141
AN ABSTRACT OF THE THESIS OF Nason J. McCullough for the degree of Master of Science in Civil Engineering presented on February 23, 1998. Title: The Seismic Vulnerability of Sheet Pile Walls. Abstract approved: Stephen E. Dickenson The seismic performance of port structures has been well documented following recent earthquakes, and indicates that port structures are highly susceptible to earthquake- induced damages. These damages are primarily due to soil liquefaction and the associated ground failures. Sheet pile bulkheads provide vital intermodal and lifeline transportation links between water-side and land-side traffic, and are waterfront structures particularly vulnerable to liquefaction-induced damages. Due to the prevalence of liquefaction- induced damages, many ports are utilizing soil improvement techniques to mitigate these hazards. Many port authorities have proposed utilizing performance-based design criteria to limit potential earthquake-induced damages. The current design method for sheet pile walls (Mononobe-Okabe) is based on simple, limit equilibrium analysis techniques, which are poorly suited for performance-based design. Recent advancements in the seismic design of sheet pile walls have addressed some of the limitations of the current design methods, but are still inadequate for performing a complete, performance-based Redacted for privacy

Nason J. McCullough for the degree of Master of Science in

  • Upload
    others

  • View
    1

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Nason J. McCullough for the degree of Master of Science in

AN ABSTRACT OF THE THESIS OF

Nason J. McCullough for the degree of Master of Science in Civil Engineering

presented on February 23, 1998. Title: The Seismic Vulnerability of Sheet Pile Walls.

Abstract approved:

Stephen E. Dickenson

The seismic performance of port structures has been well documented following

recent earthquakes, and indicates that port structures are highly susceptible to earthquake-

induced damages. These damages are primarily due to soil liquefaction and the associated

ground failures. Sheet pile bulkheads provide vital intermodal and lifeline transportation

links between water-side and land-side traffic, and are waterfront structures particularly

vulnerable to liquefaction-induced damages. Due to the prevalence of liquefaction-

induced damages, many ports are utilizing soil improvement techniques to mitigate these

hazards.

Many port authorities have proposed utilizing performance-based design criteria to

limit potential earthquake-induced damages. The current design method for sheet pile

walls (Mononobe-Okabe) is based on simple, limit equilibrium analysis techniques,

which are poorly suited for performance-based design. Recent advancements in the

seismic design of sheet pile walls have addressed some of the limitations of the current

design methods, but are still inadequate for performing a complete, performance-based

Redacted for privacy

Page 2: Nason J. McCullough for the degree of Master of Science in

design for locations that contain potentially liquefiable soils and/or where soil

improvement strategies have been instituted.

This study has focused on conducting an empirical investigation and numerical

modeling to determine the seismic performance of sheet pile walls, and the performance-

based benefit of soil improvement through densification. A case history validated,

nonlinear effective stress computer program was used to perform numerical parametric

studies on various design parameters (earthquake properties, depth of sheet pile

embedment, sheet pile wall stiffness, tie rod length, density of the backfill, and extent of

soil densification). The results have been presented as a performance-based design

method, and include a design chart that provides practitioners with a preliminary design

tool that may be used to estimate the seismic deformations of sheet pile walls with or

without soil improvement.

The study has demonstrated that soil densification can greatly reduce the seismically-

induced deformations, especially when the magnitude of soil improvement extends

beyond the location of the anchor. The study has also demonstrated that the use of soil

densification techniques for mitigating seismic hazards may not be adequate in limiting

deformations to allowable limits, and that other methods of soil improvement

(cementation, drainage, etc.) or structural improvements may also be required.

Page 3: Nason J. McCullough for the degree of Master of Science in

C Copyright by Nason J. McCullough

February 23, 1998

All Rights Reserved

Page 4: Nason J. McCullough for the degree of Master of Science in

THE SEISMIC VULNERABILITY OF SHEET PILE WALLS

by

Nason J. McCullough

A THESIS

submitted to

OREGON STATE UNIVERSITY

in partial fulfillment of

the requirements for the

degree of

MASTER OF SCIENCE

Presented February 23, 1998

Commencement June 1998

Page 5: Nason J. McCullough for the degree of Master of Science in

Master of Science thesis of Nason J. McCullough presented on February 23, 1998

APPROVED:

Major Professor, representing Civil Engineering

Head or Chair:4Department of Civil, Construction, and Environmental Engineering

Dean of Gradu School

I understand that my thesis will become part of the permanent collection of Oregon StateUniversity libraries. My signature below authorizes release of my thesis to any readerupon request.

Nason J. McCullough, Autho

Redacted for privacy

Redacted for privacy

Redacted for privacy

Redacted for privacy

Page 6: Nason J. McCullough for the degree of Master of Science in

ACKNOWLEDGMENT

The completion of this project would not have been possible without the support and

guidance of my major professor, Stephen Dickenson. My minor professor, Charles Sollitt,

has also provided much support and guidance during my academic pursuits.

I am indebted to my parents, who have supported and encouraged me in all of my

endeavors, my sister who has always been a friend, and finally, and most importantly my

wife Amelia, who I love tremendously, and is the source of my joy and happiness.

I wish to extend my sincere gratitude to Dr. Susumu Iai (Director) and Mr. Koji Ichii

(Technical Official), both with the Geotechnical Earthquake Engineering Laboratory, Port and

Harbour Research Institute of the Japan Ministry of Transport. Many aspects of this research

would not have been possible without their assistance with the case studies and valuable

insights on Japanese design and construction of anchored bulkheads.

This work was supported by grant no. NA36RG0451 (project no. R/CP-33) from the

National Oceanic and Atmospheric Administration to the Oregon State University Sea Grant

College Program and by appropriations made by the Oregon State legislature. The views

expressed herein are those of the author and do not necessarily reflect the views of NOAA or

any of its subagencies. The author gratefully acknowledges the support of NOAA and Oregon

Sea Grant.

Page 7: Nason J. McCullough for the degree of Master of Science in

ii

TABLE OF CONTENTS

Page

1 INTRODUCTION 1

1.1 Background 1

1.2 Statement of Objectives and Scope of Work 10

1.2.1 Objectives 101.2.2 Report Organization 12

2 LIQUEFACTION ASSESSMENT OF WATERFRONT SOILS 13

2.1 Introduction 13

2.2 Cyclic Shear Method of Liquefaction Assessment 14

2.2.1 Determination of the Earthquake Induced Cyclic Shear Stresses 142.2.2 Cyclic Shear Stress Required to Initiate Liquefaction 172.2.3 Evaluation of Initiation of Liquefaction 242.2.4 Effects of Liquefaction 24

3 MITIGATION OF LIQUEFACTION HAZARDS 28

3.1 Introduction 28

3.2 Techniques for Mitigating Liquefaction Hazards 28

3.3 Design of Soil Mitigation 30

3.4 Design for the Area of Soil Mitigation 33

4 SEISMIC ANALYSIS AND DESIGN OF SHEET PILE BULKHEADS 35

4.1 Introduction 35

4.2 Pseudo-Static Design 36

4.2.1 Non-Liquefiable Soil 404.2.2 Liquefiable Soil 43

Page 8: Nason J. McCullough for the degree of Master of Science in

iii

4.3 Recent Advancements in Flexible Wall Design 45

4.3.1 Non-Liquefiable Soils 454.3.2 Liquefiable Soils 48

4.4 Discussions of Sheet Pile Bulkhead Design Methods 50

5 NUMERICAL MODELING 52

5.1 Constitutive Soil Model 53

5.2 Pore Pressure Generation 55

5.3 General Modeling Parameters 57

5.3.1 Modeling of Soil Elements 575.3.2 Modeling of Structural Elements 595.3.3 Modeling of the Earthquake Motion 605.3.4 Modeling of the Water 605.3.5 Boundary Conditions 61

5.4 Validation of Numerical Model 61

5.4.1 Akita Port 625.4.2 Ishinomaki Port 725.4.3 Kushiro Port 795.4.4 Discussion of Numerical Validation Results 91

6 PARAMETRIC STUDY ON THE SEISMIC PERFORMANCE OF SHEETPILE BULKHEADS 93

6.1 Introduction 93

6.2 Depth of Sheet Pile Embedment 96

6.3 Tie Rod Length 96

6.4 Sheet pile Stiffness 97

6.5 Penetration Resistance of the Backfill 99

6.6 Soil Improvement 101

Page 9: Nason J. McCullough for the degree of Master of Science in

iv

6.7 Comparison of the Parametric Study with an Existing Deformation-BasedDesign Method 102

6.8 Development of a New Seismic Design Method 103

6.9 Recommendation Design Procedures 107

6.10 Example Design Problem 108

7 SUMMARY AND CONCLUSIONS 112

BIBLIOGRAPHY 115

APPENDICES 122

APPENDIX A (Earthquake Motions Used in the Parametric Study) 123APPENDIX B (Table of Parametric Study Results) 125

Page 10: Nason J. McCullough for the degree of Master of Science in

LIST OF FIGURES

Figure Page

1.1 Lateral Displacements at: a) Shiohama Wharf, Akita Port following the1983 Nihonkai Chubu Earthquake (PHRI, 1997) and b) Kushiro Portfollowing the 1993 Kushiro Oki Earthquake (PHRI, 1993) 2

1.2 Settlements at: a) Ohama Wharf, Akita Port following the 1983 NihonkaiChubu Earthquake (PHRI, 1997) and b) U.S. Navy Port Facilities, Guamfollowing the 1993 Guam Earthquake (Vandani et al., 1994) 3

1.3 Disruption of Crane Operations at Guam, following the 1993 GuamEarthquake (Swan and Harris, 1993) 4

1.4 Typical Anchored Sheet Pile Bulkhead 5

1.5 Geotechnical Failures of Anchored Sheet Pile Walls; a) embedmentfailure; b) anchor failure; and c) global failure 7

1.6 Structural Failures of Anchored Sheet Pile Walls; a) anchor, tie rod,and/or whale system failure; and b) sheet pile wall failure 7

2.1 Range of rd Values for Different Soil Profiles (after Seed and Idriss,1982) 15

2.2 Earthquake Induced Shear Stresses (Seed and Idriss, 1982) 16

2.3 Typical Results of Cyclic Shear Tests (PHRI, 1997) 18

2.4 Empirical Relationship Between the Cyclic Stress Ratio InitiatingLiquefaction and (N/)60 Values for Silty Sands in M=7.5 Earthquakes(after Seed et al., 1979) 21

2.5 Relationship between a and Ka (Seed & Harder, 1990) 22

2.6 Relationship between the Effective Vertical Stress and Ka (Seed &Harder, 1990) 22

2.7 Proposed CPT-based liquefaction curves based on field CPT andliquefaction data (Stark & Olson, 1995) 23

2.8 Graphical Plot on the Evaluation of the Initiation of Liquefaction 24

2.9 Relationship Between Residual Excess Pore Pressure and Factor ofSafety Against Liquefaction for Level-Ground Sites (after Marcuson andHynes, 1990) 25

2.10 Estimation of Post-Liquefaction Volumetric Strain for Clean Sands (afterIshihara and Yoshimine, 1992) 26

Page 11: Nason J. McCullough for the degree of Master of Science in

2.11 Undrained Critical Strength Ratio vs. Equivalent Clean Sand Blow

vi

Count (Stark and Mesri, 1992) 27

3.1 Example Remediation Design for a Sheet Pile Wall (PHRI, 1997) 32

3.2 Improvement area for sheet pile walls (PHRI, 1997) 33

3.3 Improvement Area for Sheet Pile Wall Anchors (PHRI, 1997) 34

4.1 Relation between Seismic Coefficient and Ground Acceleration (Noda etal., 1975) 37

4.2 Seismic Coefficient Estimated from Damaged Quaywall (after Nozu etal., 1997) 37

4.3 Definition Sketch for Pseudo-Static Design Parameters 38

4.4 Definition Sketch for the Location of the Resultant Dynamic ActiveEarth Pressure 42

4.5 Definition of Effective Anchor Index, EAI (after Gazetas et al., 1990) 47

4.6 The Developed Seismic Design Chart (Gazetas et al., 1990) 48

5.1 Basic Explicit Calculation Cycle 53

5.2 Modeled Liquefaction Resistance Curve 55

5.3 Plan of Akita Port, Japan after the 1983 M 7.7 Earthquake (Iai &Kameoka, 1993) 63

5.4 Akita Ohama Wharf 1 Wall Profile (Iai & Kameoka, 1993) 64

5.5 Akita Ohama Wharf 2 Wall Profile (Iai & Kameoka, 1993) 64

5.6 Deformed Cross-Sections for Akita Ohama Wharf 2 65

5.7 Soil Profile and Grain Size for Ohama Wharf 1 (Iai & Kameoka, 1993) 65

5.8 Soil Profile and Grain Size for Ohama Wharf 2 (Iai & Kameoka, 1993) 66

5.9 Shear Wave Velocity Profile for Ohama Wharf 3 (Iai & Kameoka, 1993) 67

5.10 Soil and Velocity Profile at the Akita Strong Motion Site (Iai &Kameoka, 1993) 68

5.11 Recorded Time Histories for the 1983 M 7.7 Nihonkai ChubuEarthquake at Akita Port 69

5.12 Input Time History for Akita Ohama Wharf 1 69

5.13 Input Time History for Akita Ohama Wharf 2 70

5.14 Deformed Grid for Akita Ohama Wharf 1 (deformations to scale) 70

5.15 Time Displacement Plot for Akita Ohama Wharf 1 71

5.16 Deformed Grid for Akita Ohama Wharf 2 (deformations to scale) 72

Page 12: Nason J. McCullough for the degree of Master of Science in

vii

5.17 Time Displacement Plot for Akita Ohama Wharf 2 72

5.18 Measured a) Horizontal and b) Vertical Displacements along the Face ofthe Bulkhead (Iai & Kameoka, 1993) 73

5.19 Layout of Ishinomaki Port during the M 7.4 Miyagi-Ken-Oki Earthquake(Iai et al, 1985) 74

5.20 Shiomi (-4.5m, Site D) Bulkhead Profile (Iai et al., 1985) 74

5.21 Boring logs at Shiomi Wharf (Iai et al., 1985) 75

5.22 Recorded Time Histories at Ishinomaki Port 77

5.23 Input Time History for Ishinomaki Wharf 77

5.24 Deformed Grid for Ishinomaki, Shiomi Wharf (deformations to scale) 78

5.25 Time Displacement Plot of Ishinomaki, Shiomi Wharf 79

5.26 Measured Horizontal Deformations at Shiomi Wharf (-4.5 m),displacements in mm (Iai et al., 1985) 80

5.27 Measured Vertical Displacements at Shiomi Wharf (-4.5 m) (Iai et al.,1985) 80

5.28 Location of Kushiro Port (Iai et al., 1994) 81

5.29 Layout of Kushiro Port (Iai et al., 1994) 81

5.30 Geometry of Kushiro Port Site C Wharf (Iai et al., 1994) 82

5.31 Geometry of Kushiro Port Site D Wharf (Iai et al., 1994) 82

5.32 Soil Profile at Kushiro Port Site C (Iai et al., 1994) 83

5.33 Soil Profile at Kushiro Port Site D (Iai et al., 1994) 84

5.34 Soil Profile at the Kushiro Port Strong Motion Site (Iai et al., 1994) 86

5.35 Earthquake Motions Recorded at the Ground Surface at Kushiro Port onJanuary 15, 1993 87

5.36 Earthquake Motions Recorded at a depth of 77 meters at Kushiro Port onJanuary 15, 1993 88

5.37 Input Time History for Kushiro Port Site C and Site D 88

5.38 Kushiro Site C Deformed Grid (deformations to scale) 89

5.39 Time Displacements for Kushiro Site C 89

5.40 Kushiro Site D Deformed Grid (deformations to scale) 90

5.41 Time Displacements for Kushiro Site D 90

6.1 Top of Wall Displacements for the Depth of Embedment Study 97

Page 13: Nason J. McCullough for the degree of Master of Science in

viii

6.2 Top of Wall Displacements for the Tie Rod Length Study 98

6.3 Top of Wall Displacements for the Sheet Pile Stiffness Study 99

6.4 Top of Wall Displacements for the Penetration Resistance of the BackfillStudy 100

6.5 Definition of Soil Improvement Variables 101

6.6 Horizontal Displacments at the Top of the Wall for Variations in theZone of Soil Improvement 102

6.7 The Developed Seismic Design Chart (Gazetas et al., 1990), IncludingData from the Parametric Study for Models with Improved Soils (solidstars) 104

6.8 Design Chart for Lateral Displacements 106

6.9 Anchored Sheet Pile Bulkhead Design Problem 108

6.10 Lateral Displacements for the Design Problem 111

Page 14: Nason J. McCullough for the degree of Master of Science in

ix

LIST OF TABLES

Table

Selected Historical Earthquake Damage to Anchored Sheet Pile

Page

1.1

Bulkheads 6

2.1 Values of CSR Correction Factor, c,. (after Seed, 1979) 18

2.2 Magnitude Scaling Factors, MSF (Arango, 1996) 21

2.3 Recommended Fines Correction for Estimation of Residual UndrainedStrength (Stark & Mesri, 1992) 27

3.1 Liquefaction Remediation Measures (after Ferritto, 1996) 29

4.1 Qualitative and Quantitative Description of the Reported Degrees ofDamage (after Kitajima and Uwabe, 1979) 47

5.1 Comparison of FLAC and Finite Element Numerical Programs 54

5.2 Model Soil Properties for Ohama Wharf 1 and 2 66

5.3 Model Structural Properties for Ohama Wharf 1 and 2 68

5.4 Model Soil Properties for Ishinomaki Port 75

5.5 Model Structural Properties for Ishinomaki Port 76

5.6 Model Soil Properties for Kushiro Port 84

5.7 Model Structural Properties for Kushiro Port 85

5.8 Summary of Case Study Results 92

6.1 Bulkhead Properties for Parametric Study 94

6.2 Parametric Study Earthquake Motions 95

6.3 Plotted Case History Validations 106

6.4 Magnitude Scaling Factors for Design Example 109

6.5 Lateral Displacements for Design Example 110

Page 15: Nason J. McCullough for the degree of Master of Science in

THE SEISMIC VULNERABILITY OF SHEET PILE WALLS

1 INTRODUCTION

1.1 BACKGROUND

Experience at ports has demonstrated that waterfront earth-retention structures are

highly susceptible to earthquake-induced damage. A significant percentage of the damage

is due to the liquefaction of adjacent soils. Liquefaction susceptible soils consist of loose,

saturated, non-cohesive soils, and are frequently found in the port environment. These

soils are prevalent in the port environment due to hydraulic backfilling of waterfront

retaining structures and/or loose sediments deposited in alluvial and marine

environments. Earthquake-induced bulkhead failures due only to large seismic inertia

forces have also been noted in locations that were reported as not experiencing

liquefaction (Werner and Hung, 1982), although it has also been noted that the occurrence

of liquefaction at depth may be masked on the ground surface by large wall deformations

(Towhata et al., 1996).

Earthquake-induced damage to waterfront retaining structures is usually manifested

as excessive lateral displacements, and/or settlements, with several recent examples

presented in Figure 1.1 and Figure 1.2. There are also many cases of widespread damage

to backland structures (e.g., gantry cranes, cargo handling components, buildings, bridges,

pavements, etc.) that are a result of the lateral displacements and settlements, as seen in

Figure 1.3 with the spreading of the gantry crane rails. The economic losses associated

with damaged port structures and the suspension of port operations due to earthquake

induced damages has been substantial. For example, the Port of Oakland following the

1989 Mw 6.9 Loma Prieta Earthquake, had repair costs estimated at $75 million (Werner,

1990). The U.S. Navy Facilities suffered a total of approximately $275 million in direct

Page 16: Nason J. McCullough for the degree of Master of Science in

2

.7.74,..74.441,..;,.., . *o....41, It y0% ,Ac".or-cs,.... ,.4,1 kel,..1,,,,.., - ,Al..,,,C,.'1-71v 0, +A, ..10.,,..-v. Il..,,V.". '.'"Ve.

+,..,,..,14,..t.,' ' ''..'''."'"*."."" t.,,...,,,"........,,,fi

:#,V...1,41.1 E'. V, V t,;:..1..i*%:.t=4,tgIt'';:.;:''',...v,.1.tYr" .. ,.,O '..-,.1, ,...lo . .1 , 0 411 .1.1-r., .... ,...,,,,,,,,, lf 1

J

r 'il ....M. A-44.4 ,, ' ',.:':tilr7,;"7.2;ia.,,,, *44 , .^ . . 4.:111-'-'1;;:1VittiHtriCr4,417 ' ,. , ,I 1,1; ....r.,, , )4{1.'.:Ii'

1

OA* r ,Ite 111401,64..,..4a11,14.,l011.r4.;111;1;01=.41

':, '.1:It('tt1::::"4;1121.1',7o"'g

- 'µ.ms

13)

Figure 1.1: Lateral Displacements at: a) Shiohama Wharf, AkitaPort following the 1983 Nihonkai Chubu Earthquake(PHRI, 1997) and b) Kushiro Port following the 1993Kushiro Oki Earthquake (PHRI, 1993)

Page 17: Nason J. McCullough for the degree of Master of Science in

b)

a)

Figure 1.2: Settlements at: a) Ohama Wharf, Akita Port followingthe 1983 Nihonkai Chubu Earthquake (PHRI, 1997)and b) U.S. Navy Port Facilities, Guam following the1993 Guam Earthquake (Vandani et al., 1994)

3

Page 18: Nason J. McCullough for the degree of Master of Science in

4

Figure 1.3: Disruption of Crane Operations at Guam, followingthe 1993 Guam Earthquake (Swan and Harris, 1993)

losses during the Loma Prieta Earthquake and the 1993 Mw 8.1 Guam Earthquake

(Ferritto, 1997). The U.S. Navy damage from both earthquakes was noted to be due

primarily to liquefaction (Ferritto, 1997). Another example of commercial port damage is

to the Port of Kobe following the 1995 Mw 6.9 Hyogoken Nanbu Earthquake, with repair

costs estimated at $5.5 billion (CKPHB, 1997), and the costs due to port downtime during

the first nine months following the earthquake estimated at $6 billion (PHRI, 1996).

Anchored sheet pile bulkheads (Figure 1.4) are particularly vulnerable to earthquake-

and liquefaction-induced damage. Seismic damage to anchored sheet pile bulkheads

following numerous international earthquakes has been well documented by Kitajima and

Uwabe (1979), Werner and Hung (1982), and Werner (1998). Table 1.1 provides a

cursory overview of earthquake damage to anchored sheet pile bulkheads, and also

includes a note on the observed occurrence of liquefaction. It is readily apparent from the

table that for the majority of anchored sheet pile bulkheads subjected to medium- to high-

Page 19: Nason J. McCullough for the degree of Master of Science in

5

intensity earthquake motions, liquefaction occurred, and was likely the primary cause of

the reported damage.

dredge line

waleAl

H

=303

tie rod

backfill soil

.).3z.

anchor

Figure 1.4: Typical Anchored Sheet Pile Bulkhead

The damaging effects of earthquakes (for cases with or without liquefaction) include

increased active pressures due to the loss of soil strength and the seismic inertia of the

backfill, and the loss of passive soil resistance adjacent to the toe of the wall beneath the

dredge line and in front of the anchor. The increase in active pressures and decrease in

passive pressures leads to possible geotechnical failures including: a) the loss of

embedment resistance, b) the loss of anchor resistance, and/or c) a deep seated global

failure for bulkheads situated on weak foundation soils (Figure 1.5). Possible structural

failures of sheet pile walls include exceeding the yield capacity of the a) tie rod, tie rod

connections, wale system; or b) sheet pile wall (Figure 1.6). A structural interlock failure

between the sheet pile sections is also possible. The geotechnical and/or structural

failures can produce lateral deformations and/or settlements in the backland.

Page 20: Nason J. McCullough for the degree of Master of Science in

6

Table 1.1: Selected Historical Earthquake Damage to AnchoredSheet Pile Bulkheads

Earthquake Date Magnitude Port LateralMovement

LiquefactionNoted?

Tonankai, Japan Dec 7, 1944 8.3NagoyaOsaka

4 m3 m

YesYes

Nankai, Japan Dec 21, 1946 8.1NagoyaOsaka

4 m3 m

YesYes

Chile May 22, 1960 8.4 Puerto Montt 1 m Yes

Alaska, USA Mar 27, 1964 8.4 Whittier Yes

Niigata, Japan Jun 16, 1964 7.5 Niigata +2 m Yes

Tokachi-Oki, Japan May 15,1968 7.8HachinoheHakodate

0.9 m0.6 m

NoYes

Nemuro-Hanto-Oki,Japan

Jun 17, 1973 7.4HanasakiKiritappu

2 mnegligible

YesYes

Miyagi-Ken-Oki,Japan

Jun 12, 1978 7.4IshinomakiYuriageSendai

1.2 m1.2 m

negligible

YesYesYes

Nihonkai- Chubu,Japan

May 26, 1983 7.7 Akita 1.8 m Yes

Kushiro-Oki, Japan Jan 15, 1993 7.8 Kushiro 0.6 m Yes

Guam Aug 8, 1993 8.1Cabras IslandApra Harbor

0.6 m0.6 m

YesYes

Limiting earthquake-induced deformations of waterfront retaining walls is a primary

seismic design issue at many ports, and subsequently, many ports have adopted

deformation-based seismic performance requirements. For example, the following

guidelines have been proposed for U.S. Naval facilities (Ferritto, 1997);

Design of anchored sheet pile retaining walls shall limit permanent displacement at the topof the sheet pile to the following;

1) Less than 2.5 cm for a Level 1 earthquake (50% probability of exceedance in 50 years)2) Less than 10 cm for a Level 2 earthquake (10% probability of exceedance in 50 years)

These standards are intended to insure that following a Level 1 (operating level)

earthquake motion, the earthquake-induced damages will be negligible and non-structural

Page 21: Nason J. McCullough for the degree of Master of Science in

a)

a)

c)

b)

7

Figure 1.5: Geotechnical Failures of Anchored Sheet Pile Walls; a)embedment failure; b) anchor failure; and c) globalfailure

b)

Figure 1.6: Structural Failures of Anchored Sheet Pile Walls; a)anchor, tie rod, and/or whale system failure; and b)sheet pile wall failure

and that the port operations will be unimpeded. They are also intended to insure that

following a larger, Level 2 (contingency level) earthquake motion, the damages will be

non-catastrophic and repairable.

Page 22: Nason J. McCullough for the degree of Master of Science in

8

In response to the liquefaction-induced damage that has occurred during recent

earthquakes, and the development of performance-based design requirements, many ports

are instigating programs to mitigate liquefaction hazards in waterfront areas. Common

remediation objectives for increasing the liquefaction resistance of soils include

densification, increased strength/stiffness, and/or improved soil drainage. These

improvement objectives are accomplished through many methods, including deep

dynamic compaction, vibro-compaction, stone columns, soil mixing, and many others.

Although the use of soil improvement methods is increasing, there are currently very few

tools available for designing the extent of ground treatment necessary to limit the

earthquake-induced damage of waterfront retaining structures. The only comprehensive

reference known to the author is provided by the Japanese Port and Harbour Research

Institute (PHRI, 1997), which provides guidance on soil improvement strategies at

waterfront facilities. The recommendations provided in this useful reference are largely

based on limit state analysis and model testing. The guidelines provided do not address

the wall deformations associated with varying design-level ground motions, a primary

concern in performance-based design.

Current "standard of practice" seismic design for anchored sheet pile bulkheads

involves using pseudo-static, limit equilibrium mechanics, developed for rigid retaining

walls by Okabe (1926) and Mononobe and Matsuo (1929). This method, which is based

on the Coulomb earth pressure theory, is frequently referred to as the Mononobe-Okabe

method. The seismic portion of the design is controlled by empirically determined

seismic coefficients, which are functions of the maximum ground accelerations. The

coefficients are used to estimate the seismically-induced inertial forces. Kitajima and

Uwabe (1979) have noted that even with increasing values of the design seismic

coefficients during the last fifty years, the percentage of earthquake-damaged sheet pile

bulkheads has not decreased. This could be due to several design factors, including:

inadequate maximum ground accelerations used in the wall design,

the use of maximum ground accelerations fails to account for the frequency and

duration of actual earthquake ground motions,

Page 23: Nason J. McCullough for the degree of Master of Science in

9

the relationships used to estimate the seismic coefficients from the maximum

accelerations may be deficient,

sheet pile walls are commonly designed using limit-equilibrium methods, but

most seismic failure criterion is based on permanent deformations,

sheet pile walls are flexible, but designed using rigid body mechanics, and

excess pore pressure generation and liquefaction is usually not accounted for,

therefore the actual active earth pressures may be much larger and the passive

earth pressures much smaller than the earth pressures used in design.

The Mononobe-Okabe method can be used to account for the presence of potentially

liquefiable soils, but only in a simplistic manner by decreasing the soil strength or

increasing the active earth pressures. The Mononobe-Okabe method is also restricted by

being limit-equilibrium based, and therefore not applicable for deformation-based

analysis.

There have been several recent additions to the Mononobe-Okabe seismic design

method. These methods may be divided into two categories; 1) procedures which include

the possibility of liquefaction, and 2) those that do not include the possibility of

liquefaction. The methods that do not include the possibility of liquefaction include

Dennehy (1985) and Gazetas et al. (1990); Neelakantan et al. (1992); and Steedman and

Zeng (1990). A method that attempts to account for liquefaction has been developed by

Towhata and Islam (1987). These recent methods have contributed to the knowledge of

seismic behavior and design of sheet pile bulkheads, but they are limited in their use as

design methods, as discussed in Section 4.4.

In summary, the prevalent issues for the deformation-based, seismic analysis of sheet

pile bulkheads includes the need to estimate lateral deformations for:

non-liquefiable soils,

potentially liquefiable soils,

varying design-level ground motions, and

bulkheads with partial soil improvement.

Page 24: Nason J. McCullough for the degree of Master of Science in

10

1.2 STATEMENT OF OBJECTIVES AND SCOPE OF WORK

1.2.1 Objectives

Primary objectives of this project were to determine the seismic vulnerability and

necessary remediation methods for flexible, waterfront retaining structures. Several

objectives were determined at the onset of this project that included; a) review the

extensive technical literature and establish a data base of case histories, b) identify

common failure modes and evaluate the applicability of current "standard of practice"

methods for the seismic design of anchored sheet pile bulkheads, c) perform numerical

soil-structure interaction studies, and d) develop design recommendations for the

performance-based seismic design of anchored sheet pile walls. The scope of work for

each of these objectives is outlined below.

1.2.1.1 Establish a Data Base of Case Histories

An extensive technical literature search was conducted to collect case histories on the

seismic performance of sheet pile bulkheads. The search was conducted intensely during

the first six months of the project, and to a lesser extent during the remainder of the

research. A review of the literature was used to evaluate the performance of sheet pile

bulkheads, and to determine the controlling variables used in design. Five of the case

histories collected during the literature search contained the necessary data for use in

validating the numerical model.

1.2.1.2 Evaluation of Current Methods

An extensive literature search was also conducted to determine the current "standard

of practice" design method, and the applicability of the design method against the failure

modes identified from the case histories. The search resulted in a literature collection of

Page 25: Nason J. McCullough for the degree of Master of Science in

11

the traditional methods used in the seismic design of sheet pile bulkheads, and recent

additions that account for limitations of the "standard of practice" design method.

1.2.1.3 Perform Analytical Soil-Structure Interaction Studies

Analytical soil-structure interaction studies were performed to model the behavior of

sheet pile retaining walls in liquefiable and improved soils. This objective utilized the

geomechanical-modeling program FLAC (Itasca Consulting Group, 1995). There were

two phases of this portion of the study, the first involved validating FLAC by modeling

case histories to assess the capabilities and accuracy of FLAC. Phase two utilized the

validated FLAC model to perform parametric studies of the seismic performance of sheet

pile bulkheads. The parametric studies were used to determine the influence of several

design and remediation factors, including:

embedment depth of the sheet pile wall,

length of the tie rod,

stiffness of the sheet pile wall,

penetration resistance of the backfill,

zone of soil densification, and

ground motion characteristics.

1.2.1.4 Development of Design Recommendations

The results of the analytical numerical analyses were used to develop an improved

seismic design procedure that incorporates all of the parametric study data, and includes

field performance data incorporated from the literature review. Design recommendations

were also developed that highlight the usage of a simple design chart and, importantly,

the assumptions and limitations of using the chart for design purposes.

Page 26: Nason J. McCullough for the degree of Master of Science in

12

1.2.2 Report Organization

The remaining chapters are organized as follows. Chapter 2 highlights the necessary

aspects of performing a liquefaction assessment for waterfront structures, and includes

aspects on field exploration, laboratory testing and liquefaction assessment. Chapter 3

highlights the current methods available and commonly used in soil improvement for

liquefaction remediation. Chapter 4 details the current design method for sheet pile

bulkheads, and also includes summaries of recent contributions to the design of sheet pile

bulkheads. Chapter 5 deals with the aspects of the numerical model used in the analytical

studies, and includes the results of the validation case histories. Chapter 6 presents the

results of the numerical parametric study and the recommended design procedures.

Chapter 7 provides a summary and conclusion of the research project and

recommendations for future work.

Page 27: Nason J. McCullough for the degree of Master of Science in

13

2 LIQUEFACTION ASSESSMENT OF WATERFRONT SOILS

2.1 INTRODUCTION

There have been several extensive reports indicating that the most significant source

of earthquake damage to waterfront structures has been due to the liquefaction of loose to

medium-dense saturated sands (Kitajima and Uwabe, 1979; Werner and Hung, 1982;

Werner, 1998). Liquefaction is the result of cyclic shearing of loose to medium-dense,

saturated granular soils. To determine the liquefaction susceptibility of soils, a

liquefaction assessment needs to be conducted. The common simplified method used to

perform a liquefaction assessment in the United States is the cyclic shear method, and is

described in the following sections of this chapter. Prior to the liquefaction assessment, it

is necessary to first complete the following:

1) Perform a Geotechnical/Geological Exploration to provide an assessment

of the soil types, stratigraphy, location of the water table, site profile and

geometry, and hydraulic conditions. It may also necessary, depending on the

liquefaction assessment analysis, to perform in-situ field testing (SPT, CPT, shear

and pressure wave velocities, etc.) and collect samples for laboratory testing.

2) Determine the Soil Engineering Properties through laboratory tests and

stress calculations. Laboratory tests for liquefaction assessment are conducted to

determine one or more of the following; density, cyclic shear resistance, damping

characteristics, soil structure, dynamic shear modulus, and grain size. It is also

necessary to determine the static soil stresses (vertical, horizontal and shear) with

depth. The sample disturbance that takes place during sample collecting,

transportation, and laboratory testing needs to be acknowledged and accounted

for in the interpretation of the test results. Descriptions on the determination of

soil engineering properties can be found in USACE (1984), Terzaghi et al.

(1996), and PHRI (1997).

Page 28: Nason J. McCullough for the degree of Master of Science in

14

3) Determine the Earthquake Motion(s) using either simple estimation

procedures that are prescribed by regulatory bodies (e.g. FEMA, UBC, etc.), or

perform a site-specific deterministic or probabilistic seismic hazard analysis. A

complete description of these methods can be found in Dickenson et al. (1998)

and Kramer (1996).

2.2 CYCLIC SHEAR METHOD OF LIQUEFACTION ASSESSMENT

Several steps are involved in using the cyclic shear method for liquefaction

assessment; 1) determine the earthquake induced cyclic shear stresses, 2) determine the

cyclic shear stress required to initiate liquefaction, and 3) compare the results of steps 1

and 2 to assess the liquefaction susceptibility of the soil. These steps are described in

Sections 2.2.1 to 2.2.3. A short discussion on the estimation of liquefaction-induced

effects (residual excess pore pressure, post-liquefaction volumetric strain, and undrained

residual shear strength) is also provided in Section 2.2.4.

2.2.1 Determination of the Earthquake Induced Cyclic Shear Stresses

The shear stresses developed at any location within a soil deposit during an

earthquake appear to be due primarily to the vertical propagation of shear waves. There

are two common methods for estimating the earthquake induced cyclic shear stresses;

1) perform a simple empirical analysis, or 2) perform a site specific ground response

analysis.

2.2.1.1 Simplified Empirical Analysis

The simplified procedure developed by Seed and Idriss (1971, 1982) uses the concept

of vertically propagating shear waves producing cyclic shear stresses within the soil

column. The first step involves estimating the maximum acceleration within the soil

column at depth z (am.,z) from the maximum ground surface acceleration (am,), using

Equation 2.1 and the data plotted in Figure 2.1.

Page 29: Nason J. McCullough for the degree of Master of Science in

15

Amax z = amax rd 2.1

a max,z

rda max

0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0

Figure 2.1: Range of rd Values for Different Soil Profiles (afterSeed and Idriss, 1982)

The empirical data plotted in Figure 2.1 indicates a wide range between the upper

and lower bound values, especially below a depth of approximately 12 m. It is

recommended that the average values (or for a conservative analysis, the upper bound

values) be used for depths between 0 and 12 m, and that a site specific dynamic ground

response analysis be conducted to estimate the maximum acceleration for depths

exceeding 12 m.

Page 30: Nason J. McCullough for the degree of Master of Science in

16

After the maximum acceleration values with depth are estimated, the earthquake-

induced cyclic shear stresses at the depth of interest (1-,,,,,) can be calculated using

Equation 2.2, where g is the acceleration due to gravity and o is the total overburden

stress. The value of rm,Z is then converted to an equivalent uniform cyclic shear stress

(ray) to account for the transient time history behavior of the shear stresses. Seed and

Idriss recommend reducing z-,, by 65% to account for the irregular loading (Figure 2.2).

The value of rav,z is then estimated using Equation 2.3. The value of ra,,, is then

normalized by the effective overburden stress (o, ') to produce the cyclic stress ratio

induced by the earthquake (CSReq), which is given in Equation 2.4.

MaXZamax, z

VOg

Tmax

T 0.65 Tav max

TIME

Figure 2.2: Earthquake Induced Shear Stresses (Seed and Idriss,1982)

1- = 0 65 z-QV, Z MaXZ

CSR = av'z

avo

2.2

2.3

2.4

Page 31: Nason J. McCullough for the degree of Master of Science in

17

2.2.1.2 Site Specific Analysis

A more in-depth method of estimating the CSReq involves calculating anza,z or i-max,z

directly using a dynamic ground response analysis. A ground response analysis is

conducted using a numerical computer model, such as the equivalent linear program

SHAKE (Schnabel et al., 1972) or a fully nonlinear dynamic program, such as DESRA-

2C (Lee and Finn, 1991) or SUMDES (Li et al., 1992). The advantages and disadvantages

of each program should be thoroughly examined before use.

2.2.2 Cyclic Shear Stress Required to Initiate Liquefaction

The cyclic shear stress ratio required to initiate liquefaction (CSRfidd) can be

determined from either laboratory or field tests. Each method is discussed in the

following sections, along with the advantages and disadvantages of each method.

2.2.2.1 Laboratory Tests

Laboratory tests are used to determine the number of loading cycles necessary to

produce liquefaction failure (N1) for various amplitudes of shear stress. Liquefaction

failure is usually defined as the point at which liquefaction is initiated, or a set amount of

strain is exceeded. It should be noted that the denser the material, the larger the values of

N1. Typical results of cyclic shear tests are shown in Figure 2.3. There are two different

laboratory tests commonly conducted to determine the cyclic stress ratio; the cyclic

simple shear and the cyclic triaxial shear. The cyclic shear strengths obtained from either

the simple shear or triaxial shear tests are normalized by the effective overburden

pressure to produce CSRSS and CSR,x, respectively. Because the cyclic simple shear and

the cyclic triaxial shear tests impose different loadings, the CSRSS and CSR,x are not

equivalent, but are related by the following equation;

CSRSS = Cr CSRIx 2.5

Page 32: Nason J. McCullough for the degree of Master of Science in

18

where recommended values of cr have been compiled in Table 2.1 as a function of the

static coefficient of earth pressure (K0).

0.4CC

w,.6, 0 . 3

cr,

uj 0.2-JUI>- (1)

0 . 1

0

0 ISHINOMAKI PORT SAND Dr 50.0%

HANASAIG PORT SAND Dr = 38.5%

0 KUSHIRO PORT SAND Di 51.3%

3 10 30 100

NUMBER OF WAVES NI

300

Figure 2.3: Typical Results of Cyclic Shear Tests (PHRI, 1997)

Table 2.1: Values of CSR Correction Factor, c (after Seed, 1979)

Reference Equation

Cr

Ko = 0.4 Ko = 1.0

Finn et al. (1971)1+ K

0.7 1.0Cr =2

Seed and Peacock (1971) varies 0.55 0.72 1.0

Castro (1975)2(1+ 21C0)

0.69 1.15cr =315

The cyclic simple shear and cyclic triaxial shear produce shear stresses in only one

direction, in contrast to the actual earthquake motions, which produce shear stresses in

different directions simultaneously. Boulanger and Seed (1995) conclude from laboratory

tests that the CSR for unidirectional simple shear tests was approximately 5 to 25% less

than multidirectional earthquake shaking. This is consistent with the recommendation of

Page 33: Nason J. McCullough for the degree of Master of Science in

19

Seed (1979), that the CSR for unidirectional simple shear tests be approximated as 10%

less than that for multidirectional shaking. Utilizing the work of these researchers, it is

concluded that multidirectional shaking reduces the liquefaction resistance of soils by an

average of approximately 10%. Therefore, the recommended laboratory CSR values are

related to the field CSR values by;

CSRfield = 0.9 . CSR,s = 0.9.c, CSRt, 2.6

Cyclic laboratory tests are normally conducted on re-constituted samples, due to the

extreme difficulty in obtaining relatively undisturbed samples of cohesionless soils. The

re-constituted cyclic laboratory tests are able to reproduce the in-situ density and effective

confining pressures, but due to the depositional and historical environment of many

waterfront areas, there are several conditions that cyclic tests are unable to reproduce

accurately. These conditions include the soil fabric, past earthquake loading, over-

consolidation ratio, lateral earth pressure coefficient, and the duration of sustained

pressure before testing. Relatively undisturbed sampling (e.g. in-situ ground freezing) is

necessary to confidently perform a liquefaction assessment using cyclic laboratory tests,

otherwise, caution should be taken in using the results of re-constituted laboratory

specimens.

2.2.2.2 Field Tests

The recommended method of characterizing the liquefaction resistance of a soil

deposit is based on the results of in-situ tests, due to the disturbance inherent in the

sampling and laboratory testing of cohesionless soils. The standard penetration test (SPT)

has historically been used for liquefaction assessments. Another method that is becoming

more common for liquefaction assessments is the cone penetration test (CPT). These

methods are discussed in detail in the following sections. Information on correlations

between other field test results (e.g. shear wave velocity, dilatometer) and the cyclic stress

ratio can be found in Kramer (1996).

Page 34: Nason J. McCullough for the degree of Master of Science in

20

Standard Penetration Test (SPT). Seed and Idriss (1982) noted that the specifics of

the SPT testing method have a very large influence on the measured results. Therefore, it

is necessary to standardize the measured results of the SPT test (N blows/30 cm). Seed et

al. (1985) provide a method for standardizing SPT results that has been used extensively

since it was first introduced. The resulting standardized SPT results are presented as

(A/1)60, in which the subscript 1 represents being normalized for an overburden pressure of

1 kg/cm2, and the subscript 60 represents being normalized for an applied energy of 60%

of the theoretical maximum free-fall energy.

Seed et al. (1979) plotted standardized SPT values versus the cyclic shear stress

ratios for earthquakes of magnitude 7.5 (CSRm=75) for locations with and without the

occurrence of liquefaction (Figure 2.4). This plot provides a very practical method of

estimating the cyclic stress ratio necessary to initiate liquefaction using normalized SPT

values and the percentage fines of the sandy soil. It should be noted that the CSRM=7.5

values in the figure are for earthquake magnitudes of 7.5 only, and for level sites that

have relatively shallow liquefaction susceptible soils. Because variations in earthquake

magnitudes lead to an increased number of earthquake induced shear cycles, the CSRA1=75

should be modified to account for different earthquake magnitudes. The CSRm=75 should

also be modified to account for the presence of large overburden pressures and for static

shear stresses (my). The CSRfieid can be calculated from the CSRm=75 with the following

equation;

CSRfield = Ka Ka MSF CSRA1,7.5 2.7

where Ka is the correction for the static shear stresses on horizontal planes in the soil

deposit (Figure 2.5), K0- is the correction for large overburden pressures (Figure 2.6), and

MSF is the correction for different magnitude earthquakes (Table 2.2). The MSF factor is

used to account for the variations in the duration (and corresponding cycles of loading)

between earthquake motions of different magnitude.

Page 35: Nason J. McCullough for the degree of Master of Science in

0.6

0.5

0.4

CSRm = 7.5 0.3

0.2

0.1

Percent fines = 35 15 <5

J I

A r

0

A

Fines content 25%Modified Chinese code proposal (clay content 5%)

Marginal NoLiquefaction Liquefaction Liquefaction

Pan-American data 0Japanese data 0Chinese data A A

00 10 20

(NOso

30 40

Figure 2.4: Empirical Relationship Between the Cyclic StressRatio Initiating Liquefaction and (A11)60 Values forSilty Sands in M=7.5 Earthquakes (after Seed et al.,1979)

Table 2.2: Magnitude Scaling Factors, MSF (Arango, 1996)

Earthquake Magnitude

5.50 6.00 7.00 7.50 8.00 8.25

MSF 3.00 2.00 1.25 1.00 0.75 0.63

50

21

Page 36: Nason J. McCullough for the degree of Master of Science in

K,

1.2

1.0

0.8

0.6

0.4

0.2

oC

rho

I

Figure 2.5: Relationship between a and KG, (Seed & Harder, 1990)

2.0 3.0 4.0 5.0 6.0 7.0 8.0

°0

\ 0\... N.

0

0 *X" 6 °..... o..4.1

...... 0 o

V ..... ,.....:,

.......0H.,..

FAIRMONT DAM N.. ......LAKE ARROWHEAD DAY ..,.

V SHEFFIELD DAM SHELL ...H., "...0 UPPER SAN LEANDRO DAM SHELL NH., ......V LONER SAN FERNANDO DAM SHELL ......,

C1 UPPER SAN FERNANDO DAM SHELLA LOS ANGELES DAM SHELL VV0

ERRiS DAM SHELL, AC 95,100%SARDIS DAM SHELL

0 SARDIS DAM FOUNDATION VTHERIAAL110 AFTERRAY DAY FOUNDATIONTHERMALITO FOREMAY DAM FOUNDATIONANTELOPE DAM IMPERVIOUS MATERIALFORT PECK DAM SHELL

0 SACRAMENTO RIVER SANG, D. SR, NO, TO, PO%MONTEREY 0 SAND, 0. SO%

0 REID BEDFORD SAND, Dr.. 40, O%0 NEw JERSEY 'MCNEILL, FPI RC OS%

0 1.0 2.0 3.0 4 0 5.0 6 0EFFECTIVE CONFINING PRESSURE (tsf ) or (kW

7.0

Figure 2.6: Relationship between the Effective Vertical Stress andK, (Seed & Harder, 1990)

8.0

22

Page 37: Nason J. McCullough for the degree of Master of Science in

23

Cone Penetration Test (CPT). The use of the CPT has increased dramatically in

recent years, due to several advantages over the SPT, including rapid and inexpensive

testing, and it also provides a continuous soil profile with depth. It is also possible to

enhance the CPT with the addition of pore pressure transducers and instruments to

measure the seismic wave velocity. The cone penetration resistance (qc) and the shaft

friction resistance (fs) are measured continuously during the testing. Stark and Olson

(1995) have proposed a relationship between liquefaction field data and CPT penetration

resistance (Figure 2.7). The seismic shear stress ratio is equivalent to CSRA,f=7.5. The value

of qa is the qc value has been normalized by a vertical effective overburden stress of

approximately 100 kPa. The CSRm=75 value can then be substituted into Equation 2.7 to

estimate the CSRfierd

0.6

0.5

0.4

0.3

02

0.1

00 5 10 15 20 25 30

CORRECTED CPT TIP RESISTANCE, cio (MPa)

Figure 2.7: Proposed CPT-Based Liquefaction Curves Based onField CPT and Liquefaction Data (Stark & Olson,1995)

Page 38: Nason J. McCullough for the degree of Master of Science in

24

2.2.3 Evaluation of Initiation of Liquefaction

After the cyclic stress ratio caused by the earthquake (CSReq) and the cyclic stress

ratio necessary to induce liquefaction (CSRfi ) are determined at the depths of interest,

the potential for liquefaction can be evaluated. A graphical representation is the simplest

method of determining the liquefaction potential, and is accomplished by plotting the

CSReq and CSRfierd versus depth (Figure 2.8). Liquefaction is likely to occur at any depth

where the CSReq exceeds the CSRfield A factor of safety against liquefaction (FS1) can also

be evaluated at specific depths with the following equation;

FS i=CSReq

CSR field

a)0

CSR

CSReq

Zone ofLiquefaction

CSRfiwd

Figure 2.8: Graphical Plot on the Evaluation of the Initiation ofLiquefaction

2.2.4 Effects of Liquefaction

2.8

The liquefaction of soils is caused by the generation of excess pore pressures, which

result in decreased effective stresses, soil stiffness and soil strength. Volumetric strains

due to the dissipation of generated excess pore pressures and consolidation of the

Page 39: Nason J. McCullough for the degree of Master of Science in

25

densified soil are also an effect of liquefaction. It should be noted that excess pore

pressures are still generated even with FS1 values in excess of one.

Marcuson and Hynes (1990) developed a figure relating the excess pore pressure

ratio (ru) to the factor of safety against liquefaction for both gravel and sand (Figure 2.9).

The excess pore pressure ratio is the relationship between the residual excess pore

pressure (uexcess) and the effective overburden pressure (cry° ') given by;

ru ='VO

U excess

1.0

0.8

0.6

0.4

0.2

010

;= Gravel

Sand

1:1:"Zzi sie,1.2 1.4 1.6 11 2.1.0 2.2

FACTOR OF SAFETY At:MIST L1OUEFAACTION. FSL

Figure 2.9: Relationship Between Residual Excess Pore Pressureand Factor of Safety Against Liquefaction for Level-Ground Sites (after Marcuson and Hynes, 1990)

2.4 '6

2.9

Several researchers have evaluated the post-liquefaction volumetric strain due to the

dissipation of excess pore pressures. An approach that utilizes the FS/ was developed by

Page 40: Nason J. McCullough for the degree of Master of Science in

26

Ishihara and Yoshimine (1992), and relates the factor of safety to the post-liquefaction

volumetric strain for clean sands of various relative densities (Dr), given in Figure 2.10.

2.0

1.8

1.6

1.4

1.2

FS, 1.0

0.8

0.6

0.4

0.2

.3

-3.5%

4%

6%

8%

( N, q

0, 60 =D, = 50

= 60= 80

/)

(NI = 20, gc, = no)

?max = 10% 13, = 80= 25,00 = 147)

Dr = 90%N, = 30

200kg cm2

D, =40N, = 6

qct = 45

Dr= 30N, = 3 .

qc, =

1,0 20 30 40Post-liquefication volumetric strain, Ev(%)

50

Figure 2.10: Estimation of Post-Liquefaction Volumetric Strain forClean Sands (after Ishihara and Yoshimine, 1992).

The most common method of estimating the residual undrained shear strength was

developed by Stark and Mesri (1992) and utilizes SPT results corrected for the percentage

of fines contained in the soil. The relationship between the clean sand SPT ((Aid 6o-cs) and

the undrained residual shear strength (normalized by the effective overburden pressure

Page 41: Nason J. McCullough for the degree of Master of Science in

27

and plotted as critical) is given in Figure 2.11. The clean sand SPT is calculated using the

data from Table 2.3 and the following equation;

(N1)60-CS = (N1)60 Ncorr

z

Table 2.3: Recommended Fines Correction for Estimation ofResidual Undrained Strength (Stark & Mesri, 1992)

0.6

0.5

0.4

0.3

0.2

0.1

Percent Fines Ncorr (blows/30 cm)0 0.0

10 2.515 4.020 5.025 6.030 6.5

35' 7.0

FIELD CASE HISTORIESO CYCUC TRIAXIAL CRITICAL STRENGTH0 CYCUC TRIAXIAL CRITICAL STRENGTH AT 100 CYCLES

CYCLIC TRIAXIAL AND BLOWCOUNT FROM EON. (6)v SIMPLE SHEAR AND TORSIONAL SHEAR TESTS

Su (YIELD. MOB)0.011 (N

1)6o_cS

CVOM7.5

:

Su (CRITICAL)0.0055 04 1

(:)" 1 60-CSVO0

4 8 12 16 20 24 28 32 36

EQUIVALENT CLEAN SAND SPT BLOWCOUNT, (N1)60 -CS

Figure 2.11: Undrained Critical Strength Ratio vs. EquivalentClean Sand Blow Count (Stark and Mesri, 1992)

40

2.10

Page 42: Nason J. McCullough for the degree of Master of Science in

28

3 MITIGATION OF LIQUEFACTION HAZARDS

3.1 INTRODUCTION

If a liquefaction assessment (Chapter 2) demonstrates that liquefaction is likely to

occur in the waterfront soils during the specified design earthquake, and the liquefaction-

induced damage exceeds the performance criteria, mitigation strategies should be

evaluated. Mitigation strategies should be designed to limit the earthquake-induced soil

displacements and settlements within acceptable levels. There are two categories of

mitigation, soil improvement and structural enhancement. Only the soil improvement

techniques are addressed herein.

3.2 TECHNIQUES FOR MITIGATING LIQUEFACTION HAZARDS

Common remediation objectives of soil improvement are to increase the liquefaction

resistance of soils through densification, increased strength, and/or improved soil

drainage. Table 3.1 presents the most common remediation methods, and the principle,

suitable soil conditions, effective treatment depth and relative costs for the tabulated

methods.

Many of the tabulated remediation methods have limited the occurrence of

liquefaction during recent earthquakes. Sand compaction piles, vibro-compaction, sand

drains and surcharge soil mitigation techniques were successful used in preventing

liquefaction in Kobe Port during the 1995 Hyogoken Nanbu Earthquake (Yasuda et al.,

1996). Sand compaction piles and gravel drains were successfully used in the Port of

Kushiro prior to the 1993 Kushiro Oki Earthquake (Iai et al., 1994). Hayden and

Boulanger (1995) describe the use of compaction grouting of liquefiable soils at a number

of different sites. Baez and Martin (1992) provide an evaluation on the use of stone

column (vibro-compaction) techniques for liquefaction mitigation. Egan and others

(1992) describe the installation of stone columns at the Port of Oakland for the mitigation

of liquefaction hazards.

Page 43: Nason J. McCullough for the degree of Master of Science in

29

Table 3.1: Liquefaction Remediation Measures (after Ferritto,1996)

Method PrincipleMost Suitable

Soil Conditionsor Types

MaximumEffective

Treatment Depth

RelativeCosts

1) VibratoryProbe

a) Terraprobe

b) Vibrorods

c) Vibrowing

Densification by vibration;liquefaction-induced settlementand settlement in dry soil underoverburden to produce a higherdensity.

Saturated or dryclean sand; sand.

20 m routinely(ineffective above 3-4 m depth); > 30 msometimes;vibrowing, 40 m.

Moderate

2) Vibro-compaction

a) Vibrofloat

b) Vibro-Composersystem.

Densification by vibration andcompaction of backfill materialof sand or gravel.

Cohesion lesssoils with lessthan 20% fines.

> 20 m Low tomoderate

3) CompactionPiles

Densification by displacement ofpile volume and by vibrationduring driving, increase inlateral effective earth pressure.

Loose sandy soil;partly saturatedclayey soil; loess.

> 20 m Moderateto high

4) Heavy tamping(dynamiccompaction)

Repeated application of high-intensity impacts at surface.

Cohesionlesssoils best, othertypes can also beimproved.

30 m (possiblydeeper)

Low

5) Displacement(compactiongrout)

Highly viscous grout acts asradial hydraulic jack whenpumped in under high pressure.

All soils. Unlimited Low tomoderate

6) Surcharge orbuttress

The weight of asurcharge/buttress increases theliquefaction resistance byincreasing the effectiveconfining pressures in thefoundation.

Can be placed onany soil surface.

Dependent on sizeofsurcharge/buttress

Moderateif verticaldrains areused

7) Drains

a) Gravel

b) Sand

c) Wick

d) Wells (forpermanentdewatering)

Relief of excess pore waterpressure to prevent liquefaction.(Wick drains have comparablepermeability to sand drains).Primarily gravel drains;sand/wick may supplementgravel drain or relieve existingexcess pore water pressure.Permanent dewatering withpumps.

Sand, silt, clay. Gravel and sand >30 m; depth limitedby vibratoryequipment; wick, >45 m

Moderateto high

8) Particulategrouting

Penetration grouting-fill soilpores with soil, cement, and/orclay.

Medium to coarsesand and gravel.

Unlimited Lowest ofgroutmethods

Page 44: Nason J. McCullough for the degree of Master of Science in

30

Table 3.1 (Continued)

Method PrincipleMost SuitableSoil Conditions

or Types

MaximumEffective

Treatment Depth

RelativeCosts

9) Chemicalgrouting

Solutions of two or morechemicals react in soil pores toform a gel or a solid precipitate.

Medium silts andcoarser.

Unlimited High

10) Pressureinjected lime

Penetration grouting-fill soilpores with lime

Medium to coarsesand and gravel.

Unlimited Low

11) Electrokineticinjection

Stabilizing chemical moved intoand fills soil pores by electro-osmosis or colloids in to poresby electrphoresis.

Saturated sands,silts, silty clays.

Unknown Expensive

12) Jet grouting High-speed jets at depthexcavate, inject, and mix astabilizer with soil to formcolumns or panels.

Sands, silts, clays. Unknown High

13) Mix-in-placepiles and walls

Lime, cement or asphaltintroduced through rotatingauger or special in-place mixer.

Sand, silts, clays,all soft or looseinorganic soils.

> 20 m (60 mobtained in Japan)

High

14) Vibro-replacementstone and sandcolumns

a) Grouted

b) Not grouted

Hole jetted into fine-grained soiland backfilled with denselycompacted gravel or sand holeformed in cohesionless soils byvibro techniques and compactionof backfilled gravel or sand. Forgrouted columns, voids filledwith a grout.

Sands, silts, clays. > 30 m (limited byvibratoryequipment)

Moderate

15) Root piles, soilnailing

Small-diameter inclusions usedto carry tension, shear, andcompression.

All soils. Unknown Moderateto high

16) Blasting Shock waves and vibrationscause limited liquefaction,displacement, remolding, andsettlement to higher density.

Saturated, cleansand; partlysaturated sandsand silts afterflooding.

> 40 m Low

3.3 DESIGN OF SOIL MITIGATION

The design of soil mitigation strategies involves investigating the cost/benefit ratio,

the performance, and the effect on adjacent structures of the mitigation technique(s). The

mitigation methods listed in Table 3.1 may be placed into one of three different

categories; compaction, drainage, and cementation. The design for each of these

Page 45: Nason J. McCullough for the degree of Master of Science in

31

remediation categories is briefly discussed below. A more in-depth discussion on the

design of soil mitigation strategies and procedures may be found in PHRI (1997).

Compaction. Compaction remediation methods mitigate liquefaction hazards through

soil densification with vibration or impact (examples include methods 1, 2, 3, 4 and 14

from Table 3.1). Compaction methods are more suitable for use in saturated, cohesionless

soils with a limited percentage of fines. Compaction remediation methods cause noise

and vibration during installation, and also increase horizontal earth pressures against

adjacent structures. The increase of horizontal earth pressures is the major disadvantage

of compaction methods in close proximity to retaining walls, and must be accounted for

during the design of a soil improvement program. The major advantage of compaction

methods for soil improvement is the relatively low cost/benefit ratio. The degree of

compaction that is necessary can be evaluated using penetration resistances that have

been back-calculated from an acceptable factor of safety against liquefaction (Chapter 2).

Drainage. Drainage remediation methods mitigate liquefaction hazards by enhancing

the rate of excess pore pressure dissipation. The most common methods of drainage

remediation are through the use of gravel, sand or wick drains. Drains are suitable for use

in sands, silts or clays. One of the greatest advantages of drains is that they induce

relatively small horizontal earth pressures during installation. Therefore, they are suitable

for use adjacent to sensitive existing structures. In the design of drains, it is necessary to

select a suitable drain material that has a coefficient of permeability substantially larger

than the in-situ soils.

Cementation. Cementation remediation methods mitigate liquefaction hazards

through increased soil strength. The soil strength is increased with the addition of acementatious material (i.e. cement, grout, lime, chemicals, asphalt). Cementation

techniques (methods 5, 8, 9, 10, 12, and 14 from Table 3.1) can be used with any type of

soil. Cementation methods are advantageous because the installation methods are

relatively quiet and induce relatively small vibrations compared with compaction

methods. The induced horizontal earth pressures are also less with cementation methods

than with compaction methods, but are greater than induced pressures from drainage

Page 46: Nason J. McCullough for the degree of Master of Science in

32

methods. The disadvantage of cementation methods is their relatively high cost/benefit

compared with compaction and drainage methods.

The relative performance of a specific improvement method is also of concern in the

design of a mitigation program. Experience has demonstrated that compaction and

cementation techniques reduce the liquefaction susceptibility of soils to a larger extent

than drainage methods.

The design of a mitigation strategy usually includes the combination of two or more

improvement techniques. An example is provided in Figure 3.1, in which a drainage

method was used adjacent to the sheet pile wall to limit the induced horizontal earth

pressures, and compaction techniques were used in front of the wall and behind the

anchor, likely to reduce the economic costs and for the better performance offered by

compaction methods in increasing the liquefaction resistance of the soil.

STEEL

17m

+4.00

H W.L* 2.00mL.W1,+ 0.00m

PIPE SHEET PILE0711. /

SEABED -corn

20m 25m

20.00m

23m

5

-21 .5m

Figure 3.1: Example Remediation Design for a Sheet Pile Wall(PHRI, 1997)

The influence on existing structures during soil improvement is a primary design

consideration. The construction of mitigation methods may lead to increased horizontal

earth pressures, which can result in increased bending stresses in sheet pile walls.

Mitigation methods may also induce excess pore pressures and vibration, which will also

Page 47: Nason J. McCullough for the degree of Master of Science in

33

affect the sheet pile wall and adjacent structures. Compaction techniques have the largest

influence on adjacent structures, followed by cementation and drainage methods, which

have the least amount of influence.

3.4 DESIGN FOR THE AREA OF SOIL MITIGATION

The Japan Port and Harbour Research Institute (PHRI, 1997) has produced one of the

few design guidelines that exist for specifying the extent of soil improvement adjacent to

sheet pile bulkheads. This recommended extent of treated soil is shown in Figure 3.2 for a

sheet pile wall and in Figure 3.3 for sheet pile wall anchors. The area of soil improvement

indicated was developed primarily to limit the effect of the propagation of excess pore

pressures. It was noted by PHRI (1997) that during shaking table tests, the area enclosed

by the 30° triangle exhibited unstable characteristics, and therefore led to the

recommended design guidelines.

IMPROVEMENT AREA

LIQUEFACTION

LAYER

NONLIQUEFACTION LAYER

Figure 3.2: Improvement area for sheet pile walls (PHRI, 1997)

There is much uncertainty in the effectiveness of varying ground treatment strategies

for limiting lateral deformations of sheet pile bulkheads due to the lack of data on the

performance of improved soil sites during design-level earthquakes. A limited amount of

research has been performed in evaluating seismically-induced lateral deformations of

Page 48: Nason J. McCullough for the degree of Master of Science in

34

improved soil sites. The guidelines provided by PHRI, though not deformation-based,

provide an applicable method of estimating the necessary extent of ground treatment that

is based on "state-of-the-art" numerical and laboratory modeling.

IMPROVEMENT AREA

AS I V EFAILURESURFACEDURINGEARTHQUAKE

ACTIVE FAILURE. SURFACEDURING EARTHQUAKE

LIQUEFACTIONLAYER

NONLIQUEFACTION LAYER

(a) FOR ANCHOR PLATE

IMPROVEMENT AREA

211111111.1111illai'lliV

30°PASSIVE FAILURE

SURFACEDURING

EARTHQUAKE30.

2.,13

LIQUEFACTIONLAYER

NONLIQUEFACTION LAYER

(b) FOR ANCHOR PILE

IMPROVEMENT AREA

1.1: POINT OF CONTRAFLEXURE

LIQUEFACTIONLAYER

AVNONLIQUEFACTION LAYER

(c) FOR BATTER ANCHOR PILES

Figure 3.3: Improvement Area for Sheet Pile Wall Anchors(PHRI, 1997)

Page 49: Nason J. McCullough for the degree of Master of Science in

35

4 SEISMIC ANALYSIS AND DESIGN OF SHEET PILE BULKHEADS

4.1 INTRODUCTION

The static design of flexible retaining walls involves using limit-equilibrium methods

that have been developed for rigid walls. An addition to the rigid-body design procedures

has been made to account for the reduced moment observed in flexible walls (Rowe,

1952). The static design procedures are very well established and will not be discussed

herein. For a complete description of static design, the reader is referred to Terzaghi

(1954), United States Steel (1975), Arbed (1991), Ebeling and Morrison (1993), or

USACE (1994).

The "standard of practice" design method for incorporating earthquake forces on

flexible retaining walls utilizes pseudo-static earth pressures. Pseudo-static pressures are

a function of the estimated maximum horizontal acceleration, and an addition to the static

Coulomb earth pressures. The pseudo-static design of flexible walls also uses Rowe's

(1952) flexible wall moment reduction, and for potentially liquefiable soils, empirical

estimations of pore pressure generation are also included.

There are several limitations of pseudo-static design, which includes the assumptions

inherent in Coulomb's earth pressure theory (e.g. planar failure surface, soil is isotropic

and homogeneous, the failure wedge behaves as an elastic-perfectly plastic rigid body,

etc.); 1) the soil profiles must be simple (only one soil layer) and uniform into the

backland; 2) the transient seismic motion is modeled with a single parameter; 3) very

simplified inclusion of pore pressure generation; and, 4) pseudo-static design is not

deformation-based. The "standard of practice" pseudo-static design procedures are

summarized in Section 4.2. There have been several recent modifications, or different

approaches to the pseudo-static design of flexible walls that can be categorized into either

design methods for potentially liquefiable or non-liquefiable soils. Several of these

methods are summarized in Section 4.3

Page 50: Nason J. McCullough for the degree of Master of Science in

36

4.2 PSEUDO-STATIC DESIGN

The standard of practice for earthquake design of flexible retaining walls utilizes

pseudo-static earth pressures calculated from a modified representation of the Coulomb

earth pressure theory developed by Okabe (1926) and Mononobe and Matsuo (1929).

This method is commonly referred to as the Mononobe-Okabe method. The pseudo-static

earth pressures are represented as static earth pressures increased by the earthquake

induced inertia of the soil. The inertial forces are estimated using the horizontal (kh) and

vertical (10 seismic coefficients. The seismic coefficients are empirical functions of the

estimated maximum acceleration at the bulkhead. One method commonly used to

estimate the horizontal seismic coefficient is to take 65% of the maximum acceleration

(Ebeling and Morrison, 1993). Another common method for estimating kh was developed

by Noda et al. (1975), and is given by Equation 4.1, utilizing the data of Figure 4.1.

akh=g

1(a)3 g

3

0.2g

a> 0.2g

4.1

where a is the maximum earthquake acceleration and g is the acceleration due to gravity.

The figure developed by Noda and others (1975) has been updated by Nozu et al. (1997),

and is shown in Figure 4.2, and includes a more recent assessment of the seismic

performance of retaining structures.

There has been very little research into the effect of 1c, on the pseudo-static earth

pressures, even though the vertical acceleration values can equal or exceed the horizontal

values, especially in epicentral regions (Ebeling and Morrison, 1993). The difference

between the upward and downward vertical accelerations also needs to also be addressed

during design. Ebeling and Morrison recommend that a vertical seismic coefficient be

used for anchored sheet pile walls when the horizontal seismic coefficient exceeds 0.05g.

Page 51: Nason J. McCullough for the degree of Master of Science in

0.3

0.2

0.1

0

th-a/s =-1-(g )I3

t23 28(1935) 24

t (195?) )' 26/7(19

I 15 8(19 3::11 2' 9:134 3 2

17.(1964)

?Al

...".

it t 1 Pt.Mont t .* 4 i .1-

33 1 /1'24t5(1973) A !..3.'...29.. 21

731,1 "...14

"..._ 28(1930)e ,.

2719

716

27 ''21'1)1,13.5a/8

123

18

s Gravitational Acceleration

Ground AccelOD d Acceleration

100 200 300 400

Ground Accelerationgals

Figure 4.1: Relation between Seismic Coefficient and GroundAcceleration (Noda et al., 1975)

035

0.30

0.25

0.20

a)00

U)0.1

0.0

0.

500

I

iI

V V! ! A5 I i

9 1

ieii

AvA $ elIca

1h=. It. °e 3

161,1

---t-i

1

( a )3I

gc I A iAI tibia j 'qv , bti

1 v° v1 1 I iiI

AZ: Aiflk! A I

1 A I1 Al

1

Upper and lower limits of seismic coefficient estimated fromstability analysis of damaged quaywalls without liquefaction. i-

i

Lower limits of seismic coefficient of 7 quaywalls estimated from 1

stability analysis of damaged quaywalls.1

i I 1 I1

100 200 300 400 500

Peak Ground Acceleration for SMAC-type Accelerograph (gal)

Figure 4.2: Seismic Coefficient Estimated from DamagedQuaywall (after Nozu et al., 1997)

600

37

Page 52: Nason J. McCullough for the degree of Master of Science in

38

They do not, however, recommend any relationships or values for kv. The vertical seismic

coefficient is often neglected in design.

The Mononobe-Okabe method is a modification of the static Coulomb earth pressure

theory and the resultant forces are the combined static and dynamic forces. The resulting

dynamic earth pressure forces and angles are defined in Figure 4.3 and given in Equations

4.2 through 4.8 for the passive and active cases of dry, uniform soils.

rPee = Ka,

1[y t (1 1 cv)1H2

1Ppe = Kpe v1(1-k)]H2

where: P = dynamic plus static earth pressure forceK = coefficient of dynamic earth pressure

= total unit weight of the soillc, = vertical seismic coefficientH = height of the wall

4.2

4.3

Figure 4.3: Definition Sketch for Pseudo-Static Design Parameters

Page 53: Nason J. McCullough for the degree of Master of Science in

The coefficients of earth pressures are given by the following;

where:

39

cos2 (0 V)= 4.4ae 2

Kpe

cos coo +0[1

+ Ilsin(fb + sin(q$ 99)

cos(8 + (o)cosp

cos2 (0 V)

cos co cos(8 0[1 ilsn° 4- sin(Ch fi CI))

cos(8 + go)cosfl

K =

=

/6

2

active and passive dynamic earth pressure coefficientangle of internal friction for the soilseismic inertia angleangle of friction between the soil and wallslope of the backfill

The seismic inertia angle is given by;

= arctan[ khk, )1

The angle of the failure plane to the horizontal is given by the following;

tan(0 co /6)+a = 0 co + arctan[1 + [tan(8 + yokan(( P)+ cot(0 P)A1

tan(fb /6)+ C2ape = 0 + p arctank[tan(8 + gokan(0 + /6)+ cot(93 P))]]

4.5

4.6

4.7

4.8

Page 54: Nason J. McCullough for the degree of Master of Science in

40

where;

= Vtan(0 /3)[tan(cb co M+ cot(q + tan(8 + Ocot(g) co)]

c2 = Vtan(0 p +13){tan(0 co + ,(3)+ cot(q$ 41+ tan(8 + yo)cot(0 (0)]

These equations are applicable for dry backfill only. If the backfill is saturated, it is

necessary to modify the above equations (as described in the following sections),

depending on whether the backfill soils are potentially non-liquefiable or liquefiable.

4.2.1 Non-Liquefiable Soil

The following equations are used for non-liquefiable soils, and they are dependent on

the behavior of the pore water relative to the soil matrix. If the hydraulic conductivity of

the soil is less than 1x10-3 cm/sec, the pore water is assumed to be restrained and if the

hydraulic conductivity is greater than 1x10"3

cm/sec, the pore water is assumed to be free.

The term restrained is used to define the condition where the pore water moves with the

soil skeleton and there is no fluid flow, and the term free is the condition where the pore

water moves relative to the surrounding soil during shaking. These conditions do not

affect the generation of excess pore pressures, and are only used to approximately

describe the pore water behavior during an earthquake.

4.2.1.1 Restrained Water Case

For submerged backfill with restrained water and without excess pore pressure

generation the following equation is commonly used to determine an effective horizontal

seismic coefficient;

71 7,nh'=--mhYb

4.9

Page 55: Nason J. McCullough for the degree of Master of Science in

41

where: y, = total unit weight of the soilYb = buoyant unit weight of the soil

The effective horizontal seismic coefficient (kh') should then be substituted for kh in

Equation 4.6. It is also necessary to use yb in Equations 4.2 and 4.3, and to include the

hydrostatic water forces on the wall.

4.2.1.2 Free Water Case

For the case of free pore water, Equations 4.10 and 4.11 should be used to estimate

the effective seismic coefficient and hydrodynamic force.

k h ' = 21 khYb

where: yd = dry unit weight of the soilYb = buoyant unit weight of the soil

4.10

If Equation 4.10 is used to calculate Equation 4.6, it is necessary to use yb in

calculating Pa, and Ppe. Because the water is now treated as acting independently from the

soil, it is necessary to include the inertia force of the water within the backfill. The inertia

forces from the free water are estimated using the following relationship developed by

Westergaard (1931). The resultant P,, is assumed to act at 0.4(H+D) above the base of the

wall.

P, = ukhr (H + D)2

where: Pi =kh =y, =H =D =

inertial water pressure forcehorizontal seismic coefficientunit weight of the waterwall heightdepth of embedment

4.11

Page 56: Nason J. McCullough for the degree of Master of Science in

42

The location of the resultant earth pressure forces from the Mononobe-Okabe method

cannot be determined from the above relationships. To determine the point of action for

the resultant forces, the following procedures developed by Seed and Whitman (1970) are

commonly used. Equation 4.12 (Seed and Whitman, 1970) provides the dynamic active

earth pressure force (Pae) as the sum of the static active earth pressure force (Pa) and the

dynamic active earth pressure force increment (AP,e);

Pae = Pa + APae 4.12

The static analysis for a uniform soil deposit, gives Pa acting at one third the total

wall height above the bottom of the wall. Seed and Whitman (1970) present APae as

acting at 0.6 times the total wall height above the bottom of the wall. Utilizing the known

points of action and Equation 4.12, Equation 4.13 was derived by Seed and Whitman to

determine the point of action for Pae above the bottom of the wall. A definition sketch is

given in Figure 4.4.

Y=pa

(H +APae[0.6(H + D)]

3

Pae4.13

APae

g

0.6(H+D)

1/3(H+D)

Figure 4.4: Definition Sketch for the Location of the ResultantDynamic Active Earth Pressure

Page 57: Nason J. McCullough for the degree of Master of Science in

43

9.2.2 Liquefiable Soil

A potentially liquefiable soil is defined as a soil having the possibility of excess pore

pressure generation during cyclic shaking. The common method for representing excess

pore pressure generation in design was given by Equation 2.8 (ru = uexcessl crvo), where ru

represents the excess pore pressure ratio. Full liquefaction is defined as the state when the

excess pore pressures equal the effective overburden pressure, or rt, is equal to unity.

Potentially liquefiable soils will behave in one of three ways during dynamic excitation;

1) no excess pore pressure generation will occur (ru= 0),

2) partial excess pore pressures will be generated, but full liquefaction will not

occur (0 < ru < 1), or

3) full liquefaction will occur (ru = 1).

There is also the question of whether the pore water will flow freely or be restrained

during the cyclic loading, similar to the non-liquefaction case. The different procedures

are described below.

4.2.2.1 Restrained Water Case

There are two ways to analyze the restrained water case with excess pore pressure

generation using simple analytical procedures. The first is to have the excess pore

pressures reduce the unit weight of the soil and increase the unit weight of the water, until

full liquefaction, where the unit weight of the water equals the summation of the unit

weights of the soil and water. The effective unit weight of the soil and water are given by;

7 6' = 7 b(1 ru) 4.14

Twt= 7w + I b ru 4.15

Page 58: Nason J. McCullough for the degree of Master of Science in

44

It is then necessary to compute an effective horizontal seismic coefficient (Equation

4.16) and corresponding seismic inertial angle (Equation 4.6).

'ch.= 11--, khYb

4.16

The effective unit weights should be substituted in all preceding calculations, in

place of the soil and water unit weights. As rz, approaches unity, m' approaches zero, and

7w' becomes the total weight of the soil and water combined. Therefore, during full

liquefaction the soil-water matrix is treated as a heavy fluid. When full liquefaction

occurs, the unit weight of the heavy fluid (70 should be used with Equation 4.11 to

determine the dynamic fluid pressure.

The second method of including pore pressure generation involves decreasing the

effective angle of internal friction of the soil, as given in Equation 4.17. Ebeling and

Morrison (1993) describe this procedure in detail, and based on their calculations, it is not

recommended that this procedure be used because it over predicts Pae, notably when kh

approaches zero.

tan 01,q = (1 ru) tan 0'

4.2.2.2 Free Water Case

4.17

If the water acts independently from the soil, the pressures are due to the thrust from

the soil, and the thrust from the water. The thrust from the soil is estimated by using an

effective horizontal seismic coefficient (Equation 4.18), calculated from the effective soil

unit weight (Equation 4.14) and substituted into Equation 4.6.

kh'=L-1 ,kh

b4.18

Page 59: Nason J. McCullough for the degree of Master of Science in

45

The thrust from the water is estimated using Equation 4.11 with 7w, and the static

water pressure using 7w' from Equation 4.15.

4.3 RECENT ADVANCEMENTS IN FLEXIBLE WALL DESIGN

There have been numerous recent advancements in the design of flexible walls that

either enhance the pseudo-static design methods that propose alternative methods of

analysis. These methods are used for two different potential cases, liquefiable and non-

liquefiable soils. The following sections summarize these recent advancements for both

cases, and the sub-sections are differentiated by the lead researchers.

4.3.1 Non-Liquefiable Soils

Neelakantan et al. Neelakantan et al. (1992) developed a balanced seismic design

method for anchored bulkheads without liquefaction, and validated the design method

with large shaking table tests. The method determines a factor of safety against wall

failure by dividing the driving moment (active seismic earth pressures) by the resisting

moment (passive seismic earth pressures). A balanced seismic design method was also

developed for use in anchor design. The researchers note that the balanced seismic design

method for anchored sheet pile walls enhances the seismic resistance as obtained from the

Mononobe-Okabe method, based on the results of the large shaking table tests.

Steedman and Zeng. Steedman and Zeng (1990) performed centrifuge model studies

and developed an analytical design method for anchored walls. The effects of varying the

shear modulus with depth, ground motion amplification with depth, the hydrodynamic

water pressure, and the natural frequency of the wall were analyzed. The study indicates

that their results are in good agreement with the Mononobe-Okabe design method. It was

also noted that in certain situations, the actual hydrodynamic water pressure on the wall is

larger than that predicted by Westergaard's (1931) formulation. It was also noted that the

response of a flexible wall during earthquake ground motions is highly dependent on the

natural frequency of the bulkhead.

Page 60: Nason J. McCullough for the degree of Master of Science in

46

Dennehy, Gazetas, and Dakoulas. In a modification of the "standard of practice"

pseudo-static design method to include deformation based analysis, Dennehy (1985) and

Gazetas et al. (1990), developed an empirical relationship between sheet pile damage and

sheet pile wall geometry for locations that were reported as not experiencing liquefaction

flow failures during the specific earthquake. Two non-dimensional factors, the effective

anchor index (EAI) and the embedment participation index (EPI) were related to the

degree of damage recorded for the bulkheads. Referring to Figure 4.5, EAI is used to

quantify the amount of available

EAI =dH

The EPI provides the contribution

P eEPI

anchor capacity

of the

be approximated

with;

4.19

wall embedment as;

4.20

as;

4.21

=Pae

For uniform soils,

EPI KPe

f + H

EPI can

f ) (1=Kae f+H f+H

The relationships between the EAI and EPI factors were plotted for 75 Japanese case

histories (Figure 4.6), with the descriptions of the degrees of damage given in Table 4.1.

Although this chart has greatly enhanced the deformation based design of sheet pile

bulkheads, several limitations are noted;

1) the design chart applies only to non-liquefiable soils,

2) the criteria for the occurrence of liquefaction is based on surface evidence

(whereas the lateral displacements of retaining structures masks the evidence of

Page 61: Nason J. McCullough for the degree of Master of Science in

47

liquefaction at depth), it is therefore questionable whether liquefaction did not

occur at all of the 75 case histories,

3) there is no direct contribution from the wall stiffness or earthquake characteristics

(i.e. intensity, frequency, duration) to the deformations, and

4) the EAI and EPI factors are indirectly a function of the seismic coefficients used

in design, which has been noted by Kitajima and Uwabe (1979) to be

unsatisfactory in relation to seismically induced deformations.

effective "point"of rotation

active failure surface

Figure 4.5: Definition of Effective Anchor Index, EAI (afterGazetas et al., 1990)

Table 4.1: Qualitative and Quantitative Description of theReported Degrees of Damage (after Kitajima andUwabe, 1979)

Degree ofDamage Description of Damage

Permanent Displacementat Top of Sheet pile (cm)

0 no damage < 21 negligible damage to the wall itself; noticeable damage

to related structures (i.e. concrete apron)10

2 noticeable damage to the wall itself 303 general shape of anchored sheet pile preserved, but

significantly damaged60

4 complete destruction, no recognizable shape of wallremaining

120

Page 62: Nason J. McCullough for the degree of Master of Science in

48

2

1.5

1

0.5

0

-0.5

0 0.5 1

Embedment participation index (EPI)

Figure 4.6: The Developed Seismic Design Chart (Gazetas et al.,1990)

The design chart provides a useful method of estimating deformations for flexible

bulkheads embedded in non-liquefiable soils. For cases where liquefaction is expected,

this method is unsatisfactory for deformation-based design.

4.3.2 Liquefiable Soils

Towhata and Islam. Towhata and Islam (1987) developed a method of estimating the

critical seismic coefficient necessary to initiate displacements. Their approach is similar

to the method developed by Newmark (1965) to estimate lateral deformations of rigid

Page 63: Nason J. McCullough for the degree of Master of Science in

49

retaining structures. Towhata and Islam used the simplified expressions of the dynamic

earth pressure coefficients proposed by Seed and Whitman (1970), given as;

3Kae = Ka + AK = Ka +-4kh

17Kpe = Kp + AKpe =K p 8 kh

4.22

4.23

Neglecting wall friction and and equilibrating the forces in the vertical and

horizontal, Towhata and Islam developed the following equation for estimating the

critical horizontal seismic coefficient (kcr).

kcr =

where:

a tanaae btan(cb aae)(1+btanaae)1+ ctanaae

m7; + PP

+1 ,(h, + D) 2 +

a=W,

yw(hw tan0+ AU sinaaeb = 2 tang,

W.

1c =

Wra

23mn7; 17 Pp)/ t 7 ywhw2

8(Kp Ka) 8Kpr b 12

rWm = -i[rt(h.-FD)2 ±rd(H h,)(H + 2D + hi,)1

4.24

Page 64: Nason J. McCullough for the degree of Master of Science in

50

where: m = 0 for no anchor capacity, 1 for full anchor capacity (dependant on theamount of excess pore pressure generation around the anchor)

T, = the static anchor forceAU= excess pore water pressure due to cyclic shearingn =1 when the anchor block is above the water table, and y/yb when the

anchor is completely submerged

Using Equation 4.24 it is possible to determine the limiting kh for lateral

deformations to occur. Towhata and Islam also performed numerical studies on the

effects of the anchor capacity and depth of embedment for a specific case. No generalized

guidelines are provided.

Byrne et al. Byrne et al. (1994) developed a rational method of estimating the seismic

displacements of earth dams using post-liquefaction stress-strain relations and energy

concepts. Even though this method was not developed specifically for the seismic design

of sheet pile bulkheads, it may be applied (with considerable judgement) to a

performance-based design of sheet pile bulkheads.

4.4 DISCUSSIONS OF SHEET PILE BULKHEAD DESIGN METHODS

The seismic performance of anchored sheet pile walls in non-liquefiable soils

designed with Mononobe-Okabe has been mixed. The cases of non-liquefaction failures

have been due primarily to inadequate depth of embedment, anchor design and/or the

partial generation of excess pore pressures. These deficiencies have been addressed to a

degree in the current design methods (Ebeling and Morrison, 1993) and by the work of

Dennehy (1985) and Gazetas et al. (1990). Neelakantan and his co-workers developed a

straightforward balanced seismic design method that also accounts for the past seismic

failures due to inadequate anchor design. Steedman and Zeng (1990) have highlighted

other limitations, such as the need to include the dynamic amplification of ground

motions due to the response of the backfill and phase effects on the behavior of flexible

bulkheads. These methods, though useful for non-liquefaction design (competent or

improved soils), are severely limited in that the majority of anchored bulkhead failures

Page 65: Nason J. McCullough for the degree of Master of Science in

51

have been caused by the liquefaction of backfill soils (which is only addressed in a

simplified manner by the Mononobe-Okabe method and its derivatives). The majority of

these methods also lack the ability to evaluate deformations (except for the method by

Dennehy and his co-workers). The only design methods that account for liquefaction are

Mononobe-Okabe (in a simplified method) and Towhata and Islam, neither method

allowing for deformation estimations. A serious limit to all of these methods is that the

transient earthquake motions are approximated by single seismic coefficients that do not

account for the duration or frequency content of actual ground motion time histories.

Page 66: Nason J. McCullough for the degree of Master of Science in

52

5 NUMERICAL MODELING

The database of case histories on the seismic performance of sheet pile bulkheads is

very limited. In order to supplement the case history data, a numerical modeling study

was conducted to analyze the seismic performance of sheet pile bulkheads. The numerical

model is advantageous for this study because numerous scenarios can be analyzed, and

the various design parameters can easily be adjusted to determine their influence on the

seismic performance of the bulkhead. The major concern in applying a numerical model

to soil-structure interaction problems is the numerical uncertainty, which is limited in this

study by performing a series of validation studies using the available case history data on

the seismic performance of sheet pile bulkheads.

The numerical modeling was accomplished utilizing a commercial finite difference

computer program entitled Fast Lagrangian Analysis of Continua (FLAC) version 3.30

(Itasca Consulting Group, 1995). FLAC is a non-linear, two-dimensional finite difference

program capable of modeling both static and dynamic situations. Elements or zones

represent the materials (structural and soil), with all of the elements and zones

constituting the grid (mesh). The numerical formulation of FLAC utilizes a time-

marching scheme, where during each timestep the following procedures take place within

FLAC (Figure 5.1); 1) nodal velocities and displacements are calculated from stresses and

forces using the equations of motion, 2) a constitutive model is then used to calculate

strain rates from the velocities and stresses from the strain rates. These two procedures

are then repeated until the computed unbalanced forces within the mesh are within a user

specified limit. FLAC uses an explicit method, where the calculation timestep is very

short compared with the time necessary for information (acceleration, velocity,

displacement) to physically pass from one element to another. FLAC also utilizes a

Lagrangian formulation, in which the incremental displacements are added to the

coordinates at each timestep so that the grid moves and deforms with the material that it

represents.

Page 67: Nason J. McCullough for the degree of Master of Science in

53

There are several advantages and disadvantages in using FLAC, compared to implicit

finite element programs as outlined in Table 5.1. Because of the explicit Lagrangian

methodology and the explicit use of the equations of motion, FLAC is advantageous in

modeling nonlinear, large strain, physically unstable situations (Itasca, 1995).

new velocitiesand

displacements

Equilibrium Equation(Equation of Motion)

Stress/Strain Relationship(Constitutive Model)

Figure 5.1: Basic Explicit Calculation Cycle

newstresses or

forces

This chapter presents the aspects of the numerical modeling program used in this

study. Section 5.1 describes the constitutive soil model, Section 5.2 describes the pore

pressure generation scheme, Section 5.3 describes the general modeling parameters (for

soils, structures, water, earthquake, and the boundary conditions), and Section 5.4

describes the extensive program used to validate FLAC for use in modeling seismically-

induced liquefaction and the performance sheet pile bulkheads.

5.1 CONSTITUTIVE SOIL MODEL

An effective stress Mohr-Coulomb constitutive model was used for this study. The

constitutive model is able to model plastic deformations utilizing a plastic flow rule. The

Page 68: Nason J. McCullough for the degree of Master of Science in

54

Table 5.1: Comparison of FLAC and Finite Element NumericalPrograms

FLAC (explicit) Finite Element (implicit)

Timestep must be smaller than a critical value forstability.

Timestep can be arbitrarily large, withunconditionally stable schemes.

Small amount of computation effort per timestep. Large amount of computational effort pertimestep.

No significant numerical damping introduced fordynamic solution.

Numerical damping dependent on timesteppresent with unconditionally stable schemes.

No iterations necessary to follow nonlinearconstitutive law.

Iterative procedure necessary to follow nonlinearconstitutive law.

Provided that the timestep criterion is alwayssatisfied, nonlinear laws are always followed in avalid physical way.

Always necessary to demonstrate that the abovementioned procedure is a) stable, and b) followsthe physically correct path (for path-sensitiveproblems).

Matrices are never formed. Memory requirementsare always at a minimum. No band-widthlimitations.

Stiffness matrices must be stored. Ways must befound to overcome associated problems such asband-width. Memory requirements tend to belarge.

Since matrices are never formed, largedisplacements and strains are accommodatedwithout additional computing effort.

Additional computing effort needed to followlarge displacements and strains.

elastic behavior of the soil is defined by the bulk and shear modulus, and the strength is

defined by the angle of friction and cohesion. This significantly simplifies the dynamic

soil behavior and is not capable of accounting for the strain dependent dynamic properties

such as damping and shear modulus. Despite the use of this simplified constitutive soil

model, it has been demonstrated to yield satisfactory displacement results for a variety of

applications involving seismically-induced deformations of earth structures and retaining

walls (e.g. Roth et al., 1986; Roth & Inel, 1993; Dickenson & Yang, 1998). The elastic

and strength properties of the soil were estimated from established correlations with

normalized SPT values ((Nd60), unless otherwise noted.

Page 69: Nason J. McCullough for the degree of Master of Science in

55

5.2 PORE PRESSURE GENERATION

The pore pressure generation scheme utilized in these analyses uses several empirical

procedures developed by Seed and others over the last 25 years (Martin et al., 1975, Seed

et al., 1976, Seed 1979). During each timestep of the dynamic analysis, the effective

stresses decrease and pore pressures gradually increase in liquefiable soils, until a state of

full liquefaction is reached. Liquefaction resistance curves are developed using the cyclic

stress ratio (CSRfi ) and number of cycles to liquefaction (Nijq) at 3 and 30 cycles (Figure

5.2). The values of 3 and 30 are used because they cover the range of number of shear

cycles for earthquake magnitudes in the range of engineering interest (M = 5 to 8). The

linear relationship provides a fairly good representation of the majority of CSRfied curves

(Figure 2.3).

CSR

CSR @3 cycles

CSR @30 cycles

3 30

Number of Cycles Required for Liquefaction (Nhz)

Figure 5.2: Modeled Liquefaction Resistance Curve

There are two methods of inputting the CSRfidd data; 1) directly from the results of

cyclic testing (e.g. simple shear, triaxial), or 2) utilize in-situ penetration resistances (e.g.

SPT, CPT) and empirical relationships. Method 1 is straightforward, as long as the

CSRfieid is corrected for the testing method, in-situ stresses and various earthquake

magnitudes, as described in Section 2.2.2. For method 2, the CSRfidd is corrected for

initial static shear stresses and initial overburden stresses (Section 2.2.2), and the values

Page 70: Nason J. McCullough for the degree of Master of Science in

56

of CSR at 3 and 30 cycles are determined by multiplying CSRfidd value by 1.5 and 0.89,

respectively. The multiplication factors were determined from Arango (1996) for various

numbers of shear cycles.

Cyclic stress ratios caused by the earthquake motion (CSReq) are monitored during

each timestep, and if CSReq exceeds CSRfield, a numerical curve fitting is conducted to

determine Nhq at CSRfieid. The earthquake-induced damage (a function of the excess pore

pressure generation) is monitored in the FLAC model using a damage parameter (D). The

damage parameter is updated at each shear cycle, and is defined as the summation of the

inverse number of cycles to liquefaction (Equation 5.1). After each cycle, D is related to

the pore pressure by an empirical function. The simplest function was used in the

current study, based on the satisfactory results of Roth and Inel (1993) for liquefaction

studies.

D =E 1

Niiq5.1

A simplification was made in the models in which pore pressures were allowed to

generate using the above formulation, but were not allowed to dissipate. This

simplification was used because the results from validation case studies in which the

dynamic motion and groundwater dissipation where coupled proved to be unsatisfactory.

The reasons for the unsatisfactory modeling of dynamic motion and groundwater

dissipation are unclear. The simplification that was made can be somewhat justified by

noting that expected flow in sand during the 10 to 40 seconds of earthquake motion

would be very small (-0.1 to 0.4 cm3/cm2). It should also be noted that the largest effect

of pore pressure dissipation is ground surface settlement, which was calculated following

the analyses using the proposed post-liquefaction volumetric strain relationship by

Ishihara and Yoshimine (1992). Therefore, for this study, the neglect of pore pressure

dissipation is an acceptable simplification.

Page 71: Nason J. McCullough for the degree of Master of Science in

57

5.3 GENERAL MODELING PARAMETERS

The numerical modeling involved two solutions; static and dynamic. After

generating the model geometry, a static solution was sought to equilibrate the initial

stresses within the soil. After the static solution, the dynamic modeling was performed.

The modeling for the validation case histories and the parametric studies (Chapter 6)

followed the same procedures, as outlined in the following sections.

5.3.1 Modeling of Soil Elements

The known site properties for most of the analysis included the uncorrected standard

penetration test (SPT) results, and the elevation of the soil layers. In order to conduct a

FLAC analysis, the soil density (p), angle of internal friction (0), cohesion (c), bulk

modulus (K), shear modulus (G), Poisson's ratio ( v), corrected penetration resistance

(Wd60), relative density (DR) and undrained residual shear strength of the liquefied soil

(Sr) are also needed. The soil density and angle of internal friction were estimated from

the SPT results using well-established correlations, when the values were not provided in

the references. The low-strain shear modulus (Gmar) was calculated from the shear wave

velocity (Vs) and total density of the soil (Equation 5.2). The shear wave velocity was

estimated from the SPT, when not provided in the references. Several different

established methods were used to estimate the shear wave velocity of the soils, depending

on the specific site profile and soil conditions.

Gmax = Vs2 Ptotal 5.2

Once the shear modulus was determined, Equation 5.3 was used to estimate the bulk

modulus of the soil, using Poisson's ratio.

K =2G(1+ 0)

3(1 2v)5.3

Page 72: Nason J. McCullough for the degree of Master of Science in

58

The corrected penetration resistance was calculated using Equation 5.4. The initial

static analysis provided the effective overburden pressures (o-') in the model, which was

then used to determine CN using Equation 5.5 (Liao & Whitman, 1986). Because all of

the case histories were from Japanese cases, the efficiency (Er) of the SPT method was

assumed to be 72% of the maximum, which is the average of the free-fall and throw

release type hammers, as given by Seed et al. (1985).

(N1)60 N 'LNEr60

CN =95760 Pa

avo

5.4

5.5

If the shear wave velocity was unknown, it was estimated from SPT data using the

relationship given in Equation 5.6, formulated by Imai et al. (1975) from Japanese case

history data.

Vs = 89.8N0341 5.6

The relative density of the soil was estimated using Equation 5.7 (Yoshida et al.,

1988), with the following; SPT values uncorrected for overburden pressure (N60), the

effective overburden pressure (cr), and the best fit values of Co = 25, Cl = 0.12 and

C2 = 0.46. Equation 5.7 was used to estimate relative density because it was formulated

from Japanese case history data, and all of the validation case studies were from Japanese

locations.

5.7

Page 73: Nason J. McCullough for the degree of Master of Science in

59

The undrained residual shear strength was estimated using Equation 5.8 (from Figure

2.11), developed by Stark and Mesri (1992) for SPT results corrected for fines content.

S,. = 0.0055 avo '*(N1)60-cs 5.8

The constitutive model used during the numerical analysis followed the Mohr-

Coulomb failure criteria (r=cr'tan0), and as the pore pressures increase during cyclic

shearing, the effective stresses decrease, until the limiting shear stress value (undrained

residual shear strength) is reached. The strength of the soil at full liquefaction was then as

the undrained residual shear strength.

5.3.2 Modeling of Structural Elements

The sheet pile walls and anchors were modeled using FLAC pile elements. These

elements allow relative movement between the soil and pile elements. The relationship

between the pile and soil elements is simulated with springs that have a stiffness and

yield value for both the normal and shear directions. The stiffness was calculated as the

minimum value of Equation 5.9 for all soil elements adjacent to the pile elements. The

same calculated stiffness value was used for both the shear and normal directions. The

yield value for the normal direction was set very large to prevent the soil moving through

the wall, whereas the yield value for the shear direction was set equal to zero, assuming

that there was no static friction between the soil and pile. A frictional shear resistance

angle (b) was input for the dynamic (not earthquake) friction between the wall and soil. If

the wall was continuous (sheet pile wall, or a continuous anchor wall), the modulus of

elasticity was increased to account for the effect of plane strain using Equation 5.10

(Itasca, 1995) to determine the new modulus. If the tie rods or anchor were not

continuous, the modulus of the tie rod and anchor were divided by the spacing of the tie

rod or anchor to provide a modulus value per unit depth. The tie rods were modeled as

cable elements, which are able to transfer tension loads, but unable to transmit moments.

Page 74: Nason J. McCullough for the degree of Master of Science in

K +yG3Stiffness =

62min

E plane strainE

1 02

5.3.3 Modeling of the Earthquake Motion

60

5.9

5.10

The input earthquake motions were recorded accelerograms that had been corrected

for the recording instrument. For simplification, the vertical motion was not used, and the

horizontal motions were vectorally summed to the motion perpendicular to the face of the

wall. The motions were input at the base of each model as acceleration time histories. The

damping of the earthquake motions for numerical stability utilized the Rayleigh damping

method within FLAC. The acceptable Rayleigh damping used in the models was

determined from the validation studies to be 5% at 5 Hz.

There is an inherent displacement in recorded earthquake motions due to various

factors (e.g. movement of the recording instrument, permanent earthquake

displacements), and is termed baseline shift. The baseline shift has been removed from all

of the presented FLAC displacements.

It should also be noted that the applied acceleration time histories do not always

correspond to the same time scale as the recorded motions. This is because some of the

small amplitude portions of the recorded motions were removed to decrease the solution

time for the numerical runs.

5.3.4 Modeling of the Water

The water in front of the wall was modeled indirectly by including the resulting water

pressures along the boundaries. This allows a simple modeling of the water, but does not

account for the dynamic interaction of the wall and water. The water within the soil was

modeled directly, and was allowed to flow during the static solutions. During the dynamic

Page 75: Nason J. McCullough for the degree of Master of Science in

61

solutions excess pore pressures were allowed to generate, but the dissipation of these pore

pressures was not modeled (as discussed in Section 5.2).

5.3.5 Boundary Conditions

The boundary conditions for the static solutions consisted of the bottom boundary

being fixed in both the horizontal and vertical directions, while the sides of the model

were treated as rollers (by fixing only the horizontal direction). During the dynamic

analysis the bottom boundary was freed in the horizontal direction to allow application of

the horizontal acceleration, and the sidewalls were treated as an infinite medium (free

field), having the same properties as the adjacent model perimeter zones.

The lateral dimensions of the model were determined as optimum from the validation

case studies as seven times the total wall height (H+D) in the backland and three times

the total wall height in front of the wall.

5.4 VALIDATION OF NUMERICAL MODEL

The use of FLAC for the seismic modeling of generic sheet pile walls in ports

required that the numerical model be validated and calibrated. The validation of FLAC

for this study, involved modeling well-documented case histories and comparing the

calculated and measured displacements. Five case histories were chosen, including: two

wharves at Akita Port during the MJMA 7.7 1983 Nihonkai Chubu Earthquake, a wharf at

Ishinomaki Port during the MJMA 7.4 1973 Miyagi Ken Oki Earthquake, and two wharves

at Kushiro Port during the MJMA 7.8 1993 Kushiro Oki Earthquake. These were the most

detailed case histories available for typical sheet pile bulkheads. The modeled geometries

were the same as given in the references, except for the noted differences and

simplifications.

Several notes on the calculated displacements include; a) the horizontal

displacements were measured at the top of the sheet pile wall. The vertical displacements

were measured at the top of the soil column, adjacent to the wall, therefore the reported

Page 76: Nason J. McCullough for the degree of Master of Science in

62

vertical displacements take into account the relative vertical movement between the wall

and soil. b) the vertical time history displacement plots are a result of the volume change

due to horizontal movements only, and do not account for the volumetric strain due to

pore pressure dissipation. The volumetric strain due to pore pressure dissipation was

calculated following the analysis using the proposed post-liquefaction volumetric strain

relationship by Ishihara and Yoshimine (1992). The post-liquefaction vertical

displacements are noted in the text.

5.4.1 Akita Port

The 1983 Mjmik 7.7 Nihonkai Chubu Earthquake struck Akita Port, Japan, causing

severe liquefaction damage in some locations, while other locations were left undamaged.

Two case histories were chosen from this earthquake. One site suffered severe damage,

while the other suffered negligible damage. There is considerable data available on the

earthquake, site conditions, and structure properties for both case histories (Iai &

Kameoka, 1993). The layout of the port (Figure 5.3) includes the location of the strong

motion observation station where the modeled earthquake acceleration time histories

were recorded. Ohama Wharf 1 and 2 bulkheads are almost identical. The differences

include the soil properties of the backfill, depth of wall penetration, and structural details

of the anchors. Ohama Wharf 1 and 2 bulkhead geometries are show in Figure 5.4 and

Figure 5.5, respectively.

5.4.1.1 Soil Conditions

The soil conditions, including standard penetration and sieve analysis results, for

Wharf 1 and 2 are shown in Figure 5.7 and Figure 5.8, respectively. The numerical

modeling utilized the following soil layers and conditions. For Wharf 1, the soil profile

was simplified into two layers for the numerical analysis, the top layer, a hydraulically

backfilled sand, extends to a depth of approximately 22.5 m below the ground surface.

The second layer is also a sand layer, and extends to the base of the model (24 m). For

Wharf 2, four soil layers were modeled. The hydraulically backfilled sand layer, which

Page 77: Nason J. McCullough for the degree of Master of Science in

63

has a sloping base (Figure 5.5), is 12 m thick adjacent to the sheet pile wall. The backfill

is underlain by 6 m of naturally deposited sandy soil. The third layer, a medium stiff clay,

is 4.5 m thick, and underlain by another sand layer to the base of the model (24 m). The

soil properties used in the numerical analysis are presented in Table 5.2. The strength

properties were based on the laboratory test results by Iai and Kameoka (1993).

I30°E 135° 14.°

awn, AkitaJapan Sea 0

okyo

(a) Location of Akita (b) Closer View

Ohoma No.l Wharf

Shinio RiverOhamo No. 2 Wharf

Akita part51-

'Japan Sea , /, ;/ S,trong,-Motion

.14 o o'4 Observationca

i.,k'

;--oSto ion,

Contours of Base , ...yo.

Liquefied Area p 7 .,,e ,.-Jorn0 ,

(c) Plan of Akita Port

Figure 5.3: Plan of Akita Port, Japan after the 1983 M 7.7Earthquake (Iai & Kameoka, 1993)

Page 78: Nason J. McCullough for the degree of Master of Science in

64

20.00

12.00

0.00_

-10.00

-15.50

20.50

" SEMI HIGH TENSILESTRENGTH STEELTIE ROD

Layer 2 ( Sand)

0

0:

-10.10

Unit in Meters

Loyer 4 ( Sand)

Figure 5.4: Akita Ohama Wharf 1 Wall Profile (Iai & Kameoka,1993)

20.00

t 2.00

_zr I 0 0010 HIGH STRENGTH STEEL TIE ROD 11

_11

-11

CRA C

-10.00 I

-20.50

Layer I

Bockfill Sand)

Layer 2 ( Sand)

Layer 3 ( Cloy)

Layer 4 (Sand)

O

W

a.

-14.50

- BEFORE THE EARTHQUAKEAFTER THE EARTHQUAKE

UNIT IN METERS

Figure 5.5: Akita Ohama Wharf 2 Wall Profile (Iai & Kameoka,1993)

Page 79: Nason J. McCullough for the degree of Master of Science in

(m)3 2

(m) (m) (m)2 r-1 0 2 EI 0 I r

01

ILI

I 0 1

m)10 \-110 1-110(m) (m) (m)

Figure 5.6: Deformed Cross-Sections for Akita Ohama Wharf 2

Cr 1009060

U" X 70(9 60

Z,W 50030

03 20W 100- °0001 001

0

5

10

(m)

Iiv-(No.58-22) --- -7'DEPTH3m_ I,/ , ---6

7-.---- 1

.-_ ...,,/I i

CC 100

L 70

y a05.g

201000001 0.01

01 10Imm)

10.0 500

DEPTH7 9m-1

---- 10-16

01 1.0InIrn1

Figure 5.7: Soil Profile and Grain Size for Ohama Wharf 1 (Iai &Kameoka, 1993)

10.0 50.0

65

The SPT values were input exactly as shown in the figures. Shear wave velocity (V3)

measurements were obtained from an exploratory borehole at a neighboring wharf,

Ohama Wharf 3. The measured shear wave velocity and soil profiles at Ohama Wharf 3

are given in Figure 5.9. Because the SPT data for the soil layers at Ohama Wharves 1, 2

and 3 are similar (except for the backfill), the shear wave velocities were assumed the

same between the three sites.

Page 80: Nason J. McCullough for the degree of Master of Science in

Elev-otion(m)

SoilType

S PT N -VALUE10 20 30 40 500.01

D50 ( mm )0.1 1.0

0 ''''' +0.33m

i`.Wf ,

..10

< ,-a

-15o v

tk-25

o-o 48 5

-30 58 - 2

LC5 100Z 90LT: 80

x 7060

w

W40

y 30W 20

LI 10o

-(No 58-2 ) 7/' DEPTH2m

ORIGINAL GROUNDSOIL

V 4-10- 12

- 1/ BACKFILL SAND

0.001 0.01

ix 10090

8:7,71 60

5040

to 32100

0.001 0.01

0 11mm) IO 10.0 500

f----.--- I DEPTH-(No.48-10) 16m-_ ...-- , 18/ %

20

..------------

__-

0 1 1.0

Figure 5.8: Soil Profile and Grain Size for Ohama Wharf 2 (Iai &Kameoka, 1993)

Table 5.2: Model Soil Properties for Ohama Wharf 1 and 2

10.0 50.0

66

Wharf 1 Wharf 2Layer 1 Layer 2 Layer 1 Layer 2 Layer 3 Layer 4

Density (kg/m3) 1500 1500 1500 1500 1500 1600

Friction Angle (0) 43 44 37 41 39 44Cohesion (kPa) 0 0 0 0 48 0

Poisson's Ratio 0.3 0.3 0.3 0.3 0.3 0.3Porosity 0.4 0.4 0.4 0.4 0.4 0.4

5.4.1.2 Structural Properties

Ohama Wharf 1 and Wharf 2 were constructed with the same wall section, FSP 6L,

manufactured by Nippon Steel. The properties of the modeled structural elements are

presented in Figure 5.9. The anchor and tie rod spacing was 2 m for both Wharf 1 and

Wharf 2 (Iai & Ichii, 1997). The friction between the sheet pile wall and the soil was

modeled as 14 degrees for Wharf 1, where limited pore pressure generation was expected,

and 0 degrees for Wharf 2, where significant pore pressure generation was expected.

Page 81: Nason J. McCullough for the degree of Master of Science in

67

5.4.1.3 Earthquake Motion

The acceleration time histories recorded at the strong motion site (Figure 5.3) were

vectorally summed to yield the acceleration time history normal to the wall. The motion

normal to the wall was then deconvolved to bedrock using the equivalent linear dynamic

soil response model, SHAKE91 (Idriss and Sun, 1992). The soil profile at the strong

motion site is shown in Figure 5.10. The computed bedrock motion was then propagated

to the base of the models (24 m below the ground surface) for Ohama Wharves 1 and 2

using SHAKE91. The earthquake time histories recorded at the strong motion site are

shown in Figure 5.11. The transformed time histories for Wharf 1 and 2 are shown in

Figure 5.12 and Figure 5.13, respectively.

x1-

w0(m)

tai0.,-I-

(1.3 ---- Vp (m/sec)SPT

1090 2000.7,1,

0cr)

N-VALUE

10203040500Vs ( m / sec I

400

0

15

20

25

30

lo

j

11,

11

4.

1:

1;..

13 _

(90

1200

L

290

210

-I

1900

lI

liI

1760

, 4

vt

Figure 5.9: Shear Wave Velocity Profile for Ohama Wharf 3 (Iai& Kameoka, 1993)

Page 82: Nason J. McCullough for the degree of Master of Science in

68

Table 5.3: Model Structural Properties for Ohama Wharf 1and 2

Property Wharf 1 Wharf 2Sheet Pile Wall:

Type FSP6L FSP6LMoment of Inertia, I (m4) 8.60e-4 8.60e-4Modulus of Elasticity, E (kPa) 2.20e11 2.20e 1 1

Cross-Sectional Area (m2) 0.0306 0.0306Anchor:

Type Pipe Pile 2 Identical Pipe PilesDiameter (mm) 750 550Thickness (mm) 10 12

Moment of Inertia, I (m4) 1.66e-3 7.84e-4Modulus of Elasticity, E (kPa) 1.00el1 1.00el1Cross-Sectional Area (m2) 0.0236 0.0207

Cable:Modulus of Elasticity, E (kPa) 1.00el1 1.00el1Radius (m) 0.10 0.10

DE( m 1

SOIL TYPE SPYN -VALUEI 203040500

S WAVE P WAVEm/s ) 0 (m/s )

200 400 ECO I0 ec.

' CCIARSE SAND

570

900

II

'SiE=TIE

5

I 0

15

20

25

1.132iiGRAVEL MIXEDWITH SAND

SANDY GRAVEL

. SAND MIXED

I

.5/

SANDY GRAVEL

GRAYEL MIX 0WITH SAND

0

SAND STONE

Figure 5.10: Soil and Velocity Profile at the Akita Strong MotionSite (Iai & Kameoka, 1993)

Page 83: Nason J. McCullough for the degree of Master of Science in

69

North

3.0

M2.0

1.0

2 0.0

e -1.0

§ -2.0'

3.0

3.0

'a 2.0C

s0.0

ti -1.0

IMIRIPPIIIMPIPPIMITSITANIIIITINATIMIF 7/

East

IMIIMPIPIIMIT111113WIT1111.1111INFRNIfillTrir11141111,1,1111MIMEALMia

-2.0

-3.0

3.0

2.0

1.0

0.0

-1.0

-2.0

-3.0

1.0

0.5

0.0

-0.5

-1.0

Vertical

0 5 10 15 20

Time (seconds)

25 30

Figure 5.11: Recorded Time Histories for the 1983 M 7.7 NihonkaiChubu Earthquake at Akita Port

35 40

A I MINI& 111111111% MAI Al A A0111111111111111Wall " w r 11"

5 10 15 20

Time (seconds)

25

Figure 5.12: Input Time History for Akita Ohama Wharf 1

30 35

Page 84: Nason J. McCullough for the degree of Master of Science in

1.0

0.5

C0.0

-0.5

-1.0

70

WilliglillEMIEMIFIVI AI A

5 10 15

Time (seconds)

20

Figure 5.13: Input Time History for Akita Ohama Wharf 2

5.4.1.4 Results of Ohama Wharf 1

25 30

The deformed grid for Wharf 1 is shown in Figure 5.14, with calculated horizontal

and vertical displacements of 0.03 m 0.055 m, respectively. The calculated vertical

displacement, including the post-liquefaction volumetric strain, was 0.055 m. The time

histories of the displacements (not including the post-volumetric strain) are shown in

Figure 5.15. The maximum measured horizontal displacement was 0.05 m. The measured

vertical displacements were not available. The results for Ohama Wharf 1 are consistent

with the measured displacements.

Figure 5.14: Deformed Grid for Akita Ohama Wharf 1(deformations to scale)

Page 85: Nason J. McCullough for the degree of Master of Science in

71

0.01

0.00

-0.01

Et -0.02E

a.es

-0.03

-0.04

-0.05

-0.06

- Horizontal DisplacementVertical Displacement

1\\*NOVO

0 5 10 15 20 25 30

Time (seconds)

Figure 5.15: Time Displacement Plot for Akita Ohama Wharf 1

5.4.1.5 Results of Ohama Wharf 2

The deformed grid for Wharf 2 is shown in Figure 5.16, with the horizontal and

vertical time displacements are given in Figure 5.17. The measured displacements are

presented in Figure 5.6 and Figure 5.18. The maximum horizontal displacement

calculated in FLAC was 1.30 m at the top of the wall, where the measured displacements

had a maximum of about 1.6 m and an average of about 1.1 m. The calculated vertical

displacements in the backfill adjacent to the wall, including and not including post-

liquefaction volumetric strain, were 0.78 and 0.60 m, respectively. The measured vertical

displacements ranged between an average of 0.5 and a maximum of 1.4 m. It should be

noted that the formation of a plastic hinge was evident during post-earthquake inspections

of the bulkhead (Figure 5.6). A plastic hinge developed only to a small degree in the

numerical model. The difference between the measured and calculated profiles could be

contributed to several reasons; 1) inadequacies in the numerical model, 2) lack of

structural integrity in the field due to corrosion, physical imperfections, etc.

Page 86: Nason J. McCullough for the degree of Master of Science in

10 meters

72

Figure 5.16: Deformed Grid for Akita Ohama Wharf 2(deformations to scale)

0.20

0.00

-0.20

-0.40

-0.6000.

-0.80

-Horizontal DisplacementVertical Displacement

5 10 15 20

Time (seconds)

25 30

Figure 5.17: Time Displacement Plot for Akita Ohama Wharf 2

5.4.2 Ishinomaki Port

35

The 1978 MimA 7.4 Miyagi-Ken-Oki Earthquake caused severe liquefaction damage

in the Ishinomaki Port. The case study utilizes the data from Shiomi Wharf (-4.5 m) that

was presented by Iai, Tsuchida and Finn (1985). The general location of the city, strong

Page 87: Nason J. McCullough for the degree of Master of Science in

APPROACH SECTION B SECTION A APPROACH30m 80m 105m 20m

I1

Cm> Horizontal Displacement at the Top of Sheet Pile (m)a) 2I 1

b)

Elevation of Sheet Pile(m)+1

+0.5

Design Elevation (m)+ I

+0.5

-0.5

Figure 5.18: Measured a) Horizontal and b) Vertical Displacementsalong the Face of the Bulkhead (Iai & Kameoka, 1993)

-0.5

73

motion site and Shiomi Wharf (-4.5 m, Site D), are shown in Figure 5.19. The bulkhead

profile is presented in Figure 5.20.

5.4.2.1 Soil Conditions

Four borings were conducted along the wharf, one previous to (CB-10 dashed), and

three after (CB-2, CB-6, CB-10 solid) the earthquake (Figure 5.21). The boring before the

earthquake was used to determine the soil properties to approximately 14 m below the

ground surface, where the combination of the other three borings were used for the deeper

soil profile. The generalized profile used in the FLAC model consisted of two soil layers,

a medium to fine sand backfill extending from the ground surface to a depth of 15 m, and

underlain by a sandy silt that extends to the base of the model (23 m). A rubble mound is

located directly behind the wall and in front of the anchor. The mound behind the wall is

small and was not modeled, but the mound in front of the anchor was incorporated into

the FLAC model. The uncorrected SPT values were input directly from the boring logs.

The modeled soil properties are given in Table 5.4. The shear wave velocities were

estimated from the SPT values.

Page 88: Nason J. McCullough for the degree of Master of Science in

74

39°141°E Ofunato 142°

N

IshinomakiSend 2i

38°

I

Fault plane

hiogama XjEpicenter

Pacific Ocean

Strong-motionAccelerogroph Site

Kaihoku Bridge

ISHINOMAKI CITY

ISHINOMAKI PORT.Site-C(Non-Liquefaction)

Site -D(Liquefaction

Site- B(Liquefaction )Site-F(Liquefaction

Site -A(Liquefaction) Site -E(Non Liquefaction

ISHINOMAKI BAY 9 icrr/

mi

Ishinomaki Fisting Port000ctooorn000cloocIT

Figure 5.19: Layout of Ishinomaki Port during the M 7.4 Miyagi-Ken-Oki Earthquake (Iai et al, 1985)

K.P+3.00

K.P +1.50

15.000 14.20

325 +328 +3.50

li emu Hugh Tensile Stren

020.0(TP-0.8745)01.0.0(KP-0.025)

K.P :4.50

TO

Ao:.

-10.00

t :20.80mth Steel

+250

Steel sheet pile anchorage1) type 400x 10x 10.5 =4.0m

Steel sheet pileIIA type 400x 150x 13.1 C: 11.5m

400

-1.50

Figure 5.20: Shiomi (-4.5m, Site D) Bulkhead Profile (Iai et al.,1985)

Page 89: Nason J. McCullough for the degree of Master of Science in

'ii

ra

CB-2

Soil Tyne

au-value(kgt/cmt)-0-0.6 09 1.0 1.2 14 1.6

'N -Value ___w___'

0 10 20 30 40 50

ffIr41333:22

O'ir

:..4.-,

''

:,--4

1.:''.1

;''',,;:..0::,.,

San;ii'')

(Medium

Sandy

c'.

'''''''''d1

Asphalt

11111111111111/11111111111M1111111111E11111111111111111111111111111111111111111111111111111111111111111E11111111111111

Scm soli

(F'n.sad)

11M 111111111111111111111111111111111111111111111111111111111mo111111111111

1aiCloy

(Script 1111111111v

(a ) CB -2

N - volu4

1 0 203040 50

93

Brown

Fine toMedium Sand

Silt mixedwith sells

2( 03m4,N4; I

11 11140m N.6

1 1 1 1

60m N =13(3,5,5)

8I

I /0 m IN= 16

(5,5,6

IIIII100m N =13(4,4,5)

11111(20w N =16(5,7,4)

4.5mI

N.I5(5,4,6)1

6.2m N.8( IIIII1

181m N.10( 2,2,6)

20.3m N 6

1 1 1 1

2.3m N.5

III III24.0m N.5

2825m N.7

30.0m N 8(1,1 1.11

(b) CB-6

CB-I0Soil Type

Block Asphalt

0.94

N - value

20 3040 50

75

GravelGravel Stonerk Groy

MediumStone

2., Medium send- Fine sand

20m N.4N.2

II I

4.0rn N.7(2,3,2)"N.13 I I

%,6.0m NOB)1(2,4,2)

-506

N.I6

'11111110m N.)

(3,4,4)

o Very fine xnd

N

Dark grey N

-14

-1506

-12

Fine sand

lue gray

Sandy Silt

Dark gray

Medium sandmixed with

iBlue gr ayshells

Tirrwith shells

--e-Dark gray

10.0m N.15(4,5,6)11

111111112.0m N.18

1 1 1 1 1

140m N.16

8m WI18 III I5.

(4,6,6

Measured112967.

Measured in 97

18.0m N =S

11 112t/Om N.6

220m

24.0m

27.5m N-9N.8

29.0

(4,23)

/4 4

11111m N.4

Figure 5.21: Boring logs at Shiomi Wharf (Iai et al., 1985)

Table 5.4: Model Soil Properties for Ishinomaki Port

Cc) CB -10

Shionzi Wharf (-4.5m,Layer 2

Site D)RubbleLayer 1

Density (kg/m3) 1500 1500 1800

Friction Angle (0) 36 36 42

Poisson's Ratio 0.30 0.30 0.20Porosity 0.40 0.40 0.40

Page 90: Nason J. McCullough for the degree of Master of Science in

76

5.4.2.2 Structural Properties

The wall section was constructed from Nippon FSP3A sheet pile sections, and the

anchor was constructed from Nippon FSP2 sheet piles sections. The total height of the

sheet pile wall is 13 m, with 5.5 m embedment, the anchor embedment is 4 m, with a tie

rod length of 20 m. The spacing of the tie rod was unknown, and assumed to be one

meter. This introduces an error into the calculated displacements because of the

uncertainty in the elastic behavior of the tie rod system. Because the backfill within the

model was expected to liquefy, the friction between the sheet pile sections and soil was

assumed to be zero. The modeled structural properties are presented in Table 5.5.

Table 5.5: Model Structural Properties for Ishinomaki Port

Property Shiomi Wharf (-4.5m, Site D)Sheet Pile Wall:

Type FSP3AMoment of Inertia, I (m4) 2.281e-4Modulus of Elasticity, E (kPa) 2.20e1 lCross-Sectional Area (m2) 0.0186

Anchor:Type FSP2Moment of Inertia, I (m4) 8.740e-5Modulus of Elasticity, E (kPa) 2.20e1 lCross-Sectional Area (m2) 0.0153

Cable:Modulus of Elasticity, E (kPa) 2.00e11Radius (m) 0.10

5.4.2.3 Earthquake Motion

The strong motion time histories were recorded on bedrock, and were assumed to be

the same as the bedrock motions beneath the site. The transformed time history was

propagated from bedrock to the base of the model (23 m below the ground surface) using

SHAKE91. The earthquake time histories recorded at the strong motion site (Kaihoku

Page 91: Nason J. McCullough for the degree of Master of Science in

77

Bridge) are shown in Figure 5.22. The transformed time history for Shiomi wharf is given

in Figure 5.23.

3.0

a 2.0

S42W

3 -2.0

-3.0

3.0

2.0

1.0

0.0

-1.0

-2.0

3.0

0

W42N

1111111111,1111111111MITIFI1111,IN 111 WAIN "Nil 'UM

1.5

1.0

0.5

0.0

-0.5

-1.0

-1.5

10 12

Time (seconds)

14 16 18 20

Figure 5.22: Recorded Time Histories at Ishinomaki Port

.1 Ail Ai. 11111111 7 r. V

2

Time (seconds)

8

Figure 5.23: Input Time History for Ishinomaki Wharf

10 12

Page 92: Nason J. McCullough for the degree of Master of Science in

78

5.4.2.4 Results of Ishinomaki Shiomi (-4.5 m) Wharf

The deformed grid at 12 seconds is given in Figure 5.24. The calculated horizontal

and vertical time displacements are shown in Figure 5.25. The time displacement plot

extends for 50 seconds, where the earthquake motion was only 12 seconds long, and it

should be noted that the displacements continued to increase after 12 seconds at a steady

rate. This is theorized to be due to the modeling limitation of not allowing the excess pore

pressures to dissipate. If excess pore pressures were allowed to dissipate, the

displacements would likely taper off after 12 seconds. Therefore, the calculated

displacements are from 12 seconds. The calculated horizontal displacement is 0.60 m and

the vertical displacements, including and not including post-liquefaction volumetric strain

are 0.30 and 0.25 m, respectively.

The measured results are presented in Figure 5.26 and Figure 5.27, and indicate a

large variation in the horizontal deflections along the top of the bulkhead. This is

theorized to be partly due to the end effects of the bulkhead. It was also noted by Iai et al.

(1985) that a landslide of unknown extent was noted at the site, and may have also caused

the large variations in the displacements. The maximum measured displacement of the

wall was 1.16 m, with an average horizontal displacement of approximately 0.60 m. The

measured vertical displacements ranged between an average of 0.05 and a maximum of

0.10 m.

Figure 5.24: Deformed Grid for Ishinomaki, Shiomi Wharf(deformations to scale)

Page 93: Nason J. McCullough for the degree of Master of Science in

79

0.20

0.00

-e- -0.20

-0.40Etv

ct.) o -0.60a.co

6 -0.80

-1.00

-1.20

Horizontal Displacement

Vertical Displacement

0 10 20 30

Time (seconds)

40 50

Figure 5.25: Time Displacement Plot of Ishinomaki, Shiomi Wharf

5.4.3 Kushiro Port

The MRAA 7.8 Kushiro Earthquake occurred on January 15, 1993, and caused severe

liquefaction damage to the Port of Kushiro at Kushiro City. The majority of the data for

this case study was taken from Iai et al. (1994) and Ueda et al. (1993). The location of

Kushiro City is shown in Figure 5.28. The case study used two wharves from Kushiro

Port, indicated as Site C and Site D in Figure 5.29. The location of the strong motion site

is also indicated in Figure 5.29. The geometry of Site C is shown in Figure 5.30, and

consists of a sheet pile wall, backed by a rubble mound, and supported by a battered pile

anchor, located 20 m into the backland. The geometry of Site D is shown in Figure 5.31,

and consists of a steel pipe wall, backed by a rubble mound, and supported by a

continuous sheet pile wall, 30 m into the backland. Site D was also an area of soil

improvement prior to the earthquake, including both sand compaction piles and gravel

drains.

Page 94: Nason J. McCullough for the degree of Master of Science in

Boring CB-10

N No.16

6 No.15

23 No.1 4

40 No.l 3

37 No.12

EOo

co

Oto

ONo

Figure 5.26: Measured Horizontal Deformations at Shiomi Wharf(-4.5 m), displacements in mm (Iai et al., 1985)

(2'az 6z az az dz az 6z 6z az 6zII3

z 6z 6z oz 6zID6

ID

eizDesign I

M+3.0004.2.900+2.8004-2.700+2_600__

Figure 5.27: Measured Vertical Displacements at Shiomi Wharf (-4.5 m) (Iai et al., 1985)

levation

80

Page 95: Nason J. McCullough for the degree of Master of Science in

x Epicenter1993. 1.15(M = 7.8)

Pacific Ocean

Closer View(a) Location of Kushiro

Figure 5.28: Location of Kushiro Port (Iai et al., 1994)

f/EOSI PortOrI

Site ASite C

KUSHIRO PORT

0 500 I000m

West PortSite DSite SPt,

Figure 5.29: Layout of Kushiro Port (Iai et al., 1994)

ormer Beach Line

81

Page 96: Nason J. McCullough for the degree of Master of Science in

82

*2.7

L.W. t. 0.0

20.0 Unit (ml0

% 2.9

-7.5

E

I-

Rubble

Sheet Pile

Tie rod SS 41 065

. Filled Sand.

OOriginal Ground

co

inNo.

-14.5before eorthquokeotter eorthquoke

L22.8of

*2.2.1.0

tft

Steel Pipe

-10.3-14.3

Figure 5.30: Geometry of Kushiro Port Site C Wharf (Iai et al.,1994)

H.W.L1.54 L 0.0

3.0

-12.0

30.0

20.02% 3.4

4.rakraiorriAa

22.7

Tie- Rod ctcl 98

2.00-J

Grovel ()rain# 400mrr

u ,-

Sond Cornpoct ion Pi le 2.

# .700mm .&

Unit(m)

Figure 5.31: Geometry of Kushiro Port Site D Wharf (Iai et al.,1994)

Page 97: Nason J. McCullough for the degree of Master of Science in

83

5.4.3.1 Soil Conditions

The boring logs for Site C and Site D are shown in Figure 5.32 and Figure 5.33,

respectively. The simplified soil conditions for the numerical analysis at Site C consisted

of 10 m of hydraulically backfilled sand, underlain by the naturally deposited sandy soil

to the depth of the model (26 m). At Site D, the soil compaction was modeled as

extending infinitely into the backland. The soil layers where simplified as a coarse sand

extending to a depth of 6 m below the surface, underlain by 12 m of fine sand, which is

underlain by another layer of coarse sand to the depth of the model (32 m). The area of

soil compaction was modeled as a uniformly compacted zone, individual gravel drains

and sand compaction piles were not modeled. The rubble mounds were modeled in both

Site C and Site D as indicated in the wharf geometries. Model soil properties used in the

analyses are given in Table 5.6. The SPT values were input as shown in the boring logs,

with the shear wave velocities estimated from the SPT values.

De p thml

Soil Type SPT-N Value0 10 20304050

'Elevotiols2.34rn

-5

-10

-15

-20

Sand

GravellySand

FineSand

GravellySondFineSand

Course

Sand

SandandGrave .

Figure 5.32: Soil Profile at Kushiro Port Site C (Iai et al., 1994)

Page 98: Nason J. McCullough for the degree of Master of Science in

84

Depth I SPT - N Voluecm) Soil Type0

01020304050

levationl ", ?"'ZI42.94m J

Coarse Sond

Coarse Sond -mixed with

Gravelr,

Fine Sand -1--10 :t Fine Sand "":

:and Silt_ Fine Sond

mixed withShells

-15 Shelly Sond- Coarse Sand

mixed withe. Grovel

Coarse Sond beforecompaction__:

:

Figure 5.33: Soil Profile at Kushiro Port Site D (Iai et al., 1994)

Table 5.6: Model Soil Properties for Kushiro Port

Site C Site DLayer 1 Layer 2 Rubble Layer 1 Layer 2 Layer 3 Rubble

Density (kg/m3) 1300 1500 1800 1500 1400 1500 1800Friction Angle (q) 32 42 42 45 40 45 42Poisson's Ratio 0.30 0.30 0.20 .0.30 0.30 0.30 0.20Porosity 0.40 0.40 0.40 0.40 0.40 0.40 0.40

5.4.3.2 Structural Properties

The model structural geometry of Site C consisted of a sheet pile wall (Nippon SP-

Z25) anchored with battered pipe piles. The wall is 17.2 m tall, with 7 m embedment, and

the left and right anchors extend to a depth of 13 and 17 m below the ground surface,

respectively. The tie rod and anchor spacing were unknown, and assumed to be 1 m. The

model wall for Site D consisted of a steel pipe pile wall, anchored with a continuous sheet

pile wall (Nippon FSP4). The tie rod spacing is 1.98 m. The radius of the tie rod was

Page 99: Nason J. McCullough for the degree of Master of Science in

85

unknown for Site D, and conservatively assumed to be 1 m. The model structural

properties are given in Table 5.7.

Table 5.7: Model Structural Properties for Kushiro Port

Property Site C Site DSheet Pile Wall:

Type FSPZ25 Pipe PileDiameter (mm) -- 914.4Thickness (mm) -- 16

Moment of Inertia, I (m4) 3.824e-4 4.804E-3Modulus of Elasticity, E (kPa) 2.20e11 2.20e11Cross-Sectional Area (m2) 0.0236 0.0460

Front Anchor:Type Pipe Pile FSP4Diameter (mm) 400 --Thickness (mm) 9 --Moment of Inertia, I (m4) 2.262e-4 3.86e-4Modulus of Elasticity, E pcPa) 1.00el1 2.20e11Cross-Sectional Area (m) 0.0113 0.0240

Back Anchor:Type Pipe Pile

_--

Diameter (mm) 700mm --Thickness (mm) 12mm --Moment of Inertia, I (m4) 1.616e-3 --Modulus of Elasticity, E (kPa) 1.00e 11 --Cross- Sectional Area (m2) 0.0264 --

Cable:_ _

Modulus of Elasticity, E (kPa) 1.00ell 1.00e 11

Radius (m) 0.10 1.0

5.4.3.3 Earthquake Motion

The strong motion recording station included a surface and downhole seismograph

array. The downhole record was recorded at a depth of 77 m. The depth to bedrock at the

strong motion site was unknown, and assumed to be 77 m. The soil profile at Site C and

Site D was unknown for depths greater than approximately 20 m, but because the strong

motion site and wharf sites are in close proximity, the same profiles were used for depths

exceeding 20 m. The soil profile at the strong motion site is shown in Figure 5.34. The

Page 100: Nason J. McCullough for the degree of Master of Science in

86

Elevo-tionfrn)

SoilType

SPT DensityN-volue ( 1 /m3)

1

50 K) 203340 15 2.0

3 6

41.6

0-

is.In Groun.

SurfaceSeismograph

i 15

I 0

15

0

30

40

-45 -

-

r=--. 11111

1111 11i..

i. 111

6.,:...:

.1

---qv:

M.

Ali d.

qualg11 Ii

4....ili

1bud r

4

-70

.- 73.

OownholeSeismograph

oorseSand

r4.ineand

ravel lySand

Silt

Figure 5.34: Soil Profile at the Kushiro Port Strong Motion Site(Iai et al., 1994)

Page 101: Nason J. McCullough for the degree of Master of Science in

87

recorded time histories are given in Figure 5.35 and Figure 5.36 for the surface and

downhole array, respectively. It is evident from the North record at the surface, that at

approximately 30 seconds, the predominant period of the ground motion increases. It has

been noted by Iai et al. (1995) that this is probably due to the dilatancy behavior of the

sandy soil. Regardless, the surface histories are unsatisfactory for use in the numerical

model, therefore the downhole histories were used. The recorded downhole acceleration

time histories were propagated through the soil to the depth of the model (26 and 32 m

below the ground surface for Sites C and D, respectively) using SHAKE91. Because the

orientations of Site C and Site D are almost identical, the same transformed time history

was used for both sites. The transformed record is given in Figure 5.37

5.0

t7, 3.0

1.00

-1.0

-3.0

5.0

5.0

fi; 3.0E

0-1.0

-3.0

5.0

1.0

5.0

3.0

1.0

-1.0

-3.0

5.0

North

East

=7NIZ,1:1C:!,

Vertical

19111.1i i i ,1.1 1 71! 771rrili I, 1 I, ILL JJahl

10 20 30

Time (seconds)

40 50 60

Figure 5.35: Earthquake Motions Recorded at the Ground Surfaceat Kushiro Port on January 15, 1993

Page 102: Nason J. McCullough for the degree of Master of Science in

5.0

a 3.0

1.0

e 1.0

-3.0

5.0

5.0

3.0

1.00

1.0

8 3.0.0

5.0

5.0

3.0

1.0

-1.0

-3.0

5.00

88

North

East

Vertical

2.0

1.0

0.0

-1.0

-2.0

10 20 30

Time (seconds)

40 50

Figure 5.36: Earthquake Motions Recorded at a depth of 77 metersat Kushiro Port on January 15, 1993

60

10 15

Time (seconds)

20 25

Figure 5.37: Input Time History for Kushiro Port Site C and Site D

30

Page 103: Nason J. McCullough for the degree of Master of Science in

89

5.4.3.4 Results of Site C

The deformed grid is shown in Figure 5.38. The calculated vertical and horizontal

time displacements at the top of the wall are shown in Figure 5.39. The measured

horizontal and vertical displacements at the top of the wall were approximately 0.6 and

0.4 m, respectively. The calculated horizontal and vertical displacements are 0.45 and

0.25 m, respectively. The post-liquefaction volumetric strain was negligible, due to

limited pore pressure generation directly adjacent to the sheet pile wall.

Figure 5.38: Kushiro Site C Deformed Grid (deformations to scale)

0.10

0.00

-0.10

a-0.30

-0.40

-0.50

Displacement

Displacement

- Horizontal

A

Vertical

i..-

A.V.V

0 5 10 15

Time (seconds)

20

Figure 5.39: Time Displacements for Kushiro Site C

25 30

Page 104: Nason J. McCullough for the degree of Master of Science in

90

5.4.3.5 Results of Site D

The deformed grid is shown in Figure 5.40. There were no reported deformations at

Site D, but it was noted that no damage was observed. The calculated horizontal and

vertical time displacements are shown in Figure 5.41. The horizontal and vertical

displacements were calculated as 0.070 and 0.085 m, respectively. The post-liquefaction

volumetric strain was negligible, due to limited excess pore pressure generation in the

improved backfill soils.

Figure 5.40: Kushiro Site D Deformed Grid (deformations to scale)

0.01

0.00

-0.01

-0.02

g -0.03

m -0.04

0al -0.05ao -0.06

-0.07

-0.08

-0.09

-0.10

Horizontal Displacement

Vertical Displacement

0 5 10 15 20 25 30

Time (seconds)

Figure 5.41: Time Displacements for Kushiro Site D

Page 105: Nason J. McCullough for the degree of Master of Science in

91

5.4.4 Discussion of Numerical Validation Results

The results of the calculated permanent horizontal displacements at the top of the

wall and the vertical displacements in the backfill adjacent to the sheet pile wall are

tabulated in Table 5.8. Despite the numerical uncertainties and limitations noted in the

previous sections, the calculated and measured results compare rather well. The

calculated horizontal displacement for Akita Wharf 1 compares well with the measured

value, and the calculated vertical displacement appears to be acceptable, even though

there are no measured results. The horizontal and vertical displacements for Akita Wharf

2 fall within the range of measured values. The horizontal displacement for Ishinomaki

compares well with the measured value, while the vertical displacement is over-predicted.

The horizontal and vertical displacements for Kushiro Site C are both under-predicted,

while the calculated displacements for Kushiro Site D are fairly well predicted.

The horizontal displacements are on average under-predicted by approximately 23%

(as compared to the average of the maximum and average measured displacements). The

vertical displacements are on average under-predicted by approximately 29% (as

compared to the average of the maximum and average measured displacements), except

for Ishinomaki, which is largely over-predicted.

Several possible reasons for the variations in the calculated and measured values

include:

1) soil properties, the majority of which were determined from correlations with

standard penetration test results,

2) constitutive model, which is Mohr-Coulomb based with the addition of a plastic

flow rule,

3) input earthquake motion, which was determined using the simplified dynamic

soil analysis program SHAKE91, and

4) pore pressure generation, which was a numerical variation on the simplified

method develop by Seed and his co-workers.

Page 106: Nason J. McCullough for the degree of Master of Science in

92

There are uncertainties in the numerical model, but it is also noted that there are

numerous uncertainties in the field case histories. There is also much variability in the

lateral and vertical deformations observed in the field (maximum displacements were

between 50 to over 100% greater than the average displacements), therefore the

numerical model is judged to provide useful, representative results for the dynamic

behavior of anchored sheet pile bulkheads. Due to the satisfactory modeling of the

validation case histories, the numerical model was used to perform parametric studies on

sheet pile bulkheads, as described in the following chapter.

Table 5.8: Summary of Case Study Results

CalculatedDisplacements (m)

MeasuredDisplacements (m)

.,

Horizontal Vertical* Horizontal Vertical

Akita Wharf 1 0.03 0.06 0.05 not reported

Akita Wharf 2 1.20 0.78 1.1 to 1.6 0.5 to 1.4

Ishinomaki Shiomi 0.60 0.30 0.60 to 1.16 0.1

Kushiro Site C 0.45 0.25 0.6 0.4

Kushiro Site D 0.07 0.09no damage

(-0.05) not reported

the vertical displacements include post-liquefaction volumetric strainthe range of measured displacements represent displacementaverage to displacement.

Page 107: Nason J. McCullough for the degree of Master of Science in

93

6 PARAMETRIC STUDY ON THE SEISMIC PERFORMANCE OF SHEETPILE BULKHEADS

6.1 INTRODUCTION

After modeling the field case histories to calibrate the numerical model and evaluate

uncertainties in the computed deformations, the project focused on an extensive

parametric study using the calibrated model for anchored sheet pile bulkheads. The

parametric study included several analyses on the effects of varying; 1) zone of soil

improvement, 2) density of the backfill, 3) length of the tie rod, 4) stiffness of the sheet

pile wall, 5) the depth of sheet pile embedment, and 6) characteristics of the earthquake

motions. The study focused on walls designed using standard pseudo-static procedures

for anchored sheet pile walls without pore pressure generation (Ebeling and Morrison,

1993).

The walls were designed with heights ranging from 7.5 to 15 m, and horizontal

seismic coefficients ranging from 0.1 to 0.16. In most cases, seismic coefficients in

excess of 0.2 were not possible for the simplified walls with single anchor systems due to

the impracticable section properties required of the sheet piles. The vertical seismic

coefficient was assumed to be zero. The design followed the steps as outlined in Ebeling

and Morrison (1993), and included a factor of safety of 1.2 applied to the passive

resistance of the dredge soils in front of the sheet pile wall. Table 6.1 presents the results

of the design for the bulkheads utilized in this study.

The tie rod length was designed such that the active failure plane behind the wall

does not intersect the passive failure plane in front of the anchor. The active and passive

failure planes were assumed to originate at the base of the wall and anchor, respectively.

The anchor was assumed to be a continuous sheet pile wall, with the same stiffness as the

bulkhead wall, and the tie rod spacing was assumed to be two meters. The distance from

the ground surface to the ground water table and the tie rod was assumed to be 27 percent

of the wall height (I/), which is on average the elevation of the tie rod for many field

bulkheads.

Page 108: Nason J. McCullough for the degree of Master of Science in

94

Table 6.1: Bulkhead Properties for Parametric Study

Wall Height,H (m)

HorizontalSeismic

Coefficient, kh

Depth ofEmbedment,

D (m)

AnchorEmbedment

(m)

Tie RodLength, A (m)

Moment of4Inertia, I (m)

7.5 0.10 2.9 4.5 16 3.45x10-57.5 0.16 3.7 6.0 22 7.43 x10-515.0 0.10 5.4 7.6 30 3.82x1015.0 0.16 6.9 9.9 40 8.11x104

The foundation and backfill soils are both cohesionless materials, each layer modeled

with uniform density (except for soil improvement zones within the backfill). The range

of densities used can be correlated with stress corrected penetration resistances ((N1)60) of

a) 25 blows/30 cm in the foundation soil, b) 10 to 25 blows/30 cm in unimproved

backfill, and c) 30 blows/30 cm in improved backfill. The soil improvement was modeled

as providing a uniform increase in soil density throughout the zone of treatment. This is a

simplification of the actual pattern of densification and soil stiffness variations that would

be expected from vibro-compaction, stone columns or other densification methods,

however it adequately modeled the general soil density increase due to improvement.

Five earthquake motions covering the magnitude range of engineering interest

(M 6 to 8) were selected for the parametric study (Table 6.2). The selected acceleration

time histories are slightly conservative in the sense that each one is characterized as

having greater than average duration for that magnitude, thereby yielding slightly

conservative displacement results. Each record was scaled to different maximum

acceleration values ranging from 0.1 to 0.4 g. Plots of the unsealed records are provided

in Appendix A.

In order to account for the duration of the earthquake motions a normalized ground

motion intensity parameter has been developed. This parameter is defined as the

maximum horizontal acceleration within the backfill at the elevation of the dredge line

(A max@dredge) divided by the appropriate magnitude scaling factor (MSF) given in Table

2.2. It is recommended that if a site specific seismic study is not performed to determine

A max@dredge 5that the peak ground surface acceleration be reduced using the reduction

Page 109: Nason J. McCullough for the degree of Master of Science in

95

factor (rd) developed for estimating the variation of cyclic shear stress (or acceleration)

with depth (e.g., Seed and de Alba, 1983). The values of rd for 15 and 7.5 m walls are

approximately 0.78 and 0.95, respectively. It should be noted that the reduction factor

was developed using a linear one-dimensional dynamic soil response method and will

only yield approximate acceleration values for the two-dimensional soil-structure

interaction applications discussed herein.

Table 6.2: Parametric Study Earthquake Motions

Earthquake MomentMagnitude

RecordedAm (g)

1984 Morgan Hills EQ - Gilroy #4 6.0 0.221989 Loma Prieta EQ Capitola Fire Station 6.9 0.401989 Loma Prieta EQ Salinas 6.9 0.111992 Landers EQ - Joshua Tree Fire Station 7.4 0.271985 Michoacan Mexico EQ 8.1 0.39

Notable conditions in the parametric numerical models include:

1) no pore pressure generation in the foundation or improved backfill soils

((NI)60-30),

2) the soil improvement extends vertically to the base of the backfill (i.e. the dredge

line elevation), and

3) the tie rod and water level are located at an elevation that is 27 percent of H from

the top of the wall.

Parametric studies were then formulated to examine the effects of varying several

parameters when subjected to seismic motions. The parametric studies included the

following (with results and discussions presented in the following sections and tabulated

results in Appendix B):

Page 110: Nason J. McCullough for the degree of Master of Science in

96

1) depth of sheet pile embedment (Section 6.2),

2) tie rod length (Section 6.3),

3) sheet pile stiffness (Section 6.4),

4) density of the liquefiable backfill (Section 6.5), and

5) zone of soil improvement (Section 6.6).

6.2 DEPTH OF SHEET PILE EMBEDMENT

A study was performed to determine the effect on the lateral displacements by

varying the depth of embedment from the value determined using the 7.5 m bulkheads

designed with a kh of 0.1. The depths of embedment were halved and doubled from the

design value of 2.9 m, and the horizontal displacements at the top of the wall were

monitored. The 1984 Morgan Hills Earthquake and the 1985 Mexico Earthquake were

used. The entire backfill was modeled as improved, with (A/d60 values of 30 blows/30 cm.

The monitored horizontal displacements for varying the depths of embedment are

presented in Figure 6.1. The depth of embedment (D) is normalized by the depth of

embedment calculated using the Mononobe-Okabe method (Dm_ 0) and the earthquake

intensity at the dredge line has also been normalized by the magnitude scaling factor

(MSF). Even though there are a limited number of data points, it is noted that for these

specific cases, the depth of embedment has very little effect on the horizontal

displacements at the top of the wall.

6.3 TIE ROD LENGTH

A study was performed to determine the effect on the lateral displacements by varying the

tie rod length from the value determined using the 7.5 m bulkheads designed with a kh of

0.1. The tie rod lengths were halved and doubled from the design value of 16 m, and the

horizontal displacements at the top of the wall were monitored. The 1984 Morgan Hills

Earthquake, 1992 Landers (Joshua Tree) Earthquake and the 1985 Mexico Earthquake

were used. The entire backfill was modeled with (N 1)60 values of 30 blows/30 cm.

Page 111: Nason J. McCullough for the degree of Master of Science in

1.50

1.25

1.00

0.75

0.50

0.25

0.00

Earthquake (Amax@dredge/MSF)Morgan Hills (0.26)

Mexico (0.94)

0.0 0.5 1.0 1.5 2.0

Normalized Depth of Embedment ( DID wo

Figure 6.1: Top of Wall Displacements for the Depth ofEmbedment Study

2.5

97

The monitored horizontal displacements for varying the tie rod length are presented

in Figure 6.2. The tie rod length (A) is normalized by tie rod length calculated using the

Mononobe-Okabe method (Am_0), as defined in Section 6.1. The earthquake intensity at

the dredge line has been normalized by the magnitude scaling factor (MSF). Even though

there are a limited number of data points, it is noted that for these specific cases, the tie

rod length has a large effect on the horizontal displacements at the top of the wall. This is

especially observed when the normalized tie rod length is less than 1.0.

6.4 SHEET PILE STIFFNESS

A study was performed to determine the effect on the lateral displacements by

varying the sheet pile stiffness (El) from the value determined using the 7.5 m bulkheads

Page 112: Nason J. McCullough for the degree of Master of Science in

1.50

1.25

1.00

0.75

0.50

0.25

0.00

Earthquake (Amax@dredge/MSF)

Morgan Hills (0.29)

Joshua Tree (0.50)

Mexico (0.94)

0.0 0.5 1.0 1.5 2.0

Normalized Tie Rod Length ( A/A )

Figure 6.2: Top of Wall Displacements for the Tie Rod LengthStudy

2.5

98

designed with a kh of 0.1. The stiffness was halved and doubled from the design, and the

horizontal displacements at the top of the wall were monitored. The 1984 Morgan Hills

Earthquake and the 1985 Mexico Earthquake were used. The entire backfill was modeled

as improved, with (Nj)60 values of 30 blows/30 cm for two sets of runs. For the other run

(hollow diamonds in the figure) the backfill was modeled as partially improved to a

normalized soil improvement value ( n = SI/(H+D) ) of 2.0, while the remaining

backland had a (Nj)60 values of 10 blows/30 cm. The normalized soil improvement

parameter is discussed in more detail in Section 6.6.

The monitored horizontal displacements for varying the stiffness are presented in

Figure 6.3. The stiffness (El) is normalized by the stiffness calculated using the

Mononobe-Okabe method (E/m_o) and the earthquake intensity at the dredge line has also

Page 113: Nason J. McCullough for the degree of Master of Science in

99

been normalized by the magnitude scaling factor (MSF). Even though there are a limited

number of data points, it is noted that for these specific cases, the sheet pile stiffness has a

negligible effect on the horizontal displacements at the top of the wall. It should be noted

that if the plastic moment of the sheet pile wall is exceeded, the deformations at the top of

the wall will increase dramatically.

1.50

1.25

1.00

0.75

0.50

0.25

0.00

6

Earthquake, Wall Height, (Amax@dredge/MSF)Morgan Hills, H=7.5m (0.29)

o Morgan Hills, H=7.5m, n=2 (0.29)

Mexico, H=7.5m (0.94)

8 8

0.0 0.5 1.0 1.5

Normalized Stiffness ( El/EI m.o )

2.0

Figure 6.3: Top of Wall Displacements for the Sheet Pile StiffnessStudy

6.5 PENETRATION RESISTANCE OF THE BACKFILL

2.5

A study was performed to determine the effect on the lateral displacements by

varying the normalized penetration resistance ((N/)60) of the backfill. The study utilized

the 7.5 m bulkheads designed with a kh value of 0.1. The (N/)60 values were varied

between 10 and 25 blows/30 cm, and the horizontal displacements at the top of the wall

Page 114: Nason J. McCullough for the degree of Master of Science in

100

were monitored. The 1984 Morgan Hills Earthquake, 1992 Landers (Joshua Tree)

Earthquake and the 1985 Mexico Earthquake were used. The entire backfill was modeled

as uniformly dense.

The monitored horizontal displacements for varying the penetration resistance are

presented in Figure 6.4. Even though there are a limited number of data points, it is noted

that for these specific cases, the penetration resistance of the backfill has a large effect on

the horizontal displacements at the top of the wall. An especially large increase is noted

for the Morgan Hills record when the (Nj)60 values were less than 15 blows/30 cm. The

analyses for the Joshua Tree and Mexico earthquakes with (Nj)60 values less than 15 and

20 blows/30 cm, respectively, experienced complete flow failures of the bulkheads and

caused instability in the numerical models.

4.5

4.0

........ 35E

g 3.0Ea)

C 2.5

6 2.0To

o 1.5N

T-*0

1 . 0

0.5

0.0

Earthquake (Amax@dredge/MSF)

Morgan Hills (0.29)

Joshua Tree (0.50)

Mexico (0.94)

5 10 15 20

Penetration Resistance (blows/30 cm)

Figure 6.4: Top of Wall Displacements for the PenetrationResistance of the Backfill Study

25 30

Page 115: Nason J. McCullough for the degree of Master of Science in

101

6.6 SOIL IMPROVEMENT

The largest parametric study was to determine the effect on the lateral displacements

by varying the zone of soil improvement. The study utilized the 7.5 and 15 m bulkheads

designed with kh values of 0.1 and 0.16. The standard penetration of the improved and

backfill soils was 30 blows/30 cm, and 10 blows/30 cm, respectively. All five earthquake

motions were used, with each one scaled to three different peak accelerations.

A typical diagram of the soil improvement parameters is presented in Figure 6.5. The

distance from the wall to the most landward portion of the improved soil region (SI) has

been normalized by the total wall height (H+D). The resulting factor, n, is given as the

normalized width of soil improvement (SI/(H+D)). The monitored horizontal

displacements for varying n are presented in Figure 6.6. The data is plotted as ranges of

normalized earthquake intensity because of the numerous data points with varying

intensities. There is a definite trend noted in Figure 6.6 that as the amount of soil

improvement decreases, the displacements increase. The scatter within each intensity

range is due to the variations in the earthquake motions that were not accounted for in the

normalized factor (i.e. frequency content). There is no distinction made in the plot

between the different wall heights and different kh values due to the number of data

Figure 6.5: Definition of Soil Improvement Variables

Page 116: Nason J. McCullough for the degree of Master of Science in

102

points. It should be noted that even without a distinction between different wall heights

and designs, the scatter in the data for each range is quite small, given the wide range of

earthquake motions used in the analyses.

1.6

1.4

1.2

1.0

0.8

0.6

0.4

0.2

0.00

0

Amax@dredge/MSFo 0.0 to 0.2

0.2 to 0.4o 0.4 to 0.6

0.6 to 0.8o0.8to1.1

9

1 2 3 4 5 6

Normalized Soil Improvement, n ( S1/(H+D) )

Figure 6.6: Horizontal Displacments at the Top of the Wall forVariations in the Zone of Soil Improvement

7 8

6.7 COMPARISON OF THE PARAMETRIC STUDY WITH AN EXISTING DEFORMATION-BASED DESIGN METHOD

The results of the parametric studies which included non-liquefiable soils (the entire

backland was improved) can be compared to the design chart that has been proposed by

Dennehy (1985) and Gazetas et al. (1990). The location of four points was determined

using the method outlined in Gazetas et al. (1990). Figure 6.7 presents the comparison

between the parametric study (for non-liquefiable soils) plotted as solid stars (with the

Page 117: Nason J. McCullough for the degree of Master of Science in

103

calculated displacements in parenthesis) and the design chart presented by Dennehy and

Gazetas. One point (center-right) is the standard 7.5 m wall presented in Table 6.1. Two

points (top-right and bottom-right) are from the parametric study varying the length of the

tie rod anchor, and the last point (left-center) comes from the parametric study varying the

depth of embedment.

It is evident from Figure 6.7 that the computed deformations vary significantly at

each data point, and that some of the plotted points have calculated displacements that

both fit, and do not fit the proposed design chart (especially the point on the center-right).

The variations in the displacement values for each of the plotted parametric study points

can be attributed to variations in the earthquake motions. Larger earthquake motions

produced larger displacements, whereas the method proposed by Dennehy and Gazetas

does not directly include any earthquake motion parameters (intensity, frequency or

duration). It is significant to note that many of the computed deformations that fall in

Zone I (deformations approximately less than 10 cm) would be considered unacceptable

by many port engineers for an operating or contingency level earthquake motion (Ferritto,

1997).

A comparison can also be made between the parametric study on the depth of sheet

pile embedment (Section 6.2) and the proposed chart by Dennehy and Gazetas. It was

noted from the parametric study that the depth of embedment had very little effect on the

performance of the bulkhead over the range of the modeled values, but the contour lines

constructed by Dennehy and Gazetas show a clear variation in performance over the

range of interest (EPI = 0.25 to 0.75).

6.8 DEVELOPMENT OF A NEW SEISMIC DESIGN METHOD

The main objective of this research was the development of a new seismic design

chart and recommendations for deformation-based design. In the development of a

seismic design chart, the results of the parametric study have been synthesized into

normalized parameters, where possible, to incorporate key variables into straightforward

design parameters. For example, aspects of the wall geometry and sheet pile stiffness

Page 118: Nason J. McCullough for the degree of Master of Science in

104

2

1.5

1

0.5

0

-0.5

0 0.5

Embedment participation index (EPI)

1

Figure 6.7: The Developed Seismic Design Chart (Gazetas et al.,1990), Including Data from the Parametric Study forModels with Improved Soils (solid stars)

have been combined as into a flexibility factor, (EI /(H +D)5), where El is the stiffness of

the wall, H is the height of the wall above the dredge line, and D is the depth of sheet pile

embedment. The computed displacements at the top of the wall (La) have been combined

with the flexibility factor and the buoyant unit weight of the soil adjacent to the wall (fl),

to yield a deformation factor as follows;

Page 119: Nason J. McCullough for the degree of Master of Science in

AX EI

(H +D)5 Y b

105

6.1

The results of the parametric study are presented in Figure 6.8. The contour lines

indicate various levels of earthquake intensity for backfill soils with blowcounts of 10

and 20 blows/30 cm. The effectiveness of soil improvement for minimizing bulkhead

deformations is clearly demonstrated by the design chart. It is also noted that an

incremental benefit of soil improvement beyond n values of approximately 2.0 decreases

considerably. In comparison, the n values as determined from the PHRI (1997)

recommendations for the parametric study sheet pile bulkhead geometries are

approximately 1.9 to 2.5.

There are fourteen case histories plotted on the chart, which are arranged according

to the blowcounts of the backfill soils. The case history data is summarized in Table 6.3.

It should be noted that seven case histories are closely predicted by the chart (within

10%), five case histories are significantly over-predicted 10%) and only two of the

case histories are significantly under-predicted (5_ 10%). These results indicate that the

design chart can be used as a conservative preliminary design chart or screening tool.

The results of the study indicate that it would be very difficult to limit the

deformations to 10 cm utilizing only densification methods of soil improvement for

moderate to high earthquake motions (Amax ?_ 0.3g). In cases such as these, it may be

necessary to consider other soil improvement techniques (e.g. grouting, soil mixing, etc.)

or structural improvements to strengthen the wall and/or anchor systems.

The chart has a relative error of approximately 30% for any of the calculated values,

this is due to the variations in earthquake motion that are not accounted for (e.g.

frequency) and other unknown wall behaviors. The value of 30% was determined from

the results of the validation case histories and from the variation of values on the design

chart that should be equal (e.g. normalized displacements for different wall heights).

Page 120: Nason J. McCullough for the degree of Master of Science in

0.014

0.012

0.010

1 I1 1 1

Case History Penetration Resistance of the(N1)60 (blows/30 cm) Unimproved Backfill Soil: _

0 510 (N1)60 (blows/30 cm)

0 15 - 1020 - - 20

1

t1

%

I L Ground Motion Intensity Factors( Anwor.dg./MSF )SIM

051Ill 0I3

01

0.008(

0.006

0.004

0.002

0.0000 1 2 3 4 5 6

Normalized Soil Improvement, n ( SW(H+D) )

Figure 6.8: Design Chart for Lateral Displacements

Table 6.3: Plotted Case History Validations

7 8

106

Earthquake (N d 60A,,*,fredge/MSF

(g)

Displacement(cm)

NormalizedDisplacement

AXE1/((H+D)511)1968 Tokachi-Oki 6 0.26 12 to 23 0.01201993 Kushiro-Oki 6 0.20 19 0.00201964 Niigita 10 0.14 100 0.00311983 Nihonkai-Chubu 10 0.10 110 to 160 0.00661968 Tokachi-Oki 15 0.23 16 to 57 0.00901968 Tokachi-Oki 15 0.23 12 to 19 0.00291968 Tokachi-Oki 15 0.12 30 0.00771973 Nemuro-Hanto 15 0.16 30 0.00311978 Miyagi-Ken-Oki 15 0.10 87 to 116 0.01291968 Tokachi-Oki 20 0.26 -60 0.00491983 Nihonkai -Chubu 20 0.09 -5 0.00061993 Kushiro-Oki 20 0.16 50 to 70 0.00441993 Guam 20 0.18 61 0.00191993 Kushiro-Oki 30 0.21 no damage (-5) 0.0004

Page 121: Nason J. McCullough for the degree of Master of Science in

107

6.9 RECOMMENDATION DESIGN PROCEDURES

The design chart is useful in estimating lateral deformations, but it is not

recommended for use in specifying bulkhead properties (stiffness, depth of embedment,

tie rod length, etc.). These should be specified by utilizing the current "standard of

practice" design methods, as outlined in Chapter 4. The recommended procedures for

utilizing the results of the parametric study to estimate the permanent displacement at the

top of sheet pile walls include;

1) determine the wall specifications (H, D, EI, and anchor length), either through a

new design or from an in-situ wall,

2) determine A,,,,@dredge with a site-specific seismic study or approximate with

empirical soil amplification factors to yield the peak ground surface acceleration

and the reduction factor (rd),

3) determine the magnitude scaling factor for the specified earthquake magnitude,

and divide into Amax@dredge to yield the intensity factor, and

4) based on the standard penetration resistance of the backfill soils, the width of the

ground treatment behind the sheet pile wall, and the ground motion intensity

factor, enter Figure 6.8 and obtain the deformation factor. From this, the

deformation at the top of the wall (dX) can be estimated.

Using these recommended procedures it is possible for an engineer to determine the

deformations of a new or in-situ wall. It is also possible to estimate the zone of soil

improvement necessary to limit deformation to an acceptable limit. An example design

problem is provided in the following section.

Page 122: Nason J. McCullough for the degree of Master of Science in

108

6.10 EXAMPLE DESIGN PROBLEM

This section provides the calculations involved in determining the extent of soil

improvement necessary to limit wall deformations for a sheet pile bulkhead, using the

developed design method from Section 6.8. The example bulkhead is shown in Figure

6.9, and was designed using the Mononobe-Okabe Method. The example problem was

taken from Appendix C in The Seismic Design of Waterfront Retaining Structures by

Ebeling and Morrison (1993). The example wall was designed using the standard of

practice Mononobe-Okabe design methods with a factor of safety of 1.5 applied to the

passive earth pressure coefficient. This was a minimum design, no consideration was

given for scour, dredging, corrosion or other design considerations.

This example illustrates the use of Operating and Contingency Level earthquake

motions that are likely for design in the San Francisco-Oakland Bay area. The Operating

Level earthquake motion (OLE) is estimated as a magnitude 6.5 earthquake with peak

ground accelerations of 0.25 g. The Contingency Level earthquake motion (CLE) is

estimated as a magnitude 7.5 earthquake with peak ground accelerations of 0.45 g. The

following design example uses both levels of earthquake motion to estimate the expected

deformations for various extents of densification improved soil.

improved soil= 19 kN/m3

(Ndso 15 blows/30 cm9 kN/m3

7v1440'114k

sheet pile wall (PZ27)

foundation soil

backfill soil

Figure 6.9: Anchored Sheet Pile Bulkhead Design Problem

Page 123: Nason J. McCullough for the degree of Master of Science in

The sheet pile wall values for a PZ27 section are:

E = 2x1011 PaI= 2.51x104 m4/m

The wall height (H) is 9 m, and the wall depth (D) is 6 m.

The soil buoyant unit weight (A) directly adjacent to the wall is 9000 N/m3.

Substituting into the normalized lateral displacement factor (Equation 6.1) yields:

AX EI AX (2 x 10" Pa). (2.51x10-4 m4/m) AX

(9 m + 6 m)5 (9000 N/m3) 136(H + D)5 7b

109

The ground motion intensity factor is determined for the earthquake motions using

Table 2.2 to determine the MSF and Figure 2.1 to determine rd, using the following

equation:

Amax@ dredge Amax@ surface rd

MSF MSF

The magnitude scaling factor values for both earthquake motions are tabulated in

Table 6.4.

Table 6.4: Magnitude Scaling Factors for Design Example

A max@surface rd (at 9 m) A mar@d redge MSF A max*Iredge

MSF

OLE 0.25 0.92 0.23 1.63 0.14

CLE 0.45 0.92 0.41 1.00 0.41

Page 124: Nason J. McCullough for the degree of Master of Science in

110

The normalized lateral displacement factor is then determined as a function of soil

improvement (n) using Figure 6.8 for blowcounts of 15 and a maximum acceleration of

0.2 g, as provided in Table 6.5.

Table 6.5: Lateral Displacements for Design Example

Normalized SoilImprovement (n)

Normalized LateralDisplacement (b)

Lateral Displacement(8x 136) (m)

LateralDisplacement (cm)

OLE

0 0.0050 0.68 68

1 0.0018 0.24 24

2 0.0007 0.10 10

3 0.0005 0.07 7

8 0.0004 0.05 5

CLE

0 0.0130 1.77 177

1 0.0041 0.56 56

2 0.0021 0.29 29

3 0.0018 0.24 24

8 0.0017 0.23 23

By plotting the values tabulated above, it is possible to develop a figure relating the

extent of soil improvement to lateral displacements (Figure 6.10). It is interesting to note

that for this example, it is impossible to limit the deformations to 5 cm for an OLE

motion and 23 cm for a CLE motion, using only soil improvement techniques. Other

forms of earthquake mitigation are necessary to limit earthquake-induced deformations to

smaller values. The design tie rod length was 28 m, which is at the break in the curve in

Figure 6.10. Improvement beyond this point has little effect on the earthquake-induced

lateral displacements.

It should also be noted that the example problem varied from the developed design

method in that the tie rod and water elevations were not equal.

Page 125: Nason J. McCullough for the degree of Master of Science in

200

180

160

E140

E 120

(.3as 100c.

80

11'co 60ea

40

20

0

OLE motion

CLE motion

0 20 40 60 80

Soil Improvement (m)

Figure 6.10: Lateral Displacements for the Design Problem

100 120

111

Page 126: Nason J. McCullough for the degree of Master of Science in

112

7 SUMMARY AND CONCLUSIONS

There have been numerous recent earthquakes, which have resulted in severe damage

to many international ports. The damage has been due largely to the liquefaction of

saturated, sandy soils. Waterfront structures, particularly anchored sheet pile bulkheads

are very susceptible to liquefaction-induced damages. Due to the recent damage caused

by liquefiable soils, many port authorities are instigating soil improvement programs to

limit the generation of excess pore pressures during design level earthquakes. The case

histories of liquefaction-damaged sheet pile bulkheads have highlighted the need for a

simplified, performance-based design method.

The "standard of practice" seismic design procedures for sheet pile wall involves

using pseudo-static, limit-equilibrium methods, developed for rigid retaining walls. There

are several recent modifications to the "standard of practice" design methods, but are very

limited in their applicability to design. The method proposed by Dennehy (1985) and

Gazetas et al. (1990) is performance-based, but is limited to non-liquefiable soils. The

other recent additions to sheet pile design are limit-equilibrium based.

A study that incorporates field performance and numerical modeling has been

conducted using a validated numerical model, utilizing several empirical relationships

developed for geotechnical earthquake parameters. The study involved determining the

effects of varying several design parameters, including the depth of sheet pile

embedment, tie rod length, sheet pile stiffness, penetration resistance of the backfill,

extent of soil improvement, and ground motion characteristics. Charts were developed

depicting the effect of varying these various parameters. A final design chart was

developed utilizing the results from the study, and includes data from several case

histories. There is some inherent scatter in the data due to the variability of the earthquake

motions, but the design chart provides a reasonable method for estimating lateral

deformations of sheet pile bulkheads with or without soil improvement. This design chart

is applicable for the preliminary design of new bulkheads and as a screening tool for

existing bulkheads.

Page 127: Nason J. McCullough for the degree of Master of Science in

113

This investigation has highlighted the following pertinent aspects of the seismic

design and performance of anchored sheet pile walls;

from a practical perspective, the "failure" of an anchored sheet pile bulkhead can

be considered as corresponding to lateral deformations in the range of 10 cm to 20

cm,

the factors of safety computed with "standard of practice" design methods are not

adequately correlated with wall deformations to facilitate estimates of seismically-

induced lateral displacements,

in order to estimate the deformations of sheet pile walls, the intensity and duration

of the earthquake, and potential pore pressure generation in the backfill and

foundation soils must be evaluated,

in light of the flexible nature of anchored sheet pile walls, lateral deformations

should be anticipated even in competent, non-liquefiable soils subjected to

moderate- to high-intensity ground motions. In relationship to the Navy guidelines

proposed by Ferritto (1997) for sheet pile walls, it may be impossible to design a

sheet pile bulkhead to stay within the recommended deformations utilizing only

soil densification techniques, and

a simplified method of estimating seismically induced lateral deformations for

anchored sheet pile bulkheads in unimproved or improved soils has been

proposed.

Several recommendations for future work are also noted, including;

studies to examine the effects of different tie rod elevations and other structural

properties such as the anchor and tie rod stiffness, and different anchor

configurations (such as single pile, deadman, or battered piles),

studies that examine the effect of pore pressured dissipation, as well as generation,

Page 128: Nason J. McCullough for the degree of Master of Science in

114

studies that examine other bulkhead variables, such as liquefiable foundation

soils, cohesive foundation soils,

studies on the specific effects of various soil improvement strategies (e.g. vibro-

compaction, soil mixing, deep dynamic compaction), and

utilize scale-model testing (i.e. centrifuge) for validation with case history and

numerical model studies.

Page 129: Nason J. McCullough for the degree of Master of Science in

115

BIBLIOGRAPHY

Arango, I. 1996. "Magnitude Scaling Factors for Soil Liquefaction Evaluations." Journalof Geotechnical Engineering. American Society of Civil Engineers. Vol 122, No 11.pp 929-936.

Arbed. 1991. Practical Design of Sheet Pile Bulkheads. Trade-Arbed, Inc., Luxembourg.96 p.

Baez, J.I. and Martin, G.R. 1992. "Quantitative Evaluation of Stone Column Techniquesfor Earthquake Liquefaction Mitigation." Proceedings of the Tenth World Conferenceon Earthquake Engineering. A.A. Balkema, Rotterdam. pp 1477-1483.

Boulanger, R.W. and Seed, R.B. 1995. "Liquefaction of Sand Under BidirectionalMonotonic and Cyclic Loading." Journal of Geotechnical Engineering. AmericanSociety of Civil Engineers. Vol 121, No 12. pp 870-878.

CKPHB (City of Kobe Port and Harbor Bureau). 1997. Reconstruction and RestorationWorks at Port of Kobe. Port and Harbor Bureau. Kobe, Japan.

Dennehy, K.T. 1985. Seismic Vulnerability, Analysis and Design of Anchored Bulkheads.Ph.D. Dissertation. Department of Civil Engineering, Rensselaer PolytechnicInstitute, Troy, New York. 215 p.

Dickenson, S.D., Ferritto, J., Kaldveer, P., and Soydemir, C. 1998. "Chapter 4, Seismicand Geologic Hazards." Seismic Guidelines for Ports. ASCE-TCLEE Monograph No.12. S.D. Werner (ed.). 376 p.

Dickenson, S.D. and Yang, D.S. "Seismically-Induced Deformations of CaissonRetaining Walls in Improved Soils." Proceedings from the 1998 ASCE GeotechnicalEarthquake Engineering and Soil Dynamics Conference. ASCE. Seattle, WA. August3-6. (in press).

Ebeling, R.M. and Morrison, E.E. 1993. The Seismic Design of Waterfront RetainingStructures. U.S. Naval Civil Engineering Laboratory, Port Hueneme, CA. TechnicalReport ITL-92-11. NCEL TR-939. 329 p.

Egan, J.A., Hayden, R.F., Scheibel, L.L., Otus, M., and Serventi, G.M. 1992. "SeismicRepair at Seventh Street Marine Terminal." Proceedings from GT Division/ASCEGrouting, Soil Improvement and Geosynthetics. ASCE. New Orleans, LA. February25-28. pp. 867-878.

Ferritto, J. 1997. Seismic Design Criteria for Soil Liquefaction. Naval Facilities Engr. ServiceCenter. Technical Report TR-2077-SHR. Port Hueneme, CA.58 p.

Page 130: Nason J. McCullough for the degree of Master of Science in

116

Gazetas, G., Dakoulas, P., and Dennehy, K. 1990. "Empirical Seismic Design Method forWaterfront Anchored Sheetpile Walls." Proceedings of the ASCE SpecialtyConference on Design and Performance of Earth Retaining Structures. ASCEGeotechnical Special Publication. No 25. pp 232-250.

Hayden, R.F., and Boulanger, R.W. 1995. "Aspects of Compaction Grouting ofLiquefiable Soil." Journal of Geotechnical Engineering. American Society of CivilEngineers. Vol 121, No 12. pp 844-855.

Iai, S., Tsuchida, H. and Finn, W.D Liam. 1985. "An Effective Stress Analysis ofLiquefaction at Ishinomaki Port during the 1978 Miyagi-Ken-Oki Earthquake."Report of the Port and Harbour Research Institute. Vol 24, No 2. pp 3-83.

Iai, S., and Kameoka, T. 1993 (March). "Finite Element Analysis of Earthquake InducedDamage to Anchored Sheet Pile Quay Walls." Soils and Foundations. JapaneseSociety of Soil Mechanics and Foundation Engineering. Vol 33, No 1. pp 71-91.

Iai, S., Matsunaga, Y., Morita, T. , Sakurai , H., Oishi , H., Ogura , H., Ando , Y., Tanaka,Y., and Kato, M. 1994. "Effects of Remedial Measures against Liquefaction at 1993Kushiro-Oki Earthquake." Proceedings of the Fifth US-Japan Workshop onEarthquake Resistant Design of Lifeline Facilities and Countermeasures Against SiteLiquefaction. T.D. O'Rourke and M. Hamada (eds.). National Center for EarthquakeEngineering Research. NCEER-94-0026. Statue University of New York, Buffalo,NY. pp 135-152.

Iai, S., Morita, T., Kameoka, T., Matsunaga, Y., and Abiko, K. 1995. "Response of aDense Sand Deposit During 1993 Kushiro-Oki Earthquake." Soils and Foundations.Japanese Society of Soil Mechanics and Foundation Engineering. Vol 35, No 1. pp115-131.

Iai, S. and Ichii, K. 1997 (April). Personal Communication. Port and Harbor ResearchInstitute, Japan.

Idriss, I.M., and Sun, J.I. 1992. User's Manual for SHAKE91. Center for GeotechnicalModeling. Department of Civil and Environmental Engineering, University ofCalifornia, Davis.

Imai, T., Fumoto, H. and Yokota, K. 1975. "Relations of Mechanical Properties of Soilsto P and S wave Velocities in Japan." Proceedings of the Fourth Japan EarthquakeEngineering Symposium. pp 89-96. (in Japanese).

Ishihara, K., and Yohimine, M. 1992. "Evaluation of Settlements in Sand DepositsFollowing Liquefaction During Earthquakes." Soils and Foundations. JapaneseSociety of Soil Mechanics and Foundation Engineering. Vol 32, No 1. pp 173-188.

Page 131: Nason J. McCullough for the degree of Master of Science in

117

Itasca Consulting Group. 1995. User's Manual for FLAC, v. 3.3. Itasca. Minneapolis,Minnesota.

Kitajima, S., and Uwabe, T. 1979. "Analysis on Seismic Damage in Anchored Sheet-piling Bulkheads." Report of the Port and Harbour Research Institute. Vol 18, No 1.pp 67-130. (in Japanese).

Kramer, S.L. 1996. Geotechnical Earthquake Engineering. Prentice Hall, Inc. UpperSaddle River, NJ. 653 p.

Lee, M.K.W., and Finn, W.D.L. 1991. DESRA-2C: Dynamic Effective Stress ResponseAnalysis of Soil Deposits with Energy Transmitting Boundary Including Assessmentof Liquefaction Potential. University of British Columbia, Faculty of AppliedScience. 34 p.

Li, X.S., Wang, Z.L., and Shen, C.K. 1992. SUMDES: A Nonlinear Procedure forResponse Analysis of Horizontally-Layered Sites Subjected to Multi-DirectionalEarthquake Shaking. Department of Civil Engineering, University of California,Davis. 81 p.

Liao, S.S., and Whitman, R.V. 1986. "Overburden Correction Factors for Sand." Journalof the Geotechnical Engineering Division. American Society of Civil Engineers. Vol112, No 3. March. pp 373-377.

Marcuson, W.F., and Hynes, M.E. 1990. "Stability of Slopes and Embankments DuringEarthquakes." Proceedings of the ASCE/Penn. DOT Geotechnical Seminar. Hershey,Penn. April 10-11.

Martin, G.R., Finn, W.D.L., and Seed, H.B. 1975. "Fundamentals of Liquefaction UnderCyclic Loading." Journal of the Geotechnical Engineering Division. AmericanSociety of Civil Engineers. Vol 101, No GT5. May. pp 423-438.

Mitchell, J.K., and Tseng, D. 1990. "Assessment of Liquefaction Potential by ConePenetration Resistance." H Bolton Seed Memorial Proceedings. Editor J. MichaelDuncan. BiTech Publishers, Ltd, Vancouver, BC. Vol 2. pp 335-350.

Mononobe, N., and Matsuo, H. 1929. "On the Determination of Earth Pressures DuringEarthquakes." Proceedings of the 9th World Engineering Conference.

Neelakantan, G., Budhu, M., and Richards, R. Jr. 1992. "Balanced Seismic Design ofAnchored Retaining Walls." Journal of Geotechnical Engineering. ASCE. Vol 118,No 6. pp 873-888.

Newmark, N. 1965. "Effects of Earthquakes on Dams and Embankments." Geotechnique.Vol 15, No 2. pp 139-160.

Page 132: Nason J. McCullough for the degree of Master of Science in

118

Noda, S. Uwabe, T., and Chiba, T. 1975. "Relation between Seismic Coefficient andGround Acceleration for Gravity Quaywall." Report of the Port and HarbourResearch Institute. Vol 14, No 4. pp 67-111.

Nozu, A., Uwabe, T., Sato, Y. and Shinozawa, T. 1997. "Relation Between SeismicCoefficient and Peak Ground Acceleration Estimated from Attenuation Relation."Technical Note of the Port and Harbour Research Institute. Ministry of Transport,Japan. (in print). (in Japanese).

Okabe, S. 1926. "General Theory of Earth Pressures." Journal of Japan Society of CivilEngineering. Vol 12, No 1.

PHRI (Port and Harbour Research Institute). 1993. Damage to Port Structures by the1993 Kushiro-Oki Earthquake. Technical Report of the Port and Harbour ResearchInstitute. Ministry of Transport, Japan. December. No 766. 454 pp. (in Japanese).

PHRI (Port and Harbour Research Institute). 1996. The Great Hanshin-Awaji EarthquakeResearch Council Report. Port and Harbour Research Institute. Ministry of Transportand the Third District Port Construction Bureau. Yokosuka, Japan. (in Japanese).

PHRI (Port and Harbour Research Institute). 1997. Handbook on LiquefactionRemediation of Reclaimed Land. A.A. Balkema. Rotterdam, Netherlands. 312 p.

Roth, W.H., Scott, R.F., and Cundall, P.A. 1986. "Nonlinear Dynamics Analysis of aCentrifuge Model Embankment." Proceedings of the 3rd U.S. National Conference onEarthquake Engineering. Charleston, South Carolina.

Roth, W.H., and Inel, S. 1993. "An Engineering Approach to the Analysis of VELACSCentrifuge Tests." Verifications of Numerical Procedures for the Analysis of SoilLiquefaction Problems. Arulanandan and Scott (eds.). A.A. Balkema, Rotterdam. pp1209-1229.

Rowe, P.W. 1952. "Anchored Sheet-Pile Walls." Proceedings of the Institution of CivilEngineers. Paper No 5788. Vol 1, No 1. pp 27-70.

Schnabel, P.B., Lysmer, J., and Seed, H.B. 1972. SHAKE: A Computer Program forEarthquake Response of Horizontally Layered Sites. Earthquake EngineeringResearch Center. Report No. EERC/72-12. University of California, Berkeley. 88 p.

Seed, H.B., and Whitman, R. 1970. "Design of Earth Retaining Structures for DynamicLoads." ASCE Specialty Conference on Lateral Stresses in the Ground and Design ofEarth Retaining Structures. pp 103-147.

Seed, H.B., Martin, P.P, and Lysmer, J. 1976 (April). "Pore-Pressure Changes DuringSoil Liquefaction." Journal of the Geotechnical Engineering Division. AmericanSociety of Civil Engineers. Vol 102, No GT4. pp 323-345.

Page 133: Nason J. McCullough for the degree of Master of Science in

119

Seed, H.B. 1979. "Soil Liquefaction and Cyclic Mobility Evaluation for Level GroundDuring Earthquakes." Journal of the Geotechnical Engineering Division. AmericanSociety of Civil Engineers. Vol 105, No GT2. pp 201-255.

Seed, H.B., and Idriss, I.M. 1982. Ground Motions and Soil Liquefaction DuringEarthquakes. Earthquake Engineering Research Institute Monograph Series. Oakland,California.134 p.

Seed, H.B., Tokimatsu, K., Harder, L.F. and Chung, R.M. 1985. "Influence of SPTProcedures in Soil Liquefaction Resistance Evaluations." Journal of the GeotechnicalEngineering. ASCE, Vol 111, No 12. Paper No 20215. pp 1425-1445.

Seed, H.B., and De Alba, P. 1986. "Use of SPT and CPT Tests for Evaluating theLiquefaction Resistance of Soils." Proceedings Insitu 1986. ASCE SpecialPublication.

Seed, R.B. and Harder, L.F. 1990. "SPT-Based Analysis of Cyclic Pore PressureGeneration and Undrained Residual Strength." H. Bolton Seed MemorialProceedings. Editor J. Michael Duncan. Vol 2. pp 351-376.

Stark, T.D., and Mesri, G. 1992. "Undrained Shear Strength of Liquefied Sands forStability Analysis." Journal of Geotechnical Engineering. ASCE. Vol 118, No 11. pp1727-1747.

Stark, T.D., and Olson, S.M. 1995. "Liquefaction Resistance Using CPT and Field CaseHistories." Journal of the Geotechnical Engineering. ASCE, Vol 121, No 12. pp 856-869.

Steedman, R.S., and Zeng, X. 1990. "The Seismic Response of Waterfront RetainingWalls." Proceedings of the ASCE Specialty Conference on Design and Performanceof Earth Retaining Structures. ASCE Geotechnical Special Publication. No 25. pp872-886.

Swan, S.W., and Harris, S.K. 1993. The Island of Guam Earthquake of August 8, 1993.National Center for Earthquake Engineering Research. Technical Report NCEER-93-0017. September 30. State University of New York, Buffalo, NY.

Terzaghi, K. 1954 . "Ancored Bulkheads." Transaction of the American Society of CivilEngineers. Vol 119. pp 1243-1324

Terzaghi, K., Peck, R.B., and Mesri, G. 1996. Soil Mechanics in Engineering Practice.Third edition. John Wiley and Sons, Inc. New York, NY. 549 p.

Page 134: Nason J. McCullough for the degree of Master of Science in

120

Towhata, I., and Islam, Md. S. 1987 (December). "Prediction of Lateral Displacement ofAnchored Bulkheads Induced by Seismic Liquefaction." Soils and Foundations.Japanese Society of Soil Mechanics and Foundation Engineering. Vol 27, No 4. pp137-147.

Towhata, I., Ghalandarzadeh, A., Sundarraj, K.P., and Vargas-Monge, W. 1996."Dynamic Failures of Subsoils Observed in Waterfront Areas." Special Issue of Soilsand Foundations. Japanese Geotechnical Society. pp 149-160.

Ueda, S., Inatomi, T., Uwabe, T., Iai, S., Kazama, M., Matsunaga, Y., Hujimoto, T.,Kikuchi, Y., Miyai, S., Sekiguchi, S., and Hujimoto, Y. 1993 (December). "Damageto Port Structures by the 1993 Kushiro-oki Earthquake." Technical Note of the Portand Harbour Research Institute. Ministry of Transport, Japan. No 766. (in Japanese).

United States Steel. 1975. US Steel Sheet Piling Design Manual.

USACE (US Army Corps of Engineers). 1984. Geotechnical Investigations. EM 1110 -1-1804. Department of the Army.

USACE (US Army Corps of Engineers). 1994. Design of Sheet Pile Walls. EM 1110 -2-2504. Department of the Army.

Vandani, S., Pyke, R., and Siriprusanen, U. 1994. "Liquefaction of Calcareous Sands andLateral Spreading Experienced in Guam as a Result of the 1993 Guam Earthquake."Proceedings of the Fifth US-Japan Workshop on Earthquake Resistant Design ofLifeline Facilities and Countermeasures Against Site Liquefaction. T.D. O'Rourkeand M. Hamada (eds.). National Center for Earthquake Engineering Research.NCEER-94-0026. Statue University of New York, Buffalo, NY. pp 117-134.

Werner, S.D., and Hung, S.J. 1982. Seismic Response of Port and Harbor Facilities.National Science Foundation. R-8122-5395. Agbabian Associates. El Segundo,California. 348 p.

Werner, S.D. 1990. "Seismic Risks to Port and Harbor Facilities." Proceedings of thePOLA Seismic Workshop on Seismic Engineering. R.D. Wittkop and G.R. Martin(eds.). Port of Los Angeles. San Pedro, CA. March 21-23.

Werner, S.D. (ed.) 1998. Seismic Guidelines for Ports. ASCLE-TCLEE Monograph No12. ASCLE-TCLEE Ports and Harbors Committee. 376 p.

Westergaard, H. 1931. "Water Pressure on Dams During Earthquakes." Transactions ofASCE. Paper No 1835. pp 418-433.

Page 135: Nason J. McCullough for the degree of Master of Science in

121

Yasuda, S, Ishihara, K., Harada, K., and Shinkawa, N. 1996. "Effect of Soil Improvementon Ground Subsidence due to Liquefaction." Special Issue of Soils and Foundationson Geotechnical Aspects of the January 17, 1995 Hyogoken-Nanbu Earthquake.Japanese Geotechnical Society. January. pp 99-107.

Yoshida, I. et al. 1988. "Empirical Formulas of SPT Blow Counts for Gravely Soils."First International Symposium on Penetration Testing. Institution of Civil Engineers,Vol 1. pp 381-387.

Page 136: Nason J. McCullough for the degree of Master of Science in

122

APPENDICES

Page 137: Nason J. McCullough for the degree of Master of Science in

123

APPENDIX A(EARTHQUAKE MOTIONS USED IN THE PARAMETRIC STUDY)

Page 138: Nason J. McCullough for the degree of Master of Science in

4

E 2

0

O 2

4

E 2

o 0

0O 2

4

0

-2

1984 Morgan Hill EQ- Gilroy #4 (Mw.0, Am ax.222g)

1989 Loma Prieta EQ - Captola Fire Station (Mw.9, Amax=-0.399)

111111111111

iTIMIP,111111111,011111.1111

11E111989 Loma Prieta EQ - Salinas (Mw.9, Amax=-0.112)

1992 Landers EQ - Joshua Tree Fire Station (Mw=7.4, Am ax).274)

1985 Michoacan Mexico EQ (Mw=8.1, Scaled to Amax.350g)

RIrip

I Ig rrirlT prrr0

INN'S 2

au I

124

0 10 20 30

Time (seconds)

40 50 60

Page 139: Nason J. McCullough for the degree of Master of Science in

125

APPENDIX B(TABLE OF PARAMETRIC STUDY RESULTS)

Page 140: Nason J. McCullough for the degree of Master of Science in

126

Ex1101.6;

Ti

Variations

7.5

SE

is

0.10

g

1Z

.2's 2-2 E

12 TI)ocr .1 En.6.c e ga Z a

Depth of Embedment

''36

(a

0.52

7-r,_'c'o-

I-a

0.26

g(7)

..:'.E'

0

E

'5co

72.8

.:-I" .a .}!

i I. .-E a. ..1

2 Ed, aca

E. c =tu 3 "a

.. o f,E. E0 .2 oct I- 2

E'Eg

&' 5c a0...... - cr0 z r5 4. 50. :-. tu

- -t s' - 1.3) 0 E 2E.., E -g e ,?, T, eTs. . o. ge ' ° e .,3 0'ci 3. O« m:1:. is 5 "ot., 02 gTe

..t.,.. .,.., 2 .R. Er, a . r, . .x 1- Z a) 2 ..?..

-6- -"fp °it .- 20z u.z a I'

p., cil lxLL 12 0 X

c 'SI r." of.2' ce 21 E El2 ., r, x x0 i= 0 O. -1

3.45Morgan all 0.30 5.8 15.9 30 10.4 6.0 0.29 0.37 1.4 2.2 2.4 2.6 0.0005

7.5 0.10 Morgan all 0.30 0.52 0.26 72.8 2.9 15.9 3.45 30 10.4 6.0 0.32 0.40 1.4 2.2 2.4 2.6 0.0017

7.5 0.10 Morgan all 0.30 0.52 0.26 72.8 1.5 15.9 3.45 30 10.4 6.0 0.31 0.38 1.3 2.2 2.4 2.6 0.0036

7.5 0.10 Mexico all 0.30 0.66 0.94 72.8 1.5 15.9 3.45 30 10.4 8.1 0.64 0.82 1.3 2.2 2.3 2.6 0.0074

7.5 0.10 Mexico all 0.30 0.66 0.94 72.8 2.9 15.9 3.45 30 10.4 8.1 0.61 0.80 -2.3 2.2 2.5 2.6 0.0033

7.5

Variations

0.10

is

Mexico

Stiffness

Morgan

all

all

0.30

0.30

0.66

0.59

0.94

0.29

72.8

72.8

5.8

2.9

15.9

15.9

3.45

1.73

30

30

10.4

10.4

8.1

6.0

0.59

0.34

0.78

0.42

-2.1

1.2

2.2

2.2

2.5

2.3

2.6

2.6

0.0009

0.00097.5 0.10

7.5 0.10 Morgan all 0.30 0.59 0.29 72.8 2.9 15.9 3.45 30 10.4 6.0 0.31 0.28 1.2 2.2 2.3 2.6 0.0017

7.5 0.10 Morgan all 0.30 0.59 0.29 72.8 2.9 15.9 6.90 30 10.4 6.0 0.30 0.37 -2.0 2.2 2.4 2.6 0.0033

7.5 0.10 Mexico all 0.30 0.66 0.94 72.8 2.9 15.9 1,73 30 10.4 8.1 0.63 0.85 -1.5 2.2 2.4 2.6 0.0017

7.5 0.10 Mexico all 0.30 0.66 0.94 72.8 2.9 15.9 3.45 30 10.4 8.1 0.62 0.67 -2.2 2.2 2.3 2.6 0.0034

7.5 0.10 Mexico all 0.30 0.66 0.94 72.8 2.9 15.9 6.90 30 10.4 8.1 0.62 0.75 -3.0 2.2 2.6 2.6 0.0067

15.0 0.10 Morgan all 0.30 0.44 0.22 142.0 5.4 30.0 19.12 30 10.4 6.0 0.43 0.54 8.4 13.0 7.9 8.7 0.0004

15.0 0.10 Morgan all 0.30 0.44 0.22 142.0 5.4 30.0 76.48 30 10.4 6.0 0.36 0.46 -9.4 13.0 8.2 8.7 0.0015

15.0 0.10 Mexico all 0.30 0.48 0.68 142.0 5.4 30.0 19.12 30 10.4 8.1 0.61 0.81 9.1 13.0 8.1 8.7 0.0006

7.5 0.10 Morgan 1.9 0.30 0.59 0.29 20.2 2.9 15.9 1.73 10 10.4 6.0 0.35 0.40 1.0 2.2 1.9 2.6 0.0010

7.5 0.10 Morgan 1.9 0.30 0.59 0.29 20.2 2.9 15.9 3.45 10 10.4 6.0 0.34 0.37 1.1 2.2 1.9 2.6 0.0018

7.5

Variations

0.10

is

Morgan

Tie Rod Length

Morgan

1 9

all

0.30

0.10

0.59

0.20

0.29

0.10

20.2

72.8

2.9

2.9

15.9

8.0

6.90

3.45

10

30

10,4

10.4

6.0

6.0

0.33

0.01

0.35

0.01

-1.7

0.8

2.2

2.2

2.0

1.7

2.6

2.6

0.0036

0.00017.5 0.10

7.5 0.10 Morgan all 0.30 0.59 0.29 72.8 2.9 8.0 3.45 30 10.4 6.0 0.48 0.44 -1.1 2.2 1.1 2.6 0.0026

7.5 0.10 Morgan all 0.30 0.59 0.29 72.8 2.9 15.9 3.45 30 10.4 6.0 0.22 0.28 1.2 2.2 2.3 2.6 0.0012

7.5 0.10 Morgan all 0.30 0_59 029 72.8 2.9 31.8 3.45 30 10.4 6.0 0.07 0.22 1.4 2.2 2.6 2.6 0.0004

7.5 0.10 Joshua all 027 0.52 0.50 72.8 2.9 8.0 3.45 30 10.4 7.4 0.69 0.60 1.4 2.2 2.7 2.6 0.0038

7.5 0.10 Joshua all 0.27 0.52 0.50 72.8 2.9 15.9 3.45 30 10.4 7.4 0.31 0.37 1.3 2.2 2.4 2.6 0.0017

7.5 0.10 Mexico all 0.30 0.66 0.94 72.8 2.9 8.0 3.45 30 10.4 8.1 1.19 1.00 -1.5 2.2 2.5 2.6 0.0065

7.5 0.10 Mexico all 0.30 0.66 0.94 72.8 2.9 15.9 3.45 30 10.4 8.1 0.50 0.67 -2.2 2.2 2.3 2.6 0.0028

7.5

Variations

0.10

in

Mexico

the Penet

all

ation

0.0

0.30

Resistance

0.30

0.66

of

0.59

0.94

the Backfill

0.29

72.8

0.0

2.9

2.9

31.8

15.9

3.45

3.45

30

10

10.4

7.2

8.1

6.0

0.13

3.50

0.51

1.79

-2.5

3.2

2.2

2.2

2.4

1.6

2.6

2.6

0.0007

0.02767.5 0.10 Morgan

7.5 0.10 Morgan 0.0 0.30 0.59 0.29 0.0 2.9 15.9 3.45 15 8.4 6.0 1.51 0.94 -1.5 2.2 1.8 2.6 0.0102

7.5 0.10 Morgan 0.0 0.30 0.59 0.29 0.0 2.9 15.9 3.45 20 9.2 6.0 0.82 0.69 -2.1 2.2 1.8 2.6 0.0051

7.5 0.10 Morgan 0.0 0.30 0.59 0.29 20.2 2.9 15.9 3.45 20 10.4 6.0 0.31 0.35 1.1 2.2 2.0 2.6 0.0017

7.5 0.10 Morgan 0.0 0.30 0.59 0.29 0.0 2.9 15.9 3.45 25 9.8 6.0 0.50 0.51 -1.9 2.2 1.9 2.6 0.0029

7.5 0.10 Joshua 0.0 0.27 0.52 0.50 0.0 2.9 15.9 3.45 15 8.4 7.4 2.86 1.55 1.9 2.2 1.7 2.6 0.0194

7.5 0.10 Joshua 0.0 0.27 0.52 0.50 0.0 2.9 15.9 3.45 25 9.8 7.4 0.79 0.72 -2.1 2.2 1.9 2.6 0.0045

7.5 0.10 Mexico 0.0 0.30 0.66 0.94 0.0 2.9 las 3.45 20 9.2 8.1 2.57 1.60 -1.5 2.2 1.9 2.6 0.0159

7.5

Variations

15.0

0.10

in

0.10

Mexico

the Zone

Morgan

0.0

of Soil

1.9

0.30

Improvement

0.10

0.66

0.15

0.94

0.07

0.0

38.7

2.9

5.4

15.9

30.0

3.45

38.24

25

10

9.8

10.4

8.1

6.0

1.56

0.05

1.19

0.06

-2.2

4.7

2.2

13.0

1.8

5.4

2.6

8.7

0.0090

0.0001

7.5 0.10 Morgan 1.9 0.10 0.20 0.10 20.2 2.9 15.9 3.45 10 10.4 6.0 0.04 0.05 0.7 2.2 1.6 2.6 0.0002

15.0 0.10 Capitola 1.9 0.10 0.14 0.11 38.7 5.4 30.0 38.24 10 10.4 6.9 0.05 0.06 4.5 13.0 5.3 8.7 0.0001

15.0 0.10 Mexico 1.9 0.05 0.08 0.11 38.7 5.4 30.0 38.24 10 10.4 8.1 0.02 0.02 3.9 13.0 4.8 8.7 0.0000

7.5 0.10 Salinas 0.8 0.11 0.16 0.12 8.0 2.9 15.9 3.45 10 10.4 6.9 0.36 0.28 -1.1 2.2 1.4 2.6 0.0020

7.5 0.10 Salinas 1.5 0.11 0.16 0.12 15.9 2.9 15.9 3.45 10 10.4 6.9 0.13 0.13 0.8 2.2 1.5 2.6 0.0007

7.5 0.10 Salinas 1.9 0.11 0.16 0.12 20.2 2.9 15.9 3.45 10 10.4 6.9 0.08 0.09 0.8 2.2 1.6 2.6 0.0004

7.5 0.16 Salinas 2.4 0.11 0.16 0.12 26.3 3.7 22.0 7.43 10 10.4 6.9 0.04 0.06 1.0 2.9 2.0 3.6 0.0004

7.5 0.10 Salinas 2.9 0.11 0.16 0.12 30.0 2.9 15.9 3.45 10 10.4 6.9 0.05 0.06 0.9 2.2 1.9 2.6 0.0003

Page 141: Nason J. McCullough for the degree of Master of Science in

127

1x:a'cra

T.,x7,1

7.5 0.10

ci

-132.ra=cr,ztaW

Salinas

Z

all

17ii

1<

0.11

a

0.16

1:., 1

u..co

-a,

"Ietg<

0.12

N

72.8

O

2.9

T<.eth

3-oo

C f.

I'1-

15.9 3.45

d

30 10.4

W

6.9

-ss'sa,E1-4 ..i.a.E., m

I' 4E,ti 0i ,' a= 1-

0.05

-ist -o' EE -!au 70

,7%.

. 2 .

3"a8 E.,,, 25 15'S m

0.06 1.0 2.2 2.0 2.6 0.0002

15.0 0.10 Salinas 1.9 0.11 0.18 0.14 38.7 5.4 30.0 38.24 10 10.4 6.9 0.07 0.09 5.3 13.0 5.7 8.7 0.0002

15.0 0.16 Salinas 2.2 0.11 0.18 0.14 48.2 6.9 39.5 92.45 10 10.4 6.9 0.04 0.06 6.3 17.9 6.4 11.9 0.0001

15.0 0.10 Morgan 1.9 0.22 0.33 0.16 38.7 5.4 30.0 38.24 10 10.4 6.0 0.25 0.28 6.8 13.0 6.4 8.7 0.0005

15.0 0.16 Morgan 2.2 0.22 0.33 0.16 48.2 6.9 39.5 92.45 10 10.4 6.0 0.17 0.22 8.1 17.9 7.5 11.9 0.0006

15.0 0.10 Capitols 1.9 0.20 0.28 0.21 38.7 5.4 30.0 38.24 10 10.4 6.9 0.23 0.23 6.2 13.0 6.3 8.7 0.0005

7.5 0.10 Morgan 0.8 0.22 0.43 0.22 8.0 2.9 15.9 3.45 10 10.4 6.0 0.74 0.58 -1.6 2.2 1.2 2.6 0.0041

7.5 0.10 Morgan 1.5 0.22 0.43 0.22 15.9 2.9 15.9 3.45 10 10.4 6.0 0.30 0.31 -1.1 2.2 1.6 2.6 0.0016

7.5 0.10 Morgan 1.9 0.22 0.43 0.22 20.2 2.9 15.9 3.45 10 10.4 6.0 0.20 0.23 1.0 2.2 1.7 2.6 0.0011

7.5 0.16 Morgan 2.4 0.22 0.43 0.22 26.3 3.7 22.0 7.43 10 10.4 6.0 0.15 0.19 -1.3 2.9 2.0 3.6 0.0012

7.5 0.10 Morgan 2.9 0.22 0.43 0.22 30.0 2.9 15.9 3.45 10 10.4 6.0 0.15 0.18 1.0 2.2 2.0 2.6 0.0008

7.5 0.10 Morgan all 0.22 0.43 0.22 72.8 2.9 15.9 3.45 30 10.4 6.0 0.14 0.18 1.1 2.2 2.1 2.6 0.0008

15.0 0.10 Morgan 1.9 0.30 0.44 0.22 38.7 5.4 30.0 38.24 10 10.4 6.0 0.42 0.48 7.5 13.0 6.6 8.7 0.0009

15.0 0.10 Mexico 1.9 0.10 0.16 0.23 38.7 5.4 30.0 38.24 10 10.4 8.1 0.10 0.10 5.6 13.0 6.1 8.7 0.0002

7.5 0.10 Morgan 1.9 0.30 0.59 0.29 20.2 2.9 15.9 3.45 10 10.4 6.0 0.34 0.37 1.1 2.2 1.9 2.6 0.0018

7.5 0.10 Morgan all 0.30 0.59 0.29 72.8 2.9 15.9 3.45 30 10.4 6.0 0.22 0.28 1.2 2.2 2.3 2.6 0.0012

15.0 0.10 Morgan 1.9 0.40 0.59 0.29 38.7 5.4 30.0 38.24 10 10.4 6.0 0.66 0.75 8.0 13.0 7.1 8.7 0.0014

7.5 0.10 Mexico 1.9 0.10 0.22 0.31 20.2 2.9 15.9 3.45 10 10.4 8.1 0.12 0.13 0.9 2.2 1.7 2.6 0.0007

7.5 0.10 Mexico 2.9 0.10 0.22 0.31 30.0 2.9 15.9 3.45 10 10.4 8.1 0.07 0.09 1.0 2.2 2.0 2.6 0.0004

15.0 0.10 Capitola 1.9 0.30 0.43 0.32 38.7 5.4 30.0 38.24 10 10.4 6.9 0.42 0.44 7.2 13.0 6.9 8.7 0.0009

15.0 0.10 Joshua 1.9 0.27 0.35 0.33 38.7 5.4 30.0 38.24 10 10.4 7.4 0.67 0.68 8.3 13.0 7.5 8.7 0.0014

15.0 0.16 Joshua 2.2 0.27 0.35 0.33 48.2 6.9 39.5 92.45 10 10.4 7.4 0.44 0.51 -11.0 17.9 7.8 11.9 0.0016

7.5 0.10 Morgan 1.9 0.40 0.78 0.39 20.2 2.9 15.9 3.45 10 10.4 6.0 0.51 0.56 1.4 2.2 2.0 2.6 0.0028

15.0 0.10 Capitols 1.9 0.40 0.57 0.43 38.7 5.4 30.0 38.24 10 10.4 6.9 0.62 0.66 7.7 13.0 7.1 8.7 0.0013

15.0 0.16 Capitola 2.2 0.40 0.57 0.43 48.2 6.9 39.5 92.45 10 10.4 6.9 0.48 0.56 -10.1 17.9 7.9 11.9 0.0017

15.0 0.10 Mexico 1.9 0.20 0.32 0.45 38.7 5.4 30.0 38.24 10 10.4 8.1 0.41 0.43 7.9 13.0 7.3 8.7 0.0009

7.5 0.10 Joshua 1.5 0.27 0.52 0.50 15.9 2.9 15.9 3.45 10 10.4 7.4 0.73 0.69 -1.7 2.2 1.5 2.6 0.0040

7.5 0.10 Joshua 1.9 0.27 0.52 0.50 20.2 2.9 15.9 3.45 10 10.4 7.4 0.52 0.52 -1.5 2.2 2.3 2.6 0.0028

7.5 0.16 Joshua 2.4 0.27 0.52 0.50 26.3 3.7 22.0 7.43 10 10.4 7.4 0.38 0.43 -2.2 2.9 1.9 3.6 0.0031

7.5 0.10 Joshua 2.9 0.27 0.52 0.50 30.0 2.9 15.9 3.45 10 10.4 7.4 0.36 0.39 -1.3 2.2 2.2 2.6 0.0019

7.5 0.10 Joshua all 0.27 0.52 0.50 72.8 2.9 15.9 3.45 30 10.4 7.4 0.31 0.37 1.3 2.2 2.4 2.6 0.0017

7.5 0.10 Capitols 0.8 0.40 0.76 0.58 8.0 2.9 15.9 3.45 10 10.4 6.9 1.39 1.10 -1.9 2.2 1.2 2.6 0.0076

7.5 0.10 Capitols 1.5 0.40 0.76 0.58 15.9 2.9 15.9 3.45 10 10.4 6.9 0.72 0.72 -1.6 2.2 1.5 2.6 0.0039

7.5 0.10 Capitola 1.9 0.40 0.76 0.58 20.2 2.9 15.9 3.45 10 10.4 6.9 0.57 0.60 -1.3 2.2 2.1 2.6 0.0031

7.5 0.16 Capitola 2.4 0.40 0.76 0.58 26.3 3.7 22.0 7.43 10 10.4 6.9 0.46 0.53 -2.2 2.9 1.7 3.6 0.0038

7.5 0.10 Capitola 2.9 0.40 0.76 0.58 30.0 2.9 15.9 3.45 10 10.4 6.9 0.47 0.52 -1.3 2.2 2.1 2.6 0.0026

7.5 0.10 Capitols all 0.40 0.76 0.58 72.8 2.9 15.9 3.45 30 10.4 6.9 0.45 0.52 1.3 2.2 2.4 2.6 0.0024

7.5 0.10 Mexico 1.9 0.20 0.44 0.62 20.2 2.9 15.9 3.45 10 10.4 8.1 0.49 0.51 -1.5 2.2 2.1 2.6 0.0027

15.0 0.10 Mexico 1.9 0.30 0.48 0.68 38.7 5.4 30.0 38.24 10 10.4 8.1 0.79 0.87 -9.0 13.0 7.4 8.7 0.0017

15.0 0.10 Mexico 1.9 0.35 0.56 0.79 38.7 5.4 30.0 38.24 10 10.4 8.1 0.96 1.08 -10.4 13.0 7.4 8.7 0.0020

15.0 0.16 Mexico 2.2 0.35 0.56 0.79 48.2 6.9 39.5 92.45 10 10.4 8.1 0.70 0.84 -15.5 17.9 8.2 11.9 0.0025

15.0 0.10 Mexico 1.9 0.40 0.64 0.91 38.7 5.4 30.0 38.24 10 10.4 8.1 1.14 1.30 -11.6 13.0 7.4 81 0.0024

7.5 0.10 Mexico all 0.30 0.66 0.94 72.8 2.9 15.9 3.45 30 10.4 8.1 0.50 0.67 -2.2 2.2 2.3 2.6 0.0028

7.5 0.10 Mexico 0.8 0.35 0.77 1.09 8.0 2.9 15.9 3.45 10 10.4 8.1 3.60 2.09 -1.4 2.2 1.5 2.6 0.0197

7.5 0.10 Mexico 1.5 0.35 0.77 1.09 15.9 2.9 15.9 3.45 10 10.4 8.1 1.49 1.37 -2.7 2.2 1.5 2.6 0.0081

7.5 0.10 Mexico 1.9 0.35 0.77 1.09 20.2 2.9 15.9 3.45 10 10.4 8.1 1.07 1.12 -2.5 2.2 2.1 2.6 0.0059

7.5 0.16 Mexico 2.4 0.35 0.77 1.09 26.3 3.7 22.0 7.43 10 10.4 8.1 0.87 0.96 -3.3 2.9 1.8 3.6 0.0071

7.5 0.10 Mexico 2.9 0.35 0.77 1.09 30.0 2.9 15.9 3.45 10 10.4 8.1 0.83 0.95 -2.7 2.2 2.4 2.6 0.0045

7.5 0.10 Mexico all 0.35 0.77 1.09 72.8 2.9 15.9 3.45 30 10.4 8.1 0.61 0.80 -2.5 2.2 2.4 2.6 0.0033