viscous behavior at critical state

The viscous behavior of soils, discussed in the preceding sections, is associated

with creep - a slow time-dependent deformation. Creep response consists of shear and

volume changes (volumetric strains), which occur at a slow rate. The shear strains in

creep occur before failure and usually prevail in the pre-failure state.

The soil around a penetrating shaft is at CS and will flow adjacent to shaft surface

at constant volume during its continuous motion. The post-failure soil (flow) response is

similar to flow of a viscous fluid, hereafter called viscous flow. The term “plastic flow” is

used in the theory of plasticity. However, it denotes a development of plastic deformation

when the load reaches a certain limit (yield point). Plastic flow of soil is assumed after

yielding and up to failure.

In viscous fluids, application of external shear stresses induces a viscous flow

progressing at a certain velocity of finite magnitude. The stress in a viscous flow is

proportional to the velocity of flow (or the rate of change of deformation). A fluid in

which the stress is directly proportional to the rate of flow is called a Newtonian perfectly

viscous fluid. The magnitude of external shear stress to initiate viscous flow depends on

the type of fluid. For a Newtonian fluid, a shear stress greater than zero induces viscous

flow. In some fluids, the flow is initiated only when the applied shear stress exceeds

certain value, called yield stress. Such fluids are classified as yield stress fluids. Soils at

56

CS are assumed to be similar to yield stress fluids, with CS shear strength analogous to

the yield stress. A soil at CS will flow for applied shear stress greater than the CS shear

stress. CS shear strength is the yield stress required to initiate the flow of a soil at CS.

The behavior of soil as a yield stress fluid is discussed in next chapter.

Strains occurring in a viscous flow are irrecoverable. The manner in which strains

develops with time for a viscous flow is depicted in Figure 2.9. The deformation in a

viscous flow progresses at a constant rate and is characterized by straight line labeled ‘1’.

Creep behavior modeled by a combination of elastic, plastic and viscous behavior is

characterized by curve labeled ‘2’.

The post-failure flow of soil at CS is similar to viscous flow as characterized by

straight line ‘1’ (Figure 2.9). A continuous viscous flow denotes an unceasing and

unconfined change in shape. Typical in this respect is the flow of a perfectly viscous

Newtonian) fluid. Post-failure viscous flow of soil at CS can be thought of a special case

of creep, where flow is similar to that of a purely viscous liquid with no recoverable

deformation.

2.5.4 Summary

Existing studies on viscous behavior of soils is an idealization of solid body as a

plastic fluid to understand the pre-failure plastic flow response after (plastic) yielding and

prior to failure. For applied shear stresses greater than the CS shear stress, post-failure

response of a soil at CS is like viscous flow. This flow response can be modeled as pure

57

viscous fluid rather than a combination of different characteristics such as elastic, plastic

and viscous used to represent creep behavior.

2.6 Viscous shear force

2.6.1 Shear viscosity

Viscosity is a property of liquids (and gases) to resist the motion of elemental

particles with respect to one another. Shear viscosity is associated with internal friction

between two layers of liquid moving relative to each other. Viscous flow offers viscous

resistance due to shear viscosity. Newton (1687) was the first to investigate viscosity. He

found that the shear resistance in a flowing fluid resulted from internal slippage of

particles.

Viscosity is the resistance to distortion or internal friction (Lamb, 1932) that is

exhibited by all real fluids. In viscous fluids, the distortion depends on the rate of change

of shape while in solids the distortion depends on actual changes in the shape. In solids,

the resistance to distortion is termed shearing resistance. Despite this difference the

mathematical methods to describe distortion in both viscous fluid and solid are almost

indistinguishable. For example, the stresses on an infinitesimal element of viscous fluid

are (Lamb, 1932):

xx2 u v w up 23 x y z x

⎛ ⎞∂ ∂ ∂ ∂σ = − − µ + + + µ⎜ ⎟∂ ∂ ∂ ∂⎝ ⎠

[2.12]

58

zz2 u v w up 23 x y z z

⎛ ⎞∂ ∂ ∂ ∂σ = − − µ + + + µ⎜ ⎟∂ ∂ ∂ ∂⎝ ⎠

[2.13]

yzw v y z

⎛ ⎞∂ ∂τ = µ +⎜ ⎟∂ ∂⎝ ⎠

[2.14]

zxu w z x∂ ∂⎛ ⎞τ = µ +⎜ ⎟∂ ∂⎝ ⎠

[2.15]

xyv u x y

⎛ ⎞∂ ∂τ = µ +⎜ ⎟∂ ∂⎝ ⎠

[2.16]

where p is the ambient fluid pressure at rest, σ is the normal stress, τ is the shear stress, µ

is shear viscosity and u, v, and w are the velocities in the x, y, and z Cartesian directions

respectively. The subscripts refer to the planes on which the stresses act. The dimension

of µ is M L-1T–1 where M is mass, T is time and L is length. Equations [2.12] to [2.16] are

similar to the generalized stress-states in a (three-dimensional) solid body stated in most

texts on solid mechanics (e.g. Fung and Tong, 2001).

2.6.2 Viscous drag on shafts in clays

During continuous motion, the shaft surface (e.g. cone friction sleeve) will be

subjected to a viscous drag (resistance) due to the viscous flow of clay adjacent to the

shaft surface. Studies related to the dynamic penetration of clays (Turnage, 1973; Murff

and Coyle, 1973; Berry, 1988) show that viscous resistance is an important component of

59

the total resistance offered by the soil. No analysis is currently available to study or

determine the viscous drag component of the total resistance offered by the soil.

2.6.3 Creeping flow

The main aim of this study is to determine the viscous drag on a shaft penetrating

a clay. The clay flowing around the shaft is assumed as a slow viscous flow at low

Reynolds number [µ

ρVLR e = << 1, eR is the Reynolds number, ρ is density, V is the

velocity and L is the characteristic length). Such flow is called creeping flow (Happel and

Brenner, 1965). Creeping flow involves fluids of high viscosities at slow velocities,

resulting in a low Reynolds number. Soil at CS as a viscous fluid can be presumed to

flow at a low Reynolds number. Materials that exhibit creeping flow behavior include

asphalt (bitumen) at low temperatures, tar, molasses, molten lava, thick slurries and gel.

The following Section 2.6.4 includes a brief background to determine viscous drag on

bodies in creeping flow.

2.6.4 Viscous drag on bodies in creeping flow

In creeping flow, viscous forces resulting due to shearing flow predominate over

inertial forces. McNown et al. (1948) showed that inertia is important only if Re > 70.

The Reynolds number for the creeping flow considered here can be perceived to be much

60

below 70. Neglecting inertia and assuming that any conservative extraneous volume

forces are included in the pressure term, p, the two governing equations that apply to

creeping flow are the Navier-Stokes equation for incompressible fluids (constant volume)

given by:

2v p1∇ = ∇

µ [2.17]

and the continuity equation:

v 0∇⋅ = [2.18]

where ∇ is the divergence operator, v is the local mass average fluid velocity and µ is the

shearing viscosity referred as viscosity in this study. The solutions of Equations [2.17]

and [2.18] using appropriate boundary conditions are used to determine the velocity

distribution and drag on a body in a viscous flow field.

Consider a rigid body moving in a viscous fluid with flow around the body as a

creeping flow. The equation to compute viscous drag on this body is known to be of the

form given by (Lamb, 1932; Ray, 1936; Happel and Brenner, 1965; Panton, 1984):

z zf CµV= [2.19]

where zf is the viscous drag force in the z (vertical) direction, zV is vertical velocity of

the body and C is a constant, which is governed by the geometry of the body with respect

to the flow field and imposed boundary conditions. For an unbounded creeping flow past

a sphere (Figure 2.10 a) of radius, R, C 6πR= (e.g. Lamb, 1932; Panton, 1984). Ray

.

61

Figure 2.10 Creeping Flow Past (a) a Sphere (b) a circular disk (Van Dyke, 1982)

(a)

(b)

62

(1936) obtained solutions for the motion of circular disk (Figure 2.10 b) in an unbounded

viscous fluid. For a disk of radius R the derived value ofC 16R= .

2.6.5 Effect of boundaries

The drag in the presence of finite boundaries or bounded flow conditions is higher

than the drag obtained for semi-infinite unbounded flow conditions (e.g. McNown et al.,

1948). The viscous soil around a penetrating shaft is bounded by finite boundaries

(Figure 1.2 b) beyond which the soil exists at pre-failure states or relatively undisturbed.

Drag on a penetrating body (e.g. shaft) should be analyzed considering the boundary

conditions. The extent of the failure zone around the shaft should be understood to

incorporate the effects of boundary conditions on the viscous drag.

2.7 Soil state around the shaft

When a shaft penetrates into a soil, the soil in its path fails and is displaced

outwards during its advancement. A region of soil near the shaft called the influence zone

is disturbed. The influence zone classified into different sub zones according to the

created intensity of disturbance.

Zeevart (1948) describes three soil zones around driven piles, referred as shaft in

this study. The study was based on the observations of shaft driving (penetration) in the

subsoil of Mexico City (Figure 2.11). Zone I is an annulus of soil that is subjected to

63

Figure 2.11 Soil Disturbance around Penetrating Shafts in Soft Clay (Zeevart, 1948)

64

excessive disturbance (remolding). The soil in this zone reaches CS and is at a post-

failure state. This annulus of soil is referred as the CS zone for this study. If no soil is

squeezed out to the ground surface, the volume of the CS zone is at least equal to the soil

volume displaced by the occupied shaft volume. The extent of this zone according to

Zeevart (1948) is 1.4 times the shaft radius measured from the center of the shaft (Figure

2.11). The soil in this zone flows during the continuous penetration of the shaft.

Zone II is a disturbed zone with a lesser degree of disturbance. The soil in this

zone is usually at a pre-failure state. Soil movement of a point in this zone only occurs

when the shaft tip is along the same depth level and adjacent to it. The movement ceases

when the shaft tip advances further below. This zone, according to Zeevart (1948), has a

radius of about 3 times the radius of the shaft. Zone III, which is present at a distance

beyond zone II, is relatively undisturbed.

Cummins et al. (1948) conducted field and laboratory tests to determine the extent

of disturbance produced by driving shafts in soft clay deposits. They observed that close

to the shaft surface, the natural structure of the clay is excessively disturbed and the

disturbance vanishes at a distance of about 4 times the shaft radius from the center of the

shaft.

Esrig et al. (1977) and Kirby and Wroth (1977) investigated the axial capacity of

driven shafts in clay. Based on displaced volume considerations they suggested that for

driven shafts, the zone of completely remolded soil extends to at least 1.4 times the shaft

radius, measured from the center of the shaft. Kirby and Wroth (1977) stated that the

65

shear property of the soil in this remolded zone is important to understand the shear

resistance on the shaft.

From the observations on driven timber shafts and information available in the

literature, Flaate (1972) concluded that shaft penetration causes strong remolding in a

zone up to 10 to 15 cm from the shaft surface (equivalent to a CS zone radius of 2 to 2.5

times the radius of the shaft studied). Cooke and Price (1973) conducted load tests on

instrumented jacked piles (shafts) in London Clay and observed that movements around

the shaft are greatest within about 4 pile radius of the shaft (possible CS zone).

Randolph and Wroth (1978) discussed the size of influence zone around a driven

shaft by introducing a radius (ri). This radius is the distance from the center of the shaft to

a point away from shaft where the induced shear stress due to penetration becomes

negligible. An approximate value of the radius of influence zone can be computed as:

υ)2.5L(1ri −= [2.20]

where, L is the length of the shaft and υ is the Poisson’s ratio of the soil.

Randolph et al. (1979) assumed that shaft installation (penetration) can be

modeled as the undrained expansion of a cylindrical cavity under plane strain conditions.

It was shown that the radius of the CS zone in case normally consolidated soil is about 5

times the shaft radius from the center of the shaft. For an overconsolidated soil (over

consolidation ratio, OCR = 8) the radius of the CS zone is about 4 times the shaft radius

from the center of the shaft. They assumed that the soil behaves according to the

Modified Cam Clay model and established that, irrespective of consolidation history, the

66

soil adjacent to driven piles ends up as normally consolidated. This is consistent with CS

soil mechanics and the Modified Cam Clay model that embodies it.

From the above cited literature, CS zone exists near the surface of the shafts. The

radius (extent) of the CS zone ranges between 1.4 to 5 times the shaft radius from the

center of the shaft. The minimum value of about 1.4 times the shaft radius is based on

volume considerations (i.e. the zone equivalent to the shaft volume). The effects of the

size of this zone on the viscous resistance on the penetrating shaft are investigated in this

study.

2.8 Summary

Total resistance on a shaft during continuous penetration is due to static and

viscous components. The static component of resistance is studied extensively. However,

the viscous drag is not studied and is still unclear.

It appears from the literature that a CS zone exists near the penetrating shaft

surface. The soil in the CS zone is displaced and flows during the advancement of the

shaft. The penetrating shaft in clay will experience viscous drag due to the soil flow.

Soils at CS can be assumed to be similar to yield stress fluids, which will flow on

exceeding the CS shear stress. Creeping flow equations can be used to determine the

viscous drag on a penetrating shaft.

67

CHAPTER 3

MATHEMATICAL FORMULATION AND ANALYSIS

3.1 Introduction

The penetration of rigid objects such as shafts into soils creates a zone of

disturbance soil around them. The soil in this zone is remolded and has reached critical

state (CS). Soils at CS can be modeled as a viscous yield stress fluids and will flow if

stresses greater than the CS shear (fluid yield) stress are applied. In this chapter a new

theoretical approach is presented using creeping flow hydrodynamics to analyze the

viscous drag on a cylindrical shaft penetrating a clay.

3.2 Post-failure response of soil as yield stress fluid

Viscoplastic or yield stress fluids that do not flow unless they are subjected to a

certain stress are widely used in industrial applications. Well known examples are

toothpaste, gels, clay suspensions, soft glassy materials, fresh concrete and drilling fluids.

For example, when a tube of toothpaste is opened an adequate force is needed before the

toothpaste will start flowing. This force is called the critical force and the induced stress

to initiate flow is called the critical stress or yield stress.

When the applied stress in a yield stress fluid is less than the fluid yield stress, it

deforms plastically like a solid with definite strain recovery upon the removal of stress.

Applied stress exceeding the critical yield stress will cause the fluid to flow and

68

accelerate, leading to avalanches similar to those observed in granular materials (Coussot

et al., 2002). The flow behaviour yield stress fluids can be modeled using Newtonian

(e.g. Bingham) or non-Newtonian flow models.

For simple shear flow, the constitutive equations for yield stress fluids are

expressed as (Nguyen and Boger, 1992):

y0 (noflow)γ = τ <τ& [3.1]

y p yτ = τ +µ γ τ ≥τ& (Bingham fluid) [3.2]

where τ is the total (dynamic) shear stress, yτ is the yield stress, pµ is the shear viscosity

and γ& is the shear strain rate. The CS shear stress ( csτ ) is representative of the fluid yield

stress ( yτ ). When pµ is constant (Newtonian), the flow can be modeled as a Bingham

fluid represented by curve A (Figure 3.1). The constant viscosity is an ideal case for

many visco-plastic fluids at high shear stresses, when there is complete breakdown and

disruption of the material structure that is responsible for yield behavior (Nguyen and

Boger, 1992).

The yield stress definition in plasticity models (e.g. CSM) is not the same as the

fluid yield stress. In solid mechanics, yield stress is the stress at the onset of permanent

strains with some strain recovery after the removal of the applied stress. The yield stress

in a yield stress fluid is the stress at the onset of viscous flow with no strain recovery. In

this study, the yield stress refers to the fluid yield stress.

69

Figure 3.1 Typical flow curves for yield stress fluids.

Shear rate, γ&

Shea

r stre

ss,τ

Yield Stress

B

A

Non-Newtonian

Bingham (Newtonian)

y0 (noflow)γ = τ <τ&

y p yτ = τ +µ γ τ ≥τ& (Bingham fluid)

70

A soil at CS is a yield stress fluid with the yield stress equal to the critical state

shear stress. An applied external shear stress less than the CS shear stress will cause

elasto-plastic deformation of a soil. An external shear stress equal to CS shear stress will

cause a steady deformation of soil at constant volume. The soil will flow like a viscous

fluid for an applied shear stress greater than the critical shear stress. Viscous drag stress

will then act above the static shear stress. Combining soil behavior before critical state

(Figure 1.1 a) and the soil behavior of soil at critical state as yield stress fluid (Figure 3.1)

the post-failure (flow) response of clays can be represented as illustrated in Figure 3.2.

3.3 Analytical method

3.3.1 General

The current methods of calculating the resistance of shafts penetrating in soils are

based on the assumption that the soil is either a rigid or a deformable solid. These methods

allow for the determination of the static component of the total resistance. For a shaft under

continuous motion in a failed soil mass, viscous resistance is present. No analysis is

available to the knowledge of the author to determine the viscous drag on a penetrating

shaft. The analytical method to determine the viscous resistance to be described later is

based on creeping flow and the analysis requires certain assumptions listed in the next

section.

71

Figure 3.2 Analogy of the post-failure response of a soil to a yield stress fluid response.

Dilative soil

τ

(b)

Critical shear stress

τp

τcs

γ

Peak

Non-dilative soil

(a)

γ

Dilative soil

Non-dilative soil

e

Shear strain

Shea

r stre

ss

ecs

Critical void ratio

P

C

C

τd

(c)

Shea

r stre

ss

γ&

y csτ = τ d cs; 0τ = τ γ =&

Shear (flow) rate

Post-failure

72

3.3.2 Assumptions

The following assumptions are made to develop the analytical method for this

study.

1. The CS zone adjacent to the shaft exists in a concentric annular region.

2. The soil flow occurs at constant volume, consistent with critical state soil

mechanics.

3. The soil mass in the CS zone exhibits steady creeping viscous flow when

subjected to stresses higher than CS shear stress of the soil.

4. A unit length of the shaft surface is analyzed. The CS zone size along the

unit length is constant.

5. Shear viscosity of the soil in the CS zone around the shaft is constant.

3.3.3 Viscous drag on a penetrating shaft in clay

Consider the axial motion of a cylindrical shaft of radius, 0r , in a semi-infinite soil

mass for which an annulus of soil of radius, 0R , adjacent to the shaft has reached critical

state (CS zone) as illustrated in Figure 3.3. The viscous soil within the CS zone is

laterally bounded by soil in a pre-failure state (solid boundary). The total shearing side

resistance ( τ ) on the shaft surface moving at a constant rate is due to static and viscous

components, i.e.

cs vτ = τ + τ [3.3]

73

Figure 3.3 Forces acting on a soil element in the CS zone around the shaft

Undisturbed soil

Elevation

Influence zone

Critical state (CS) zone

Plan

δr

r R0

R0

z

δz

Shaft

r

r0

x

r

p

zdzdpp δ⎟

⎠⎞

⎜⎝⎛+

Stresses acting on the element

cs p pdu d du+ µ µ rdr dr dr

⎡ ⎤⎛ ⎞τ + δ⎜ ⎟⎢ ⎥⎝ ⎠⎣ ⎦ δz

δr

cs pdu+µdr

τ

74

where and vτ is the viscous drag component. The CS shear stress is constant (static) and is

independent of the penetration rate. However, the viscous component will be influenced by

the penetration rate (velocity). For a shaft at rest (penetration rate = 0), the viscous resistance

will not act and the total resistance will be equal to the static component.

Consider a cylindrical element within the CS zone is (1) everywhere parallel to z

(vertical direction), i.e. du =0dz

and (2) a function of r, ( 0 0r r R≤ ≤ ).the drag shear stress on

the shaft surface according to Equation (3.3) is

cs pdudr

⎛ ⎞τ = τ +µ ⎜ ⎟⎝ ⎠

[3.4]

where pdudr

⎛ ⎞µ ⎜ ⎟⎝ ⎠

is the viscous drag component of the shearing resistance and pµ is the

(shear) viscosity of the soil.

The tangential stress on the vertical planes of the cylindrical element at r is

cs pdudr

⎛ ⎞τ = τ +µ ⎜ ⎟⎝ ⎠

and cs p pdu d du+ µ µ rdr dr dr

⎡ ⎤⎛ ⎞τ + δ⎜ ⎟⎢ ⎥⎝ ⎠⎣ ⎦at r r+ δ .

The difference in tangential tractions on the two curved surfaces (Figure 3.3) is

pd dudT= µ 2πrδz δrdr dr⎛ ⎞⎜ ⎟⎝ ⎠

[3.5]

The difference in normal pressures on the ends of the cylindrical element is dp zdz

δ . The

traction from this pressure difference is, dpdP 2 r z rdz

= π δ δ . For vertical equilibrium of the

75

element, the tractions due to the pressure difference must balance the tangential tractions,

i.e. dT = dP

Therefore,

pd du dpdT= µ 2πrδz δr= 2πr.δz.δrdr dr dz⎛ ⎞⎜ ⎟⎝ ⎠

[3.6]

which simplifies to

p

1 d du 1 dprr dr dr µ dz

⎛ ⎞ =⎜ ⎟⎝ ⎠

[3.7]

The solution of Equation (3.7) is

2

1 2p

r dpu= +C lnr+C4µ dz

[3.8]

Applying velocity boundary conditions, zu = V at 0r r= (the shaft radius) and u = 0 at

0r R= , we get

2 20 0

1 zp 0

0

R r dp 1C V4µ dz Rln

r

⎡ ⎤⎛ ⎞−= − +⎢ ⎥⎜ ⎟⎜ ⎟ ⎛ ⎞⎢ ⎥⎝ ⎠⎣ ⎦ ⎜ ⎟

⎝ ⎠

[3.9]

and,

2 2 20 0 0 0

2 zp p 0

0

R R r ln Rdp dpC V4µ dz 4µ dz Rln

r

⎡ ⎤⎛ ⎞−= − + +⎢ ⎥⎜ ⎟⎜ ⎟ ⎛ ⎞⎢ ⎥⎝ ⎠⎣ ⎦ ⎜ ⎟

⎝ ⎠

[3.10]

76

The net flow in the CS zone is zero, i.e. the net flow across a horizontal cross-section at any

depth within the CS zone is zero (constant volume condition consistent with CS). Thus the

volume flow boundary condition is

0

0

R

r

Q 2 urdr 0= π =∫ [3.11]

Substituting Equation (3.8) in Equation (3.11) and solving for dpdz

, we get

2 2 2 2p z 0 0 0 0 0 0

2 2 2 2 2 20 0 0 0 0 0 0 0 0 0

4µ V ( r R 2r ln r 2r ln R )dpdz (r R ){2r ln r 2r ln R (r R )(ln R ln r 1)}

− + + −=

− − + − − − [3.12]

Substituting Equation (3.9) and Equation (3.10) into Equation (3.8) we get

2 2 220 0 0

zp p p0

0

2 20 0 0

zp 0

0

R r Rr dp dp ln r dpu V4µ dz 4µ dz 4µ dzRln

r

R r ln RdpV4µ dz Rln

r

⎡ ⎤⎛ ⎞−= − + −⎢ ⎥⎜ ⎟⎜ ⎟ ⎛ ⎞⎢ ⎥⎝ ⎠⎣ ⎦ ⎜ ⎟

⎝ ⎠⎡ ⎤⎛ ⎞−

+ +⎢ ⎥⎜ ⎟⎜ ⎟ ⎛ ⎞⎢ ⎥⎝ ⎠⎣ ⎦ ⎜ ⎟⎝ ⎠

[3.13]

Substituting Equation (3.12) in Equation (3.13) and simplifying we get

{ }2 2 2 4 4 2 2 2 2 2 2 2 2z 0 0 0 0 0 0 0 0 0 0 0 0

2 2 2 2 2 20 0 0 0 0 0 0 0 0 0

V 2r (r R )lnr (r R 2r r )lnR (r R )(r R (r R )lnr)u

(r R ){2r lnr 2r lnR (r R )(lnR lnr 1)}− + + − − − − + −

=− − + − − −

[3.14]

The velocity profile in the CS zone as given by Equation (3.14) is shown in the

Figure 3.4. The soil adjacent to the shaft is dragged downwards along with it. In the far

field (after point B), the soil moves upwards (heave) as shown in the Figure 3.4. At the

shaft surface, the velocity is Vz, and at the boundary of the CS zone, the velocity is zero,

77

Figure 3.4 Velocity profile of the viscous soil in the CS zone

r

Vz

Vz

A D

r0 Ro

R1

R2

BC

Upward Movement

u

u= 0

CS Zone

Shaft

u = Vz

78

which is consistent with the imposed boundary conditions. However, the net flow across

a horizontal plane at any depth (e.g. plane ABD) is zero as given by Equation (3.11), i.e.,

01

10

RR

Rr

Q 2π urdr urdr 0⎡ ⎤

= ⎢ + ⎥ =⎢ ⎥⎣ ⎦∫ ∫ [3.15]

where R1 is the radial distance of the point of inflexion of the velocity profile (point B in

Figure 3.4)

The viscous drag stress experienced by the penetration shaft is

0

v pr r

duµdr =

⎡ ⎤τ = ⎢ ⎥⎣ ⎦ [3.16]

which, after substituting of the differential of Equation (3.8), gives

( )3 2 4 3 30 0 0 0 0 0 0 0 0

v p z 2 2 2 2 2 20 0 0 0 0 0 0 0 0 0

3r 4r R R / r 4r ln r 4r ln Rµ V

(r R ){2r ln r 2r ln R (r R )(ln R ln r 1)}− + − +

τ =− − + − − −

[3.17]

The viscous drag force (fv) per unit length of the shaft is v 0 vf 2πr= τ , [3.18]

which on substituting Equation (3.17) gives

( )3 2 4 3 30 0 0 0 0 0 0 0 0

v 0 p z 2 2 2 2 2 20 0 0 0 0 0 0 0 0 0

3r 4r R R / r 4r ln r 4r ln Rf 2πr µ V

(r R ){2r ln r 2r ln R (r R )(ln R ln r 1)}− + − +

=− − + − − −

[3.19]

Equation (3.19) is simplified as

4 2v 0 0 0

2 2p z 0 0 0 0

f 4 4 ln 32πµ V (1 ) (1 )(ln 1) 2 ln

λ − λ + λ +=

⎡ ⎤−λ −λ λ − − λ⎣ ⎦ [3.20]

where 00

0

Rλr

=

Putting 4 20 0 0

0 2 20 0 0 0

4 4ln 3β(1 ) (1 )(ln 1) 2ln

λ − λ + λ +=

⎡ ⎤−λ −λ λ − − λ⎣ ⎦, [3.21]

79

we get,

v p z 0f 2πµ V β= [3.22]

Equation (3.22) reveals that the soil’s viscosity, rate of shaft penetration and the

size of the CS zone influence the viscous resistance on the penetrating shaft. The term 0β

or 0λ quantifies the effects of size of the CS zone on the viscous drag.

The total resisting force per unit length of shaft is

z 0 cs p z 0f 2πr 2πµ V β= τ + [3.23]

or ( )p z 0z 0 cs 0 cs v

0

µ V βf 2πr 2πr

r⎛ ⎞

= τ + = τ + τ⎜ ⎟⎝ ⎠

[3.24]

where,

p z 0v

0

µ V βr

τ = [3.25]

is the viscous drag stress.

The total soil resistance on the shaft is

n

z zi ii 1

F = f l=∑ [3.26]

where i indicate the length segment of the shaft in clay and n is the total number of length

segments. If the static and viscous components are constant throughout the length of the

shaft the total resistance on the shaft is

( )n

z zi i 0 cs vi 1

F = f l 2πr L=

= τ + τ∑ [3.27]

where L is the embedment length of the shaft in the clay.

80

3.4 Results and discussion of analysis

3.4.1 Parameters influencing viscous drag

According to the theoretical analysis, size of the CS zone, shear viscosity of the soil

and the rate of penetration (velocity profile) influence the viscous resistance on a

penetrating shaft. For a given shaft radius, the viscous drag stress depends on the

product p z 0µ V β . The effects of these parameters on the viscous drag component and their

relationship with other parameters are discussed.

3.4.2 Effects of the size of CS zone

The size of the CS zone is quantified by 0β or 0λ . Higher values of 0λ result in

lower values of 0β . Consider a shaft penetrating at constant velocity. For values of 0λ less

than about 4, 0β changes rapidly and the viscous resistance increases significantly as

0λ decreases (Figure 3.5). For values of 4λ 0 > , 0β remains virtually constant indicating

that the CS zone with radius greater than about 4 times the radius of the shaft does not

significantly influence the viscous resistance for a given penetration rate. In a study of

shaft installation effects, Randolph et al. (1979) showed that 0λ is about 5 for a normally

consolidated soil and 0λ = 4 for an overconsolidated soil (OCR = 8).

81

Figure 3.5 Effect of size of the CS zone on the viscous drag on a cylindrical shaft

1 1.4 2 3 4 5 6 70

2

4

6

8

10

12

rRo

Pile

Viscous soil (CS zone)

o0

0

Rr

λ =

vo

z

f β2πµV

=

82

The minimum expected value of 0λ is about 1.4 (e.g. Zeevart, 1948; Kirby and

Wroth, 1977; Esrig et al., 1977). Using Equation (3.21), this corresponds to a value of

0β 10= .07 and a viscous shaft resistance per unit length of

v p z 0 p z p zf 2πµ V β 20.14πµ V 20πµ V= = ≅ [3.28]

Equation (3.28) may be regarded as an upper limit of the viscous drag on the shaft.

The theoretical lower limiting drag occurs at larger values of R0 ( 0 0R r>> ). For the lower

limiting drag case, Equation (3.19) reduces to

p zv

0

2πµ Vf

ln 1=

λ − [3.29]

This solution can be used to compute the limiting viscous drag on a penetrating shaft. 3.4.3 Velocity profile within CS zone

The viscous drag stress on the shaft is dependent on the velocity gradient (or strain

rate) at the shaft soil interface (Equation 3.16). The velocity of the shaft and the size of

CS zone affect the velocity profile in the CS zone. The velocity distributions on a

horizontal plane within the CS zone for different sizes of CS zone are shown in Figure

3.6. For small sizes of CS zone, the velocity gradient at the shaft soil interface

significantly influences the viscous drag on the shaft. The rate of change of velocity

gradient at the shaft soil interface decreases significantly with increasing sizes of critical

state zone and for large sizes of the CS zone, the velocity gradient at the soil-shaft

interface reaches an approximately constant value.

83

The velocity distributions shown in Figure 3.6 satisfy the zero volume flow

boundary condition imposed by Equation (3.11). The soil in the near field of the shaft is

dragged downwards, while in the far field the soil movement (velocity) is upwards

(heaving). Both, the radial distance from the center of the shaft at which the upward

movement is initiated (R1 in Figure 3.4) and the position of maximum upward motion (R2

in Figure 3.4), increase with 0λ as shown in Figure 3.7. As the size of the CS zone

decreases, the magnitude of the maximum upward velocity increases.

3.4.4 Shear viscosity of clay

The viscous drag on the penetration shaft is directly proportional to the product

p z 0µ V β (Equation 3.22). The shear viscosity of the clay in the CS zone influences the

viscous drag on the shaft surface. For 6λ o > , the product p z 0µ V β (Figure 3.5) is nearly

constant and, consequently, the viscous drag on the shaft does not change significantly.

For computing the viscous drag, shear viscosity of the soil is essential. However, the

available experimental procedures in the literature are for soils with higher water contents

and are not suitable for the soils with low water contents.

84

Figure 3.6 Velocity profiles for different sizes of the CS zone

6.0 4.0 2.0

λo=1.4

o0

0

Rr

λ =

z

uV

85

Figure 3.7 Upward movement of soil in the CS zone

Maximum upward movement

Initiation of upward movement

(b)

0

rr

86

3.5 Conclusion

The theoretical analysis of viscous drag on a shaft penetrating in clay shows that

the viscous drag depends on the size of the CS zone around the shaft, the shear viscosity

of the clay in this zone and the penetration velocity of the shaft. The viscous soil mass

close to the shaft surface is dragged downwards along with the shaft. In the far field, the

soil moves upwards (heave). The key soil parameter to estimate the viscous drag in post-

failure analysis is the soil’s shear viscosity.

87

CHAPTER 4

SHEAR VISCOSITY OF CLAYS AT CS

4.1 Introduction

Limited information is available in the literature on viscosities of soil. The

available investigations are applicable to soil flow events such as landslides, mudslides or

earth flows. The soils in such flows contain water contents higher than their liquid limits.

Viscosity of soils with high water contents is measured using viscometers, which are

difficult to operate for soils with low water contents. However, soils can reach CS at

water contents less than liquid limit and flow. The shear viscosity of these soils cannot be

measured using viscometers.

The influences of strain rate on shear behavior of clays have been studied using

shear tests. In these studies, viscous effects are observed in terms of increasing shear

stress with increasing shear strain rate. However, the shear viscosity of clay at CS has not

been specifically addressed. This chapter focuses on the use of the fall cone test to

determine shear viscosity of clays with water contents less than or near the liquid limit.

The fall cone test is widely used to interpret the index properties and the shear

strength of fine grained soils. In this research, the existing theory of the fall cone test is

extended to determine the shear viscosity of clays. It is shown in this chapter that this can

be achieved by a continuous record of the penetration depth-time data in a fall cone test.

88

4.2 Current Investigations of soils viscosity

4.2.1 General

Torrance (1987) stated that it is of geotechnical importance to know the flow

properties of soil such as viscosity for understanding post-failure response. Most

viscosity measurements for soils are done in connection with clay slurry flows, earth

(debris) flows, mudslides or landslides. The soils in these events have water contents

higher than their liquid limit. In a literature search by the author, disparate information on

viscosities of soil is available.

4.2.2 Landslides and earth flows

The solid structure of clay consists of mineral particles, glued together by ionic

bonds. These bonds are broken by ion exchange with or in the absence of external

loading. The exchange can be such that the stronger bonds are substituted by weaker

bonds. With weaker bonds water can now easily cause the soil structure to collapse and

bring the soil mass to a fluid state. Clays often form colloidal suspensions when

immersed in water. Due to this property of clay, landslides and earth flows frequently

occur in regions with heavy rainfall or in clay deposits with high water contents (e.g.

submarine clays). Sensitive clays have large proportions of water present between their

solid particles, which result in high in situ water contents, remarkably higher than their

89

liquid limit. Such clays suffer a large loss of its strength when disturbed leading to rapid

earth flows and landslides. Deposits of sensitive clays normally flow (Figure 4.1) when

disturbed by external influences such as earthquakes or heavy rainfall.

Soil behavior when flow is initiated cannot be predicted using the parameters

measured by conventional shear tests (e.g. direct shear test, triaxial shear test). To

understand the post-failure flow behavior the rheological properties, particularly viscosity

is essential. According to Edgers and Kalrsrud (1982), the shear viscosity of soil plays an

important role in landslides and earth flows. Researchers to study the soil flows in these

events, have examined the viscosities of sensitive clays, which have high water contents

and low shear strength.

Conventionally, viscometers have been used for measuring the shear strength

(Torrance, 1987; Tan et al., 1991; Fakher et al., 1999) and shear viscosity (Komamura

and Huang; 1974; Edgers and Karlsrud, 1985; Locat and Demers, 1988) of soft clays with

high water contents. Komamura and Huang (1974) used a rotational viscometer called the

mini viscometer to study the rhelogical nature of soils in rapid flow landslides. A new

rheological model was proposed based on the proportion of water content. According to

the model, a phase change in the soil occurs at the liquid limit of the soil. For water

contents higher than the liquid limit the visco-plastic soil behavior transforms into a

Newtonian viscous fluid behavior. A sudden drop in the viscosity is observed on the

phase change of the soil.

90

Figure 4.1 Soil flow of sensitive clay due to heavy rainfall (Reference: Geological Survey

of Canada)

91

Edgers and Karlsrud (1985) applied a theoretical viscous flow model to analyze

submarine slides with observed field velocities of about 7m/s to 10m/s. Their analyses of

field data provided estimates for soil viscosities in the range of 2 37 10 to1.4 10× × Pa.s.

Locat and Demers (1988) used a rotational viscometer to investigate viscosity of

sensitive clays. They recommended a relationship between viscosity and LI, which can be

used to compute viscosity of a soil with values of LI between 1.5 and 6. Their study

provided a simplistic initial approximation of the rheological parameters of clayey soils

for applications in landslide studies. They suggested based on field and experimental

observations that, for most sensitive clays the viscosity can be considered constant once

the soil reaches its (fluid) yield stress. The viscosity relationship proposed by Locat and

Demers (1988) for sensitive clays of high LI is

103(Pa.s) 0.00927 (LI)µ = × [4.1]

For soils with LI<1, viscometers are not recommended because of the difficulty of

placing the soil uniformly in the cylinder and removing the entrapped air (Fakher et. al

1999).

4.2.3 Strain rate effects

Strain rates significantly influence the viscous response of soils. The strain rates

equivalent to landslide velocities of about 7 m/s to 10 m/s are represented well in

viscometric measurements (Locat and Demers, 1988). For penetrating shafts such as a

jacked pile or a cone penetrometer sleeve the velocities are much smaller, for example in

92

CPT, the standard penetration rate is about 2 cm/sec, which induces small strain rates

compared to those occurring in landslides.

Ding et al. (2001) studied the viscous behavior of soil and obtained the shear

viscosity of soil using static and cyclic (low shear strain rate) triaxial shear tests. The soil

behavior was modeled like a non-Newtonian fluid. These tests showed a consistent trend

of decreasing viscosity with increasing strain rate. The shear viscosity was expressed as a

power function of strain rate.

Strain rate influences on shear behavior of clays in pre-failure stress conditions

have been studied using shear tests such as the triaxial test (e.g. Mitchell, 1964; Sheahan

et al., 1996) and the simple shear test (e.g. Matesic and Vucetic, 2003). To model the

shear strain rate dependent behavior of saturated clay, Sheahan et al. (1996) presented

experimental results for low strain rates (0.0001 %/s to 10 %/s) using an automated

triaxial device. They concluded that irrespective of the stress history, clays are expected

to behave as viscous materials at very high strain rates. However, the shear viscosity has

not been explicitly determined. Applied stress in these experiments was less than the

failure (CS) stress. Sheahan et al. (1996) suggested examining the behavior of clays at

higher strain rates for applications to in-situ penetration tests, where the soil reaches CS

and exhibits post-failure flow.

93

4.3 Modeling of viscous behavior

Fluids are classified as Newtonian or non-Newtonian based on their relationship

between shear stress and shear (strain) rate while in flow. A fluid with a linear

relationship between shear stress and shear rate is known as a Newtonian fluid. Water is

an example of a Newtonian fluid. Any fluid that does not obey the linear relationship

between shear stress and shear strain rate is a non-Newtonian fluid. High molecular

weight liquids such as slurries or pastes are usually non-Newtonian.

Yield stress fluids are the fluids that flow only when the applied shear stresses

exceed a critical (yield) value. The soil mass at CS is similar to a yield stress fluid which

will flow for applied stress exceeding the CS shear stress. The flow behavior of clays at

CS can be modeled using Newtonian (e.g. Bingham) or non-Newtonian flow models as

discussed in previous chapter. Flow behavior of soft clays and clay slurries has been

studied (e.g. Matsui and Ito, 1977; Inoue et al., 1990; Tan et al., 1990; Fakher et al.,

1999) by adopting various rheological models. It was observed that at low strain rates

( 17%s−< ), the behavior of soft clays can be represented by the Bingham model (Fakher

et al., 1999). In general for soils, pµ is not constant, but varies nonlinearly with the shear

strain rate, similar to curve B (Figure 3.1).

Post-failure flow of clays can be represented as illustrated in Figure 3.2. Non-

Newtonian yield stress fluids are characterized by a yield stress and slowly decreasing

viscosity with strain rate. This type of behavior can be represented by the relationship for

94

plastic flow proposed by Casson (1959). This relationship is expressed, using the notation

of CS shear stress and viscosity in this dissertation, as

1 1 12 2 2

cs p cs( ) ( ) ( ) ; forτ = τ + µ γ τ ≥τ& [4.2]

The above relationship (called Casson fluid) has been successfully applied to a

diverse range of materials similar to soft clays (Nguyen and Boger 1992). Locat and

Demers (1988) suggested that most sensitive clays behave as a Bingham or a Casson

fluid, the latter being more appropriate for less sensitive clays. Assuming Casson’s

relationship is applicable to clays at low LI (less sensitive), the values of total shear

stress, CS shear stress and the shear strain rate need to be known to calculate the shear

viscosity (Equation 4.2).

4.4 Shear viscosity using the fall cone test

4.4.1 Penetration test

Penetration tests using a needle are commonly used to determine the viscosity of

materials such as asphalts (e.g. Puzinaukas, 1967; Tons and Chritz, 1975) and paste-

extrudable explosives (e.g. Picart et al., 1999), which have consistency similar to clays.

The fall cone test (Hansbo 1957) shown in Figure 4.2, is a penetration test widely used in

geotechnical engineering. This test appears to have the potential to measure shear

viscosity of clays and broaden its usage. In this research, the fall cone test is explored to

95

Figure 4.2 The fall cone test apparatus

Soil cup Cup platform

Fall cone

Dial-needle assembly

96

estimate shear viscosity of clays at low liquidity indices. A low LI in the context of this

research refers to LI < 1.5.

4.4.2 The fall cone test

The fall cone test consists of a solid metal cone that freely penetrates a soil mass

placed in a standard size cup (Figure 4.2). The depth of cone (tip) penetration is used to

determine the liquid and plastic limits (Wroth and Wood, 1978; Wood, 1982; Budhu,

1985; Zreik et al., 1995; Feng, 2000) and the undrained shear strength (Hansbo, 1957;

Houlsby, 1982; Wood, 1985; Shimobe, 2000; Koumoto and Houlsby, 2001) of fine-

grained soils. Shear strengths greater than 0.075 kPa and corresponding LI less than 1.7

(Wroth and Wood, 1978) can be effectively measured using the fall cone test.

The conventional fall cone test apparatus is shown in Figure 4.2. The procedure to

conduct a fall cone test is briefly listed below:

1. The fall cone apparatus is stationed on a leveled surface.

2. The given soil to be tested is placed uniformly in a 55mm diameter cylindrical

soil cup. Care should be taken to avoid air voids in the sample. The top soil

surface of a completely filled cup is leveled using a flat spatula. The cup is then

placed centrally on the cup platform below the cone (Figure 4.2).

97

3. The fall cone (Figure 4.3) with an apex angle of 30o and a total mass of 80 grams

is lowered and suspended such that the cone tip is just in contact with the soil

surface.

4. The cone is then released and allowed to fall freely.

5. The final cone penetration depth, fh , into the soil, marked on the dial-needle

assembly (Figure 4.2) is recorded.

The water content corresponding to a penetration depth of 20 mm defines the

liquid limit. Four or more tests on the same soil with different water contents should be

conducted because of the difficulty to achieve a penetration of 20 mm from a single test.

The results of these tests are plotted as water content (ordinate, arithmetic scale) versus

penetration depths (absicca, logarithm scale). A best fit straight line (liquid state line)

linking the data points is drawn (Figure 4.4), and the water content (liquid limit)

corresponding to a penetration depth of 20 mm is calculated.

A recent study by Feng (2000) suggests that the results of the fall cone test for

soils are better represented by a linear relationship between the logarithm of cone

penetration and the logarithm of water content, given by

flog(w) log(c) mlog(h )= + [4.3]

where w is the water content, c is a constant that represents the water content

corresponding to fh 1mm= and m is the slope of the plotted linear relationship. Equation

4.3 can be used to determine the liquid limit ( LLw ), and the plastic limit ( PLw ). For a

.

98

Figure 4.3 Fall cone and the soil cup

99

Figure 4.4 Liquid limit from typical fall cone test results

30

35

40

45

50

55

60

10 100

Penetration of cone (mm) –logarithmic scale 20

Wat

er c

onte

nt, w

(%)

Liquid Limit ( LLw )

100

given soil, c and m can be determined from the results of as few as four fall cone tests.

The plastic limit is computed (Feng, 2000) from

mPLw c(2)= [4.4]

Hansbo (1957) conducted theoretical and experimental investigations of the fall

cone test and showed that the undrained CS shear strength is related to fh and can be

expressed as

cs 2f

KWh

τ = [4.5]

where W is weight of the cone and K is a constant called as the fall cone factor. From the

theoretical and experimental work (Houlsby, 1982; Wood, 1985; Koumoto and Houlsby,

2001) it was shown that K is influenced by the cone geometry, cone roughness, soil

heave around the cone and shear strain rate (dynamic) effects. According to Hansbo

(1957), the resistance to the penetration of the cone during free fall motion depends on

the static shearing resistance (interfacial friction) and viscous resistance (viscous flow).

The dynamic effects due to cone motion during penetration influence the shear

stress in the soil. Koumoto and Houlsby (2001) proposed a reduction factor, λ, to modify

K to be used in Equation (4.5). This modified K accounts for the effects of dynamic

component of shear resistance. The factor, 0.74λ = , was estimated by extrapolating the

published low strain rate (triaxial) test) test results to a higher level strain rate that

approximates the strain rate occurring in the fall cone test.

The depth of static equilibrium, sh , is the depth of cone penetration required so

that the cone weight is balanced by the soil resistance. Static equilibrium of the cone is

101

determined by equilibrium equations in which the cone is assumed at rest (static)

condition and shear stress in the soil is equal to the CS shear strength, csτ . The solution

of the dynamic equation of cone motion (Hansbo, 1957; Houlsby, 1982), neglecting the

inertial forces of soil, shows that, f sh 3h= . Hence the cone is in motion even after sh is

achieved. As the cone is in motion there will be an additional viscous shear stress acting

above csτ of the soil. The total dynamic shear resistance on the cone is the sum of static

shear and viscous components. The approach taken in this study is to separate the static

and viscous shear resistance in order to estimate the shear viscosity of soil.

4.4.3 Theoretical approach

When a fall cone penetrates soil, the soil just below and adjacent to the cone tip

reaches CS and flows around the cone as illustrated in Figure 4.5. The soil mass at CS

behaves like a yield stress fluid as discussed in previous chapters. The equation of motion

of a cone at any penetration depth, h, in the soil is

2ma mg F h= − τ [4.6]

where m is mass of the cone, a is acceleration of the cone at depth, h, g is the acceleration

due to gravity, τ is the (dynamic) shearing resistance and F is the non-dimensional cone

resistance factor (Houlsby, 82; Koumoto and Houlsby, 2001) expressed as

2chF N tan ( )= π θ [4.7]

102

Figure 4.5 Illustration of the soil state around a fall cone

θ

h Post-failure (Critical)

Soil in pre-peak (pre-failure) state

δ

103

where chN is the modified bearing capacity factor of the cone, which accounts for the soil

heave around the cone, and θ is the half cone angle. For a 30o (i.e. 15θ = o ) semi rough

cone, the value of chN is 7.457 (Koumoto and Houlsby, 2001).

When the free fall motion of the cone is initiated, its acceleration decreases from

an initial value, g, due to the soil resistance. At a certain depth of penetration, heq, the

acceleration of the cone becomes zero, i.e. the net force on the cone is zero. In this

investigation, this depth is referred to as the dynamic equilibrium position. The cone

thereafter decelerates further finally coming to rest at hf. The velocity of the cone

increases from the zero (beginning of test) and reaches a maximum value at heq.

Thereafter, the velocity reduces until the cone finally comes to rest. The additional

viscous stress due to cone motion causes the dynamic equilibrium (heq) to be achieved at

penetration depth lesser than the theoretical equilibrium depth, hs. If the dynamic shear

resistance at this point of equilibrium in a fall cone test is estimated, the viscous

component can be extracted by subtracting the static component. From Equation (4.6),

the dynamic equilibrium condition ( a 0= ) is

2eqmg F h= τ [4.8]

from which, we get

2 2eq eq

mg WFh Fh

τ = = [4.9]

From Equations (4.2), (4.5) and (4.9), we can express the term pµ γ& (the viscous

component of shear resistance) as

104

21 121 1 2 22 2

p cs 2 2eq f

W KWFh h

⎡ ⎤⎛ ⎞⎛ ⎞ ⎛ ⎞⎢ ⎥µ γ = τ − τ = −⎜ ⎟⎜ ⎟ ⎜ ⎟⎢ ⎥⎜ ⎟ ⎝ ⎠⎝ ⎠ ⎝ ⎠⎢ ⎥⎣ ⎦

& [4.10]

The static resistance from the fall cone test data (W and hf) can be calculated using the

expression proposed by Komoto and Houlsby (2001). The relationship among K, λ and F

can be written as (Koumoto and Houlsby, 2001)

3KFλ

= [4.11]

Using the above relationship, K = 1.33 for a 30o semi rough cone.

Substituting Equation (4.11) and 0.74λ = in Equation (4.10), and simplifying the

equation further, we get

2

peq f

0.67 1KWh h

⎛ ⎞µ γ = −⎜ ⎟⎜ ⎟

⎝ ⎠& [4.12]

To estimate the viscosity of clay using the fall cone test, γ& and eqh must be determined.

The shear strain pattern around the cone is complex and difficult to determine

accurately. Consequently, γ& during fall cone motion, which varies with h, is also difficult to

measure. However, the approximate γ& at eqh can be estimated from the expression

(Koumoto and Houlsby, 2001) given as

f

2 g 32.44 h

δγ =& [4.13]

where δ the inclination angle (in radians) of heaved soil surface (Figure 4.5). The angle

of heaved soil surface is a function of the cone angle (2θ) and the cone roughness. For a

105

30o semi rough cone, 5.77δ = o (Koumoto and Houlsby, 2001). Substituting in Equation

(4.13), γ& can be estimated as

f

10.34h

γ =& [4.14]

The shear viscosity of a given soil can then be computed using Equations (4.4.10) and (4.14)

as

2

p feq f

0.67 12.94KW hh h

⎛ ⎞µ = −⎜ ⎟⎜ ⎟

⎝ ⎠ [4.15]

4.4.4 Experiments

A standard cone of apex angle, 30o and height, 35 mm, was used for the test. The

procedure used for conducting the tests is similar to the conventional fall cone test explained

previously in the Section 4.4.2. However, to record continuous data of penetration (h) with

time (t), the dial-needle measurement assembly was replaced by a calibrated linear variable

displacement transformer (LVDT) and computer data acquisition system. Photographs of

the modified experimental setup are shown in Figure 4.6 and Figure 4.7. An LVDT

connected to the top of the cone shaft (Figure 4.7) was used to record the depth of

penetration in terms of voltage as a continuous function of time. The voltage data measured

by the LVDT was sent to a data acquisition system

106

Figure 4.6 Modified experimental setup for the fall cone test

107

Figure 4.7 LVDT connected to the top of fall cone shaft

LVDT

108

(DAS). The commercial software, TEST POINT 3.0, was implemented to log, observe

and retrieve the measurements.

Kaolin (clay) of liquid limit ( LLw ) and plastic limit ( PLw ) of 47% and 30%

respectively, was used for the test samples. Dry kaolin was mixed with distilled water to

achieve the desired water content and stored in an air tight container for 24 hours. The

wet soil was then thoroughly mixed and placed in the soil cup for the experiment to be

conducted. Water content of the soil sample placed in the soil cup was measured. Tests

were performed on soil samples with water contents in the range 35.8 % to 65.7 %. The

total mass of the cone assembly (cone, shaft and LVDT) was 93 grams. A computer data

acquisition system recorded the times and penetrations at a frequency of 1 kHz. For soil

states near the plastic limit, additional masses of 50 grams and 150 grams were added to

increase the depth of penetration. Experimental details of water contents and the masses

used for the tests are summarized in Table 1.

4.5 Results and discussion of the experimental study

The data recorded for time versus penetration was filtered to remove noise. A

MATLAB code, written by the author for the curve fitting algorithm to remove noise, is

listed in Appendix A1. It was found that except for a few initial and end points, a 5th

order polynomial was the best fit to define the penetration, h, as a continuous function of

time, t. Measured h versus t data along with the 5th order polynomial fit for Test no. 15C1

is shown in Figure 4.8. This polynomial was differentiated once and twice to obtain

109

Table 1 Experimental details

(a) Using additional mass of 150 gm (total mass 243 gm, W = 2.38 N)

(b) Using additional mass of 50 gm (total mass 143 gm, W = 1.40 N)

(c) Using cone assembly only (total mass 93 gm, W = 0.91 N)

Test No. Water content (%) 15C1 41.93 15C2 35.84 15C3 43.40 15C4 45.51 15C5 43.63 15C6 47.43

Test No. Water content (%) 5C1 54.28 5C2 54.25 5C3 59.89

Test No. Water content (%)

C1 54.25 C2 65.70 C3 58.77

110

Figure 4.8 Penetration-time relationship of the cone (Test no. 15C1)

0 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.080

0.002

0.004

0.006

0.008

0.01

0.012

0.014

Time, t (s)

Pene

tratio

n, h

(m)

5th order polynomial h f (t)=

heq

hs

o Measured data points

111

velocity and acceleration of the cone, respectively. A MATLAB code for this

computation is listed in Appendix A2.

These results for Test no. 15C1 are shown in Figure 4.9. It was observed that at

heq (point A), the velocity of the cone reaches its maximum value and the acceleration is

zero. Pertinent data of fh and eqh from the tests is summarized in Table 2. The resulting

plots of time-penetration, velocity and acceleration for Tests 15C2 to 15C6, 5C1 to 5C3

and C1 to C3 are given in Appendix B1.Assuming hf as the datum level (Figure 4.10), the

total energy (Ec) of the cone penetrated to a depth h can be written as the sum of potential

energy (PE) and kinetic energy (KE),

2c f

1E mg[h h] mv2

= − + [4.16]

PE, KE and cE as the cone penetrates the soil for Test no. 15C1 are shown in Figure

4.11. The maximum KE occurs at the dynamic equilibrium point A. The total energy

remaining at point A causes the cone to penetrate further until all the available energy is

dissipated in plastic deformation of clay.

The shear viscosity ( pµ ), calculated using Equation (4.15) for different values of

liquidity index (LI) together with csτ and τ , is summarized in Table 3. The relationship

between pµ (Pa.s) and LI for kaolin is depicted in Figure 4.12. The results fit quite well

(regression coefficient = 96%) to an exponential function expressed as

p 817.6exp( 1.43LI)µ = − [4.17]

112

Figure 4.9 (a) Velocity and (b) acceleration of the cone (Test no. 15C1)

0 0.002 0.004 0.006 0.008 0.01 0.012 0.0140

0.05

0.1

0.15

0.2

0.25

0.3

h (m)

Vel

ocity

(m/s

) (a)

0 0.002 0.004 0.006 0.008 0.01 0.012 0.014-15

-10

-5

0

5

10

15

A

Acc

eler

atio

n (m

/s2 )

h (m)

(b)

A

heq

hs

113

Table 2 Test data

Test No.

Water content

(%) LI hf (mm) heq (mm) f

shh

3= (mm)

15C1 41.93 0.70 13.77 7.46 7.95 15C2 35.84 0.34 9.07 5.00 5.24 15C3 43.40 0.79 17.88 9.57 10.33 15C4 45.51 0.80 19.44 10.51 11.22 15C5 43.63 0.91 18.18 9.70 10.49 15C6 47.43 1.02 24.55 13.13 14.17

Test No.

Water content

(%) LI hf (mm) heq (mm) f

shh

3= (mm)

5C1 54.28 1.43 20.40 10.88 11.78 5C2 54.25 1.43 23.55 12.26 13.60 5C3 59.89 1.76 27.59 14.64 15.93

Test No.

Water content

(%) LI hf (mm) heq (mm) f

shh

3= (mm)

C1 54.25 1.43 21.23 10.79 12.26 C2 65.70 2.10 30.19 16.47 17.43 C3 58.77 1.69 18.81 10.52 10.86

114

Figure 4.10 Fall cone energy (Ec)

115

Figure 4.11 PE, KE and Ec with penetration depth (Test no. 15C1)

0 0.002 0.004 0.006 0.008 0.01 0.012 0.0140

0.005

0.01

0.015

0.02

0.025

0.03

0.035

0.04

Kinetic Energy (KE)

Potential Energy (PE) Total Energy, Ec = PE +KE

A

Ener

gy (N

.m)

h (m)

116

Table 3 Estimated shear strength and shear viscosity

Test LI

csτ (kPa)

τ (kPa)

1(sec )−γ&

pµ (Pa.s)

15C1 0.70 16.71 25.64 2.90 328 15C2 0.34 38.49 57.19 3.57 516 15C3 0.79 9.91 15.60 2.54 253 15C4 0.91 8.39 12.93 2.44 200 15C5 0.80 9.59 15.18 2.52 253 15C6 1.03 5.26 8.29 2.17 158 5C1 1.43 4.48 7.10 2.38 125 5C2 1.43 3.36 5.59 2.22 127 5C3 1.76 2.45 3.92 2.05 84 C1 1.43 2.69 4.70 2.34 119 C2 2.10 1.33 2.02 1.96 36 C3 1.69 3.43 4.94 2.48 55

117

Figure 4.12 Shear viscosity - LI relationship for kaolin used in this study

LI

0

100

200

300

400

500

600

0.00 0.50 1.00 1.50 2.00 2.50

p (Pa.s)µ

p

2

817.6exp( 1.43LI)

R 0.96

µ = −

=

118

Locat and Demers (1988) proposed a similar type of relationship for computing the

viscosity of sensitive clays with 1.5 LI 6.0≤ ≤ using viscometers.

4.6 Application to the CPT

Cone penetrometer test (CPT) results have been extensively used in deriving the

static skin friction stress required to compute the axial load capacities of pile foundations.

It is often assumed that the end bearing capacity is related to the cone tip resistance and

the skin friction capacity is related to the sleeve resistance of the CPT. As the cone tip

penetrates, the soil around the cone reaches CS. When the sleeve is pushed after the cone,

viscous flow occurs adjacent to the sleeve. The viscous drag stress on the sleeve can be

computed using Equation (3.22) as

p z osv

2π V βf

[4.18]

where fsv is the viscous drag stress and D is the outer diameter of the sleeve.

The measured (dynamic) sleeve friction is a summation of the static frictional

stress and the viscous drag stress, which can be written as

s ss svf f f= + [4.19]

where, fs, is the measured sleeve skin friction and fss is the static friction stress

component.

CPT uses a friction sleeve of diameter, D = 3.57 cm (cross-sectional area of 10

cm2) and a rate of penetration of 2 cm/s. If the diameter of the CS zone is assumed to be 4

119

times the diameter of the sleeve ( o 4λ = ), we get o 1.56β = from Equation (3.21). The

viscous drag stress on the friction sleeve (Equation 4.18) is

sv pf 1.75= µ (Pa.) [4.20]

In the event of a conventional CPT being performed on the kaolin tested in this study,

and has a LI of 0.7, the viscous drag would be about 575 Pa.

The static skin friction stress can be computed from the measured cone sleeve

resistance using

ss sf (1 )f= − ξ [4.21]

where the factor, sv

s

ff

ξ = , is the ratio of viscous drag stress to measured sleeve friction

stress. The viscous drag, according to Equation (3.22), is linearly related to penetration

velocity for a given shear viscosity and the size of the CS zone. CPT results (Marsland

and Quarterman, 1982) indicate that for soft clays, ξ increases linearly with increments of

penetration rate above 2 cm/s ( ξ increases by about 10 % for every increment of 2 cm/s

above its standard penetration rate). A similar application of this research for jacked piles

in clay is illustrated with an example given in Appendix C.

4.7 Conclusion

The theory of the fall cone test, currently used to determine the index and strength

properties of soil, is extended to consider the viscous drag as the cone penetrates the soil

at low liquidity index (or water content). This extended theory shows that the shear

120

viscosity of clays can be determined using continuous record of penetration-time data in a

fall cone test. For the kaolin tested, the shear viscosity decreases exponentially with LI,

This is consistent with the variation of viscosity with LI for clays with higher liquidity

indices, reported in the literature. The shear viscosity of clay can be utilized to determine

the viscous resistance on the sleeve in CPT or jacked piles in clays.

121

CHAPTER 5

SUMMARY, CONCLUSIONS AND RECOMMENDATIONS

5.1 Summary and conclusions

A new approach based on creeping flow hydrodynamics to analyze viscous drag in

post-failure flow of clays has been described assuming the soils at critical state (CS) are

similar to yield stress fluids. This approach can be implemented to analyze the problems

associated with post-failure soil flow. Based on the theoretical solution for viscous drag

on a penetrating shaft in clay, the following conclusions are drawn.

1. Viscous resistance is an additional component above the static interfacial

resistance on a penetrating shaft in clays.

2. Viscous drag on the shaft depends on the size of the CS zone, shear viscosity of

the soil in this zone and velocity of the shaft.

3. Viscous drag on the shaft increases significantly when the size of the CS zone is

less than four times the shaft radius. Limiting viscous drag occurs when the size

of the CS zone exceeds six times the shaft radius.

4. The viscous soil mass close to the shaft surface is dragged downwards and in the

far field, the soil moves upwards (heave). The radial distance from the shaft at

which the upward movement of soil is initiated increases with increases in the

size of the CS zone.

122

5. The shear viscosity of soil at CS is an important parameter to investigate post-

failure soil response and to determine viscous resistance in soil penetration

problems.

Standard methods of defining the viscosity of soil at CS are not presently available.

The shear viscosity has been investigated for clays with higher water contents to study

soil flows in landslides using viscometers. Viscometers are difficult to operate on soils

with low water contents. In this research, it is shown that a fall cone test can be used to

determine the shear viscosity of clays at low water contents (LI < 1.5). Fall cone

experiments were conducted using kaolin samples. Based on the reevaluated theory for

the fall cone test and experimental results, it can be concluded that

1. The fall cone test is a promising tool to estimate the shear viscosity of clays at CS.

2. The shear viscosity decreases exponentially with increases in LI.

3. Casson’s model was used in this study and it is uncertain whether this model

captures the viscous response of soils (in the plastic range) accurately despite

suggestion in the literature of its potential appropriateness to clays.

The results of this research are in agreement with the hypotheses proposed.

123

5.2 Key Contributions

The following are the key contributions of this study.

1. A new theoretical approach based on hydrodynamics to study viscous resistance

in soil penetration problems.

2. Advancement of the theory and experimental procedures of the fall cone test to

estimate shear viscosity of clays at low water contents.

3. Establishment of a relationship between the shear viscosity and LI of clays.

4. Bring attention to the importance of viscous behavior of soils at CS for post-

failure response.

5.3 Potential applications

Potential applications of this research include:

1. Determination of the viscous soil resistance on jacked piles in clays.

2. Extraction of the static friction resistance on the sleeve of a CPT to design pile

foundations.

3. Estimation of the penetration force required in advancing of casing in soil for

drilling or tunneling operations.

124

5.4 Recommendations for future research

For future research, the following is recommended.

1. Comparison of the results of this theoretical solution with field and lab tests.

2. Conducting experimental studies to investigate the influences of shaft cross-

section, penetration velocity and the in situ soil properties on the size of the CS

zone.

3. Investigation of rheological flow models for their applications to the post-failure

flow of clays.

4. Comparison and validation of shear viscosities of clay measured by fall cone test

with other potential measurement techniques.

5. Determination of shear viscosities for different types of clays and examining their

relationships with well known soil parameters for practical applications.

125

APPENDIX A

MATLAB CODES

126

A1. MATLAB CODE TO FILTER THE RAW EXPERIMENT DATA % MATLAB DATA FILTER CODE – FALL CONE EXPERIMENT % INPUT OF RAW DATA input = xlsread('filter.xls'); time_data = input(:,20); time_data = time_data'; voltage_data = input(:,21); voltage_data = voltage_data'; time = time_data; % START AND END POINT OF EXPERIMENT 3. first = 1204; last = 1300; t0=first; t1= last-first +1; %CONSTANTS : 1. LVDT CALIBRATION 2.FILTER LENGTH A= 6.0141052476402200; filterlength =14; repeat = 12; %i= index repeating median filter medrepeat =5; %k=index repeating mean filter meanrepeat =20; output = medfilt1(voltage_data,filterlength); for i= 1:medrepeat; output = medfilt1(output,repeat); end length = size(output,2);

127

%DISCARDING THE FEW INITIAL DATA POINTS WITH ERROR AND NORMALISING THE DATA WITH INITIAL VALUE output = output(filterlength:length); time_data = time_data(filterlength:length); output = output(1) - output; output(1:t0-1)=[]; time_data(1:t0-1)= []; time_data = time_data - time_data(1); output = output - output(1); %DATA AFTER MEDIAN FILTERING final_output = output(1:t1); final_time = time_data(1:t1); %MULTIPLYING BY CALIBRATION CONSTANT TO CONVERT mV to mm medh = A*final_output; medianh = medh; % MEAN FILTER FOR SMOOTH CURVE k=1; while k < meanrepeat for i = 3 : t1-2 medh(i)= mean(medh(i-2 : i+2)); medh(i)= mean(medh(i-2 : i+2)); end k = k + 1; end meanh= medh; %ORIGINAL DATA REQUIRED TO SELECT APPROPRIATE FILTER LENGTH original_data = voltage_data(filterlength:length); original_data(1:t0-1)=[]; original_data=original_data(1:t1);

128

original_data = original_data (1)- original_data; originalh = A*original_data; %PLOTS figure, plot(2,2,3), plot(meanh,final_time); title ('Time vs Penetration ') xlabel('Displacement, h (mm)'); ylabel('Time (sec)'); grid on; %SAVE & WRITE THE DATA IN EXCEL FILE final(:,1) = final_time'; final(:,2) = meanh'; dlmwrite('C:\sandeep\EXPERIMENTS\15C\15C1.xls', final, '\t');

129

A2. MATLAB CODE TO COMPUTE PENETRATION-TIME POLYNOMIAL FIT, VELOCITY ETC. % MATLAB COMPUTATION CODE – FALL CONE EXPERIMENT clear,close all; % INPUT i=1; %coefficient to read data from corresponding excel sheet sheetname = strcat('Sheet',int2str(i)); input = xlsread('15C.xls',sheetname); %Experimental Time-Penetration Data time_data = input(:,1); time = time_data'; penetration_data = input(:,2); penetration = penetration_data'; penetration_m = penetration./1000; %changing h from mm to m index = length(time); %Final penetration depth h_fmm=penetration_data(index); h_fm = penetration_m(index); t_f = time(index); h_s = h_fm/sqrt(3); m = 0.150+0.09295;% mass of cone in kg g = 9.81; W= m*g % cone weight in N % Defining 5th order polynomial order = 5; p = polyfit(time,penetration_m,order); % array dimension is 1 by order+1 % 1st derivative (i.e. velocity) is a 4th order polynomial for i=1:order p_d(i) = (order-i+1)*p(i);

130

end % 2nd derivative (i.e. velocity) is a 4th order polynomial for i=1:order-1 p_d_2(i) = (order-i)*p_d(i); end % Computing values of penetration, velocity and acceleration from % fitted expression h_fit = polyval(p,time); v_exp = polyval(p_d,time); a_exp = polyval(p_d_2,time); % equilibrium point and equlibrium depth [max_velocity,eq_point] = max(v-exp(10:index-10) % equilibrium point h_eq = h_fit(eq_point); % CONE CONSTANTS Koumoto and Houlsby (2001) % 30 Degree Semi-Rough (alpha =0.5) Cone Nch= 7.457; F=(pi*Nch*tan(pi/12)*tan(pi/12)); K=1.33; %K=0.74*(3/F); % total and CS shear stress tau = ((W)/(F* h_eq*h_eq))./1000; % in kPa t_cr = ((K*W)/(h_fm*h_fm))./1000; % in kPa % Average strain rate strain_rate = 0.34*sqrt(1/h_fm); % Shear Viscosity using Casson's Model visc1=2.94*K*W*sqrt(h_fm); visc2=(0.67/h_eq); visc3=(1/h_fm); shear_visc = visc1*((visc2-visc3)^2);

131

% Fall Cone Energy p_e = m*g*(h_fm - h_fit); k_e = 0.5*m*(v_exp).^2; total_energy = p_e + k_e; figure plot(time,h_fit,time,penetration_m,'o'); xlabel('time,t (sec)'); ylabel('h (m)'); title('15C1') grid on; legend('Polynomial Fit','Measured',2) figure plot(h_fit(20:index-20),v_exp(20:index-20)); title('15C1') xlabel('h (m)'); ylabel('v (m/s)'); grid on; filename = strcat('15C1'); save(filename,'penetration_m','h_fit','time','v_exp','a_exp','t_cr','tau','shear_visc');

132

APPENDIX B

FALL CONE TEST RESULTS

133

(a)

(b)

Figure B1 (a) Penetration-time relationship (b) velocity of the cone for test 15C2

134

(a)

(b)

Figure B2 (a) Penetration-time relationship (b) velocity of the cone for test 15C3

135

(a)

(b)

Figure B3 (a) Penetration-time relationship (b) velocity of the cone for test 15C4

136

(a)

(b)

Figure B4 (a) Penetration-time relationship (b) velocity of the cone for test 15C5

137

(a)

(b)

Figure B5 (a) Penetration-time relationship (b) velocity of the cone for test 15C6

138

(a)

(b)

Figure B6 (a) Penetration-time relationship (b) velocity of the cone for test 5C1

139

(a)

(b)

Figure B7 (a) Penetration-time relationship (b) velocity of the cone for test 5C2

140

(a)

(b)

Figure B8 (a) Penetration-time relationship (b) velocity of the cone for test 5C3

141

(a)

(b)

Figure B9(a) Penetration-time relationship (b) velocity of the cone for test C1

142

(a)

(b)

Figure B10 (a) Penetration-time relationship (b) velocity of the cone for test C2

143

(a)

(b)

Figure B11 (a) Penetration-time relationship (b) velocity of the cone for test C3

144

APPENDIX C

APPLICATION USING AN ILLUSTRATIVE EXAMPLE

145

The example given below is only intended to illustrate the potential application of this

research to practical problems such as a jacked pile in clay.

EXAMPLE :

A cylindrical steel shaft of radius 0.5 m is to be jacked with velocity ( zV ) of 5 cm/sec in

clay layer, which extends upto a depth of 20 m below the ground surface. The undrained

shear strength of this soil is 20 kPa and the LI is 0.6. The relationship between pµ (Pa.s)

and LI determined by using fall cone tests is

p 2000exp( 1.0LI)µ = − [C.1]

If the embedment length of the shaft is 10 m, compute

(a) the static and viscous resistance components on the shaft during penetration. The

jacked pile creates a CS zone of radius equal to 2 times the radius of the pile

(i.e. 0λ 2= ).

(b) The change in total penetration resistance if zV 10cm / sec= and 0λ 1.4= .

SOLUTION:

(a)

1. Static friction resistance:

The static skin frictional stress on a jacked pile in clay under undrained condition (no

volume change) is:

ss uf s= α [C.2]

146

where α is a skin friction factor and su is the undrained shear strength of the clay. For the

given clay with us 25kPa, 1.0≤ α = (Budhu, 2000). The total static skin friction

resistance on the jacked pile is: 0 ss(2 r f )L 628.3kNπ =

2. Viscous friction resistance:

The viscous skin friction resistance per unit length on a penetrating shaft using Equation

(3.22), developed in this study is, v p z 0f 2πµ V β=

For the jacked pile shaft in this example, 2zV 5 10 m / s−= × , 0 4β = (Figure 3.5).

The viscosity of clay using Equation (C.1) is p 1098Pa.sµ ≈

The total viscous resistance on the jacked pile is: p z 0(2πµ V β )L 13.8kN=

The total soil resistance on the jacked pile is: 628.3 13.8 642.1kN+ =

(b)

For this shaft, zV 0.01m / s= and 0 10.56β = (Figure 3.5).

The viscous skin friction resistance per unit length is, v p z 0f 2πµ V β=

The total viscous resistance on the jacked pile is: p z 0(2πµ V β )L 72.6kN= The total soil resistance on the jacked pile is: 628.3 72.6 700.9kN+ =

The total penetration resistance increases by 9.2%.

147

REFERENCES

Allersma, H.G.B. (1997). Optical Analysis of Stress and Strain in Photoelastic Particle

Assemblies. PhD Thesis, Department of Civil Engineering, Delft University of

Technology, Netherlands.

Baligh, M.M. (1985). “Strain path method.” J. Geotech. Eng. Div., ASCE, 111(9), 1108-

1136.

Berry, A.J. (1988) “The dynamic penetration of clays.” Proc. 1st International

Symposium on Penetration Testing ISOPT I, Orlando., A. A. Balkema, Rotterdam,

673-679.

Budhu, M. (1985). “The effect of clay content on liquid limit from fall cone and British

cup device.” Geotech. Testing J., ASTM, 1985, 8(2), 91-95.

Budhu, M. (2000) Soil mechanics and foundations. Wiley, New York.

Budhu, M., and Wu, C. (1992). “Numerical analysis of sampling disturbances in clay

soils.” Int. J. Numeical. Analytical. Methods in Geomechanics, 16, 467-492.

148

Casson W. (1959). “A flow equation for pigment-oil suspensions of printing of the

printing ink type.” Rheology of Dispersed Systems, Pergamon, London.

Cooke, R. W., and Price, G. (1973). “Strains and displacements around friction piles.”

Proc. 8th Int. Conf. on Soil Mechanics and Foundation Engineering, 2 (1), Moscow,

U.S.S.R., 53-60.

Coulomb, C.A. (1776). “Essai sur une application des règles des maximis et minimis à

quelques problèmes de statique relatifs à l’architecture.” Memories de la

Mathematique et de Physique, presentes a l’Academie Royale des Sciences, par

divers savants, et lus dans ces Assemblees., L’Imprimerie Royale, Paris, 3-8.

Coussot, P., Nguyen, Q.D., Huynh, H.T., and Bonn D. (2002). “Avalanche behavior in

yield stress fluids. Phy. Rev. Lett., 88 (17), paper no.175501 (4).

Cummins, A.E., Kerkhoff, G. O., and Peck, R. B. (1948). “Effect of driving piles into

soft clay.” Transactions, ASCE, 115, Paper No. 2400, 275-285.

Desai, C.S. (2001) Mechanics of materials and interfaces: the disturbed state concept.

CRC Press, Boca Raton.

149

Ding, D., Li, J., and Helm, D.C. (2001). “Nonlinear shear viscosity of clayey soils over

different time scales.” Proc. 3rd Int. Conf. on Soft Soil Engineering, Hong Kong, 623

- 627.

Edgers, L., and Karlsrud, K. (1985). “Viscous analysis of submarine flows.” Proceedings

of the 4th International Conference on the Behaviour of Offshore Structures, Delft,

Developments in Marine Technology, 2, 773-784

Esrig, M. I., Kirby, R. C., Bea, R. G., and Murphy, B. S. (1977). “Initial development of

a general effective stress method for prediction of axial capacity for driven piles in

clay.” Proc. of the 9th Annual Offshore Technology Conf., Houston, OTC paper 2943,

495-506.

Fakher, A., Jones, C.F.J.P., and Clarke, B.G. (1999). “Yield stress of super soft clays.” J.

Geotech. Geoenviron. Eng. ASCE, 125 (6), 499-509.

Feng, T.W. (2000). “Fall-cone penetration and water content relationships of clays.”

Geotechnique, 23 (4), 181-187.

Flaate, K. (1972). “Effects of pile driving in clays.” Can. Geotech. J., 9, 81-88.

150

Fung, Y.C., and Tong, P. (2001). Classical and computational solid mechanics. World

Scientific, New Jersey.

Hansbo, S. (1957). “A new approach to the determination of the shear strength of clay by

the fall-cone test.” Proc. of the Royal Swedish Geotechnical Institute, 14, 1-47.

Happel, J., and Brenner, H. (1965). Low Reynolds number hydrodynamics with special

applications to particulate media. Prentice Hall Inc., Eaglewood cliffs, New Jersey.

Hill, R. (1950). The mathematical theory of plasticity. Oxford University Press, London

Houlsby, G.T., (1982). “Theoretical analysis of the fall cone test.” Geotechnique, 32 (2),

111-118.

Inoue, T., Tan, T.S., and Lee, S.L. (1990). “An investigation of shear strength of slurry

clay.” Soils and Foundations, 30(4), 1-10.

Kirby, R. C., and Wroth, C. P. (1977). “Application of critical state soil mechanics to the

prediction of axial capacity for driven piles in clay.” Proc. of the 9th Annual Offshore

Technology Conf., Houston, OTC paper 2942, 483-494.

151

Komamura, F., and Huang, R.J. (1974). “A new rheological model for soil behavior.” J.

Geotech. Eng. Div. ASCE., 100(7), 807-824.

Koumoto, T., and Houlsby, G.T. (2001). “Theory and practice of the fall cone test.”

Geotechnique, 51(8), 701-712.

Lamb, H. (1932). Hydrodynamics (6th edn.). Cambridge University Press, England.

Locat, J., and Demers, D. (1988). “Viscosity, yield stress, remolded strength, and

liquidity index relationships for sensitive clays.” Canadian Geotech. J. 25, 799-806.

Marsland, A. and Quarterman, R.S.T. (1982). “Factors affecting the measurements and

interpretation of quasi static penetration tests in clays.” Proc. 2nd European

Symposium on Penetration Testing ESOPT II, 1982, 2, A. A. Balkema, Amsterdam,

697-702.

Matesic, L., and Vucetic, M. (2003). “Strain-rate effect on soil secant shear modulus at

small cyclic strains.” J. Geotech. Geoenviron. Eng., ASCE, 129(6), 536-549.

Matsui, T., and Ito, T. (1977). “Flow mechanism of clay-layer system and microscopic

meaning on shear parameters on soils.” Proc. 9th Int. Conf. on Soil Mechanics and

Foundation Engineering, Tokyo, Specialty session 9, 143-152.

152

McNown, J.S., Lee, H.M., McPherson, M.B. and Engez, S.M. (1948). “Influence of

boundary proximity on the drag of spheres.” Proc. VII Int. Congress Applied

Mechanics, 2, 17.

Meyerhof, G.G. (1951). “The ultimate bearing capacity of foundations.” Geotechnique, 2

(4), 301-332.

Mitchell, J.K. (1976). Fundamentals of soil behaviour (1st edn.). Wiley, New York.

Mitchell, J.K. (1964). “Shearing resistance of soils as a rate process.” J. Soil Mechanics

and Foundation. Div., ASCE, 90(1), 29-61.

Murff, J.D., and Coyle, H.M. (1973). “Low velocity penetration of kaolin clay.” J.

Geotech. Eng. Div., ASCE, 99 (5), 375-389.

Nguyen, Q.D., and Boger, D.V. (1992). “Measuring the flow properties of yield stress

fluids.” Annual Review Fluid Mechanics, 24, 47-88.

Panton, R.L. (1984). Incompressible flow. Wiley, New York.

153

Picart, D., Manceau, R., and Faure, J. P. (1999). “Characterization of paste-extrudable

explosives using a penetration test.” Propellants, Explosives, Pyrotechnics, 24, 227-

231.

Puzinaukas, V. (1967). “Evaluation of properties of asphalt cements with emphasis on

consistency at low temperatures.” Proc. of the Association of Asphalt Paving

Technologists, 36, 489-540.

Randolph, M. F., and Wroth, C. P. (1978). “Analysis of deformation of vertically loaded

piles.” J. Geotech. Eng. Div., ASCE, 104(12), 1465-1488.

Randolph, M. F., Carter, J. P., and Wroth, C. P. (1979). “Driven piles in clay – the

effects of installation and subsequent consolidation.” Geotechnique, 29 (4), 361-393.

Randolph, M.F., Martin, C., and Hu, Y. (2000). “Limiting resistance of a spherical

penetrometer in cohesive material.” Geotechnique, 2000, 50 (5), 573-582.

Ray, M. (1936). “Application of Bessel functions to the solution of problem of motion of

circular disc in viscous liquid.” Phil Mag. Vol. 21, Series 7, 546-564.

Roscoe, K.H., and Burland, J.B. (1968). “On the generalized stress-strain behaviour of

wet clay.” Engineering Plasticity, Cambridge Univ. Press, 535-609.

154

Sagaseta and Whittle, 2001 Sagaseta, C., and Whittle, A.J. (2001). “Ground movements

due to pile driving in clay.” J. Geotech. Geoenviron. Eng., ASCE, 127 (1), 55-66.

Schofield, A., and Wroth, C.P. (1968). Critical state soil mechanics. McGraw-Hill,

London.

Sheahan, T.C., Ladd, C.C., and Germaine, J.T. (1996). “Rate-dependent undrained shear

behavior of saturated clay.” J. Geotech. Eng., ASCE, 122 (2), 99-108.

Shimobe, S. (2000). “Correlations among liquidity index, undrained shear strength and

fall cone penetration of fine-grained soils.” Coastal Geotech. Eng. in Practice,

Balkema, Rotterdam, 141-146.

Tan, T.S., Goh, T.C., Karunaratne, G.P., Inoue, T. and Lee, S.L. (1991). “Yield stress

measurement by a penetration method.” Canadian Geotech. J., 28(4), 517-522.

Tan, T.S., Yong, K.Y., Leong, E.C., and Lee, S.L. (1990). “Behaviour of clay slurry.”

Soils and Foundations, 30(4), 105-118.

Teh, C.I., and Houlsby, G.T. (1991). “An analytical study of the cone penetration test in

clay.” Geotechnique, 1991, 41(1), 17-34.

155

Tomlinson, M.J. (1957). “The adhesion of driven piles in clay soils.” Proc. 4th

International Conference on Soil Mech. and Found. Eng., Vol. 2, 66-71.

Tons, E., and Chritz, P. (1975). “Grading of asphalt cements by viscosity.” Proc. of the

Association of Asphalt Paving Technologists, 44, 387-415.

Torrance, J.K. (1987). “Shear resistance of remoulded soils by viscometric and fall-cone

methods: A comparison for Canadian sensitive marine clays.” Canadian Geotech. J.,

24(2), 318-322.

Turnage, 1973 Turnage, G.W. (1973). “Influence of viscous-type and inertial forces on

the penetration resistances of saturated fine-grained soils.” J. of Terramechanics, 10,

No. 2, 63-76.

van den Berg, P. (1994). Analysis of soil penetration. PhD thesis, Delft University of

Technology, Netherlands.

Vesic, A.S. (1973). “Analysis of ultimate loads of shallow foundations.” J. Geotech. Eng.

Div. ASCE., 99(1), 45-73.

156

Vyalov, S.S. (1986). Rhelogical fundamentals of soil mechanics. Developments in

Geotechnical Engineering, Vol.36, Elsevier, Amsterdam.

Whitman, R.V. (1957). “The behavior of soils under transient loadings.” Proc. 4th Int.

Conf. on Soil Mechanics and Foundation Engineering, 1, Butterworths Scientific

Publications, London, 207-210.

Wood, D.M. (1982). “Cone penetrometer and liquid limit.” Geotechnique, 32(1), 152-

157.

Wood, D.M. (1985). “Some fall-cone tests.” Geotechnique, 35(1), 64-68.

Wroth, C.P., and Wood, D.M. (1978) “The correlation of index properties with some

basic engineering properties of soils.” Canadian Geotech. J., 15 (2), 137-145.

Yong, R.N., and Japp, R.D. (1967). “A flow law for in dynamic compression.” Proc.

International Symposium on Wave Propagation and Dynamic Properties of Earth

Materials, University of New Mexico Press, Albuquerque, N.M., 183-188.

Yu, H.S., Herrmann, L.R. and Boulanger, R.W. (2000). “Analysis of steady cone

penetration in clay.” J. Geotech. Geoenviron. Eng. ASCE, 126 (7), 594-605.

157

Zeevart, L. (1948). “Discussion on effect of driving piles into soft clay.” Transactions,

ASCE, 115, 286-292.

Zreik, D.A., Ladd, C.C., and Germaine, J.T. (1995). “A new fall cone device for

measuring the undrained strength of very weak cohesive soils.” Geotech. Testing J.,

ASTM, 18(4), 472-482.