Download as pdf or txt
Download as pdf or txt
You are on page 1of 171

‘ .

/' ’//L '"


CEP 702
FEB 1990

N T E R N A T I O N AL
D N S U L T A N TS

Retaining Wall Design Notes


CEP 702
FEB 1990

INTERNATIONAL
CONSULTANTS

Retaining Wall Design Notes

Published by: Group Office


PO Box 12343
Wellington

Tnis document and its contents is the property of Opus International Consultants Limited.
l ny unauthorised employment or reproduction, in full or part is forbidden

© Opus International Consultants Limited 1997


PREFACE

The aim of these notes on Retaining wall design is to provide the


designers within WORKS, guidance on design methods, bringing together the
commonly used charts. The revisions in this edition generally incorporate
up to date design methods, and include major revisions to the section on
Earthquake Earth Pressures and Design. A brief new section on the design
of sheet retaining structures is aiso included.
The section on Earthquake Earth Pressures was prepared by Dr John wood of
Phillips and wood Limited, Lower Hutt and has been incorporated into this
document.
This revised document was prepared by the Special Project Office of works
Consultancy Services, wellington, with useful comments from other
divisions of works Consultancy Services and the Geotechnical Control
Office in Hong Kong.

Special Projects Office


January 1990
1

RETAINING WALL DESIGN NOTES

CONTENTS

CONTENT S

LIST OF FIGURES

LIST OF TABLES

SYMBOLS

SECTION 1 — INTRODUCTION

Scope
Y-4l—4 II l\)|-\ Definitions and Symbols
1.3 Design Principles
1.3.1 Free Standing Retaining walls
1.3.2 Other Retaining Structures
1'4 Load Cases
1.4.1 Basic Loadings
1.4.2 Other Considerations

SECTION 2 — SOIL PROPERTIES

General
Selection and Use of Backfill
Density
Effective Stress and Pore Pressures
Shearing Strength
Base Friction
Wall Friction
Modulus of Elasticity and Poisson's Ratio
Coefficient of Subgrade Reaction
Swelling and Softening of Clays
Permeability
I\!l'JI\1lJI\>l)\>l f\)lH\)l IIO
I I-‘l—P ‘£DO)\l ‘(Jl- >(»Jl'\)—-\|\>|-AC3
OI
IIOII Liquefaction
2

Page

SECTION 3 - STATIC EARTH PRESSURE


States of Stress 32
(.00.) I l\)r—- Amount and Type of Wall Movement 32
co O co ing Equilibrium Conditions 34
The Rankine Earth Pressure Theory
The Coulomb Earth Pressure Theory
The Trial wedge Method
Earth Pressure in Soils with Cohesion
Passive Earth Pressures
Geometrical Shape of the Retaining Structure
w cow wui-' IOIOOIO
—lI w z.ow cow5 IOIIIIO
_ln \|mu1-I: w|\)+—¢-r Limited Backfill
L0 I -I-> ic Equilibrium Conditions 39
At-Rest Pressures
Over-Consolidation Pressures
Elastic Theory Methods
UJ(»0(.\)(4dl'T1 0_.|00->-§-l>-l>F-U;
10| 000-‘>UOIQl—‘l"|' Compaction Induced Earth Pressures

SECTION 4 - EARTHQUAKE EARTH PRESSURES AND DESIGN


4.1 1 ound ‘ 42
wall Categories
Soil - Structure Interaction
Simplifications for Design
Plastic Theory and Failure Modes
-h-b-l>->-Q)
k0wo0 |—I|-\—l|—=|—*O 7? U1-¥>(».l'\)|—=LQ
0 0I Resonance Effects
Design Seismic Coefficients 45
Load Combinations 46
-54>-l>UII -l>(.ul\> Factors of Safety 46
4.5 ic Forces and Pressure Distributions 46
Rigid wall
Stiff wall
Flexible wall
Displaceable wall
Forced Wall
-I3-I>-I>-h l>-l>C0J0l<
U'lU'lU'lU'lO'lO"lD Q]
00 U‘)U'l-P(.U|'\!ld5
I0000 Water Pressures
4.6 ('7 ations to Various Types of walls 53
Free Standing walls Founded on Soils
Free Standing walls Founded on Rock or Piles
Tied Back walls
Basement walls
-5-h-lb-l>-I>cI 00'0
Q§Q'§o7C§O§'C) 0n__|00U1-¥>b0I\)v- h Bridge Abutments
3

Page

SECTION 5 - EFFECT OF SURCHARGES

Uniform Surcharges 59
Line Loads 59
Point Loads 60
(J'|(J'I(J1(T| II -§UO|\]l—-\ Non-uniform Surcharges 60

SECTION 6 - EFFECTS OF WATER

General 61
Static water Level 61
Seepage Pressure 62
Dynamic water Pressure 62
Drainage Provisions 63
Filter Requirements 64
0505030703070) I000
III\lo)(J'|- >(.O|\)l—‘ Geotextiles 64

SECTION 7 - STABILITY OF RETAINING WALLS


7.1 General 66
7.2 Z g Stability 67
Base without a Key
\l\lU7 II
._a [\)[Q_..|- Q
0I |\)|-\._.|- Base with a Key
7.3 Overturning Stability 68
7.3.1 Gravity and Counterfort walls
7.3.2 Sheet walls
7.4 ation Bearing Pressures 69
Vertical Central Loads
Inclined Loads, Tilted Base and Sloping Ground
Miscellaneous Foundation Conditions
0 -I>-P-§~|>CZ
\.|\l\l\|"|1 Q
II0 oI0O ->L.oI\J|—-Q. Foundations on Rock
7.5 Slope Failure in Surrounding Soil 74

SECTION 8 - STRUCTURAL DESIGN


(D O |-1 DJ d 75
Codes
Limit State Design
Cover to Reinforcement
(DCO(I)@fi1 (D
I I- |—\|- r—\3 fl)
I 0I0 ->L».H\)l—\'1 Selection of wall Type
8.2 Toe Design 76
4

Page

8G3 Design 76
Stem Loading
Lower Section of Counterfort wall Stem
(D@@(/7 OI(-1-I(A)U.)(.0(D B(-\)|\)l—‘
II
0 Horizontal Moments in Counterfort wall Stem
8.4 Heel Slab Design 77
8.4.1 Loading
80402 Heel Slabs for Counterfort walls
Count erfort Design 78
Key D esign 78
@@Q7 IQI \|O"!(J1 Contr ol of Cracking 79

SECTION 9 - SPECIAL PROVISIONS FOR CRIB WALLS

Gener al 80
Desig n Loading 80
Found ation Depth 80
Drain age 80
Multi ple Depth walls 81
Q CQ OQDCDQ I00I0
ICDC”->OO|'\)l—\ walls Curved in Plan 81

SECTION 10 - SHEET RETAINING STRUCTURES


10.1 General 82
10.2 Strutted Excavations 82
10.3 Anchored Flexible walls 84
10.3. 1 walls Anchored Near the Top
10.3.2 Multiple Anchored walls
10.3.3 Effects of Anchor Inclination
10.4 Cantilevered walls 85

Referen CBS 86
Figures
5

LIST OF FIGURES

Loading on typical retaining wall


Rankine earth pressure, cohesionless soil, constant backfill slope
Rankine active earth pressure coefficients, cohesionless soil, uni-
form sloping backfill
Rankine earth pressure, soil with cohesion, horizontal ground
surface
Coulomb earth pressure, cohesionless soil, constant backfill slope
Coulomb active earth pressure coefficients, cohesionless
soil, uniform sloping backfill ¢ = 25°
ll ll ll ll ll ll ¢ = O

ll ll ll ll ll ll ¢ = O

ll II ll ll ll ll ¢ = O

Coulomb failure plane for active pressure, cohesionless soil with


uniform sloping backfill
Trial wedge method, cohesionless soil, irregular ground surface
Trial wedge method, cohesionless soil, Culmann's construction
Trial wedge method, soil with cohesion, irregular ground surface
Trial wedge method, layered soil and pore water pressures

(A) Approximate method for direction of Rankine earth pressure.


(B) Geometric conditions for Rankine and Coulomb methods.
Point of appiication of active pressure
Point of application of resultant pressure and pressure distribution
Earth pressure coefficients, sloping ground
Earth pressure coefficients, sloping wall
Earth pressure due to compaction
Seismic zone factor
Stability analysis of retaining wall under earthquake loading
Earthquake increment of pressure on rigid retaining wall
6

Equivalent wall height for estimating earthquake earth pressures for


wall with sloping backfill
Earthquake increment of earth pressure - stiff and flexible walls
Mononobe - Okabe earthquake earth pressure - cohesionless soil,
constant backfill slope
Mononobe - Okabe active earthquake earth pressure increments for
vertical walls : Horizontal Backfill : 8 = ¢ : ¢ = 20° - 45°
Mononobe - Okabe active earthquake earth pressure increments for
vertical walls : Sloping Backfill : 8 = ¢

(A) <l>= 20°

(B) ¢= 25°

Mononobe - Okabe active earthquake earth pressure increments for.


vertical walls : Sloping Backfill : 6 = ¢

(C) d>= 30°

(D) ¢= 35°

Mononobe - Okabe active earthquake earth pressure increments for


vertical walls : Sloping Backfill : 8 = ¢

(E) ¢= 40°

Mononobe - Okabe active earthquake earth pressure increments for


vertical walls : Horizontal Backfill : 8 = 2¢/3, ¢ = 20° — 45°
Mononobe - Okabe active earthquake earth pressure increments for
vertical walls : Sloping Backfill : 8 = 2¢/3

(A) ¢= 20°

(B) ¢= 25°

Mononobe - Okabe active earthquake earth pressure increments for


vertical wall : Sloping Backfill : 6 = 2¢/3

(C) <l>= 30° ,

(D) ¢= 35°

Mononobe - Okabe active earthquake earth pressure increments for


vertical walls : Sloping Backfill : 6 = 2o/3

(E) ¢= 40°
Mononobe - Okabe active earthquake earth pressure increments for
non-vertical walls : wall Slope = —14° : Horizontal Backfill :
6 = 2¢/3 : ¢ = 20° - 45°
7

Mononobe - Okabe active earthquake earth pressure increments for


non-vertical walls : wall Slope = -14°; Sloping Backfill; 6 = 2¢/3

(A) ¢ = 20°
(5) ¢ = 25°
Mononobe - Okabe active earthquake earth pressure increments for
non-vertical walls : wall Slope = -14° : Sloping Backfill : 8 = 2¢/3

(C) ¢ = 30°
(D) ¢ = 35°
Mononobe - Okabe active earthquake earth pressure increments for
non-vertical walls; Wall Slope = -14°; Sloping Backfill; 8 = 2¢/3
(E) ¢ = 40°
Trial wedge method for earthquake force on wall
Displaceable wall - wall displacements from Sliding Block Theory -
Mononobe - Okabe passive earthquake earth pressure for vertical
walls, horizontal backfill and zero wall friction

(A) Earthquake passive earth pressure increment.


(B) Total passive pressure.
Forced wall - earth pressure distributions
(A) Rotated wall
(B) Translated wall
Static and Earthquake Pressures on walls with Submerged Backfill
Failure surfaces for tied back wall under earthquake loading
Earthquake and gravity forces on bridge abutments
Uniform surcharge load cases
Trial wedge active pressure including line load
Lateral pressure distribution on wall due to point and line loads
Static water pressure, and seepage pressure for drained walls

Drainage system for low walls and crib wall drainage


Drainage details for backfilled walls
Permeability of drainage materials
8

Stability criteria for retaining walls


Sheet wall stability - cohesionless soils
Bearing capacity factors
(A) Loading arrangements for Brinch Hansen's bearing capacity
formulae
(B) Transformation of eccentrically loaded foundations
Brinch Hansen bearing capacity formula - shape factors
Brinch Hansen bearing capacity formula - depth factors
Brinch Hansen bearing capacity formula - load inclination factors
Design loading on heel slab
Cut-off positions of main tensile steel in counterfort
Crib wall design curves, normal loading, ¢ = 30°, 40°
Crib wall design curves, earthquake loading, o = 30°, 40°
Pressures on typical crib wall
Braced Excavation Pressure Distributions
Factor of Safety with Respect to Base Heave
Large Excavations Settlement Guide
9

LIST OF TABLES

Representative Values for Unit weight of Soils

Representative Values for the Angle of Shearing Resistance

Typical Friction Angles and Adhesion Values for Bases Without Keys

Indicative Proportions of Maximum wall Friction Developed

Modulus of Elasticity for Selected Soils (Undrained Compression)

Typical Values for Poisson's Ratio

Coefficient of Subgrade Reaction (Vertical)

Permeabilities of Soils

Movement of Wall Necessary to Produce Active Pressures

Allowable Bearing Pressure on Jointed Rock


10

SYMBOLS

A. effective area of base


B base width
B. effective base width

Co seismic response function (for a 150 year return period)


C(T) horizontal earthquake coefficient (period T)
9(0) horizontal earthquake coefficient (period T = O)
c cohesion of soil in terms of total stress

Cb adhesion at base
C. cohesion of soil in terms of effective stress
D foundation depth

D20 particle size at which 20% of the material is finer


d forward displacement of centre of mass due to an
earthquake
d¢,dq,dY foundation depth correction factors
Es modulus of elasticity of soil
e eccentricity of load on base
Fs factor of safety
9 gravitational acceleration
H,H1,etc vertical height of plane on which earth pressure is calcu-
lated (from underside of base or bottom of key to ground
surface)
H horizontal component of loading on the base
HI
vertical height of wall

hc depth below which active pressure exceeds compaction


induced earth pressure

he equivalent height of fill to uniform surcharge loads

hw piezometric head
i hydraulic gradient
'l(;.Iq.IY foundation load inclination factors
11

K0 coefficient of earth pressure at rest

KA coefficient of active earth pressure

KAE coefficient of active earthquake earth pressure (KA+AKAE)


AKAE coefficient of active earthquake earth pressure increment

Kn coefficient of passive earth pressure

Ks coefficient of subgrade reaction


k coefficient of permeability
L length of base
L. effective length of base

Ls length of retaining wall stem


I length of failure surface
N normal reaction on a soil failure surface
N-value standard penetration test blow count for 300 m
penetration
N threshold acceleration factor
N6 stability factor

NcsNqsNY bearing capacity factors

Ns slope stability number


P resultant force due to lateral pressure

PA force due to active lateral earth pressure

PAE force due to active lateral earthquake earth pressure


(PA+APAE)
PE force due to earthquake pressure component from inertia
forces in soil

PF force due to earthquake pressure component from forcing of


the wall against the soil

PFX force on forced wall due to translation of wall

PFo1 force on forced wall due to rotation of wall

PH horizontal component of force due to lateral earth


pressure

P1 inertia load acting on abutment mass

PL load from superstructune


12

P0 force due to at-rest earth pressure


Pp force due to passive earth pressure
PPN force due to net passive pressure
PQ force due to lateral earth pressure due to line or point
surcharge load (per unit length of wall)
P5 force on wall due to gravity pressure component
PV vertical component of force due to lateral earth pressure
PW force due to hydrostatic water pressure
APAE increment in force due to active earth pressure due to an
earthquake
APE increment of earthquake force
p effective surface line load due to compaction plant
pA active earth pressure
pc consolidation pressure
pw water pressure
p(z) earth pressure at depth z below ground surface
Ap(z) earthquake pressure increment at depth z below surface
Q total load
QL line load
Qp point load
q pressure on base or surcharge load intensity
qa allowable soil bearing pressure

qd flow rate through drain


qujt ultimate soil bearing pressure
qu unconfined compressive strength
R risk factor (seismic design)
R,RA,Rp,Rw,etc resultant forces
ROD rock quality designation
S total shearing resistance at underside of base
s shearing strength of soil
Su undrained shear strength of soil
13

5¢.5q,5Y foundation shape correction factors

5cL, $cB’ 5qL=


5qB» 5YL' SYB foundation shape correction factors for inclined loading
T tangential force along a failure surface
U resultant force due to pore water pressures
u pore water pressure
V vertical component of resultant of loading on the base
Vo peak ground velocity (seismic)
W weight of soil wedge used in calculation of earth
pressures

Vb weight of backfill over heel of wall

ww weight of wall

Wt total weight of wall, soil above toe and soil above heel
x horizontal translational displacement of wall
Ax displacement of superstructure
Z earthquake zone factor

Zo vertical depth of tension crack in cohesive soil


Zc critical depth for compaction pressure distribution
2 depth below ground surface

AA inclination of failure plane from the horizontal for


active state (degrees)

GAE inclination of failure plane from the horizontal for


active state under earthquake conditions

“P inclination of failure plane from the horizontal for


passive state (degrees)
B slope of back of the wall (degrees, anti-clockwise positi-
ve)

Y bulk unit weight of soil (kN/m3)


Y. submerged unit weight of soil = Ysat - Yw

Yd dry unit weight of soil


Ysat unit weight of saturated soil

Yw unit weight of water


14

A increment
s angle of wall friction (anti-clockwise positive)

$6 angle of base friction


e angle = tan'1 C(O)

91 rotation of forced wall

96 rotation of base of wall


V Poisson's ratio
P settlement
E angle measured clockwise from vertical to direction of PA
O total normal stress

°ho horizontal earth pressure


OI
effective normal stress
o'v effective vertical overburden pressure
T shear stress
¢ angle of shearing resistance in terms of total stress
¢. angle of shearing resistance in terms of effective stress
¢u angle of shearing resistance under undrained conditions

W angle of inclination of loading on base


w inclination of ground slope
Section 1

INTRODUCTION
15

SECTION 1 - INTRODUCTION

1.1 SCOPE
These notes are intended as a guide for use in the estimation of
earth pressure forces and the design and construction of
retaining walls and similar earth retaining structures. These
notes are not intended to be used as a detailed text
encompassing all aspects of retaining wall design. The
intention is to provide brief notes and recomended methods
covering most aspects of design. If a more detailed knowledge
of a particular subject is required, the references given should
prove helpful. Reference is also made to standard texts for
detailed methods such as the construction of flow nets for pore
water pressure determination, and reinforced concrete design
methods.
Aspects such as the use of classical earth pressure equations,
the effect of earthquakes on earth pressures, and allowable
bearing pressures under inclined loads are covered in detail.
Engineering judgement must always be used when applying the
theories and methods given in these notes and strict notice must
be taken of the limitations of the various assumptions.
Special retaining systems such as reinforced earth and soil
nailing are beyond the scope of these notes, and reference
should be made to specialist literature.
1.2 DEFINITIONS AND SYMBOLS
Throughout these notes, static earth pressure means the pressure
exerted by the earth due to gravity forces. Earthquake earth
pressure means the combined static and dynamic earth pressure
which acts during or because of an earthquake.
A list of symbols used, with their meanings, is included in the
front of these notes.
1.3 DESIGN PRINCIPLES
3 FFGG $I1&fld‘II'lQ RE‘C6.‘Il'l'IflQ NGIIS

In the design of free standing retaining walls, the following


aspects need to be investigated:
(a) the stability of the soil around the wall;
(b) the stability of the retaining wall itself; and
(c) the structural strength of the wall.
For these walls it is usual to assume that some outward movement
of the wall takes place so that the lateral earth pressure from
16

the retained soil is a minimum (active condition) for both


static and earthquake loadings. However, the designer should
check that the required movement can take place and that it does
not affect the serviceability or appearance of the wall or cause
damage to nearby structures or services. If the deformation
that is required to reduce the earth pressure to the active case
is not available due to the rigid nature of the structure or
foundation, either the wall must be designed to withstand a
higher pressure or some change made to the structure or
foundation. If cohesive backfill is used, the large
displacements necessary for the active condition means that the
lateral earth pressure will almost always be higher than the
active value.
For the determination of earth pressures it is usual to consider
only a unit length of the cross-section of the wall and retained
soil. A unit length is also used in the structural design of
cantilever walls and other walls with a uniform cross-section.
3 Other Retaining Structures
Where an earth retaining wall is part of a more extensive
structure (eg, a basement wall in a building or an abutment wall
of a portal structure) or is connected to another structure (eg,
a bridge abutment connected to the superstructure) the wall is
usually subject to static earth pressures which are greater than
active since the structure does not allow full "yielding" of the
soil. In these cases, the main structure generally provides the
stability for the wall which then only needs to have adequate
structural strength.
The earth pressure on this type of structure under earthquake
conditions depends on the movements of the structure and the
forces exerted on the wall by the rest of the structure as well
as the inertia forces from the soil.

1.4 LOAD CASES


4 Basic Loadings
Two basic earth pressure loadings are considered for design.
These are:
(a) Normal loading = Static effective earth pressure +
water pressure + effective pressure
due to live loads or surcharge.
(b) Earthquake loading = Earthquake earth pressure + water
pressure + surcharge (but not live
loads).
However, earth retaining structures should be designed for not
less than the pressure due to a fluid with a unit weight of
4 kN/m‘.
17

For normal loading, static earth pressure and pressure due to


surcharge should be derived in terms of effective stress unless
short term load conditions exist, see section 2.5. For short
term loading conditions earth pressures should be derived in
terms of total stress.
For many walls of lesser importance, earthquake loading need not
be applied, see section 4.
4 Other Considerations
Consideration should also be given to the possible occurrence of
other design cases or variations within the two design cases
given above, caused by the construction sequence or future
development of surrounding areas. For instance, additional
surcharges should be considered in calculating active pressures
and allowance made for any possible future removal of ground in
front of the wall. Usually the passive resistance of the
material in front of the wall is ignored in the design of
gravity or cantilever retaining walls.
Section 2

SOIL PROPERTIES
18

SECTION 2 - SOIL PROPERTIES

GENERAL

In advance of the design, tests should preferably be carried out


on the proposed backfill material and natural ground behind and
under an earth retaining structure. It is good practice to make
further soil tests on the material exposed after excavation.
For all walls higher than 6 m, especially those with sloping
backfill, the soil properties of the natural ground and backfill
should be estimated from tests on samples of the materials
involved. For less important walls, an estimation of the soil
properties may be made from previous tests on similar materials.
However, a careful visual examination of the material, particu-
larly that at the proposed foundation level, should be made with
the help of identification tests to ensure that the assumed
material type is correct.
SELECTION AND USE OF BACKFILL '

The ideal backfill is a free draining granular material of high


shearing strength. However, the final choice of material should
be based on the costs and availability balanced against the
desired properties.
In general the use of cohesive backfills is not recommended.
Clays are subject to seasonal variations, swelling (see section
2.10), and deterioration which all lead to an increase in
pressure on a wall. They are difficult to consolidate and long
term settlement problems are considerably greater than with
cohesionless materials. For cohesive backfills, special
attention must be paid to the provision of drainage to prevent
the build-up of water pressure.
The wall deflection required to produce the active state in
cohesive materials may be up to 10 times greater than that for
cohesionless materials. This, together with the fact that the
former generally have lower values of shearing strength, means
that the amount of shearing strength mobilised for any given
wall movement is considerably lower for cohesive materials than
for cohesionless materials. The corresponding active earth
pressure for a particular wall movement will therefore be higher
if cohesive soil is used for backfill.
In cases of a high seismic coefficient and for a steeply sloping
backfill, the active earth pressure will be substantially
reduced if the failure plane occurs in a material with a high
angle of shearing resistance. In some circumstances it may be
economical to replace the existing soil by a material with a
high angle of shearing resistance so that the above situation
occurs. However, also see section 3.3.6.
19

It is essential to specify and supervise the placing of backfill


to ensure that its properties (¢, c and Y) agree with the design
assumptions both for lateral earth pressure and dead weight
calculations.
DENSITY
The density of soil depends on the specific gravity of the solid
particles and the proportions of solids, air and water in the
soil. The typical specific gravity of soil particles is about
2.65 for sand or rock and 2.70 for clays. However this may vary
from area to area. The proportion of the total volume that is
made up of this solid material is dependent on the degree of
compaction or consolidation.

An estimate of the density of backfill material to be used


behind a retaining structure may be obtained from laboratory
compaction tests on samples of the material. The density chosen
must correspond to the compaction and moisture conditions that
will apply in the actual situation. .

The density of natural soil should be obtained from undisturbed


samples kept at the field moisture content, and volume. In
earth pressure calculations, density must be in force units, ie,
Unit weight (N/m’ or kN/m’). For low, relatively unimportant
walls, the unit weight of the soil behind the wall may be
estimated from the typical values given in Table 1. In general
the saturated unit weight should be used in calculations
involving clay.
EFFECTIVE STRESS AND PORE PRESSURE
An effective stress may be considered to be the stress trans-
mitted through the points of contact between the solid particles
of the soil. It is this stress that determines the shearing
resistance of the soil. The effective stress, 0', at any point
in a saturated soil mass may be obtained by subtracting the
pressure transmitted by water in the voids, u (pore water
pressure), from the total stress, o, thus:
o‘ = 0 - u

An increased pore water pressure gives a reduced effective


stress and therefore a reduced soil shearing resistance. This
leads to an increased force against a wall in the active case.
Conversely, an increase in the negative pore pressure (ie, a
pore suction) gives an increased shearing resistance and reduces
the force against a wall in the active case.
2O

TABLE 1: REPRESENTATIVE VALUES FOR UNIT HEIGHT OF SOILS


(Basic Data from Terzaghi and Peck (1967) and
US Department of the Navy (1982))

Unit height

Material Dry: Ysat


(kN/m‘) (kN/mi)
Clean gravel or rock -
loose 16 - 17 19 — 20
dense, poorly graded 18 - 20 20 - 22
dense, well graded 20 - 21
well graded, clean san ds,
gravelly sands —
loose 14 — 16 19
dense 17 — 2O 20

Poorly graded clean sa nd,


sand-gravel mix - loos e 16 - 17 20
dens e 17 - 19 21

Clayey sand -
loose, poorly graded 14 - 17
dense, poorly graded 16,- 18
Fine silty sands and s ilt -
loose 14 - 16
dense 17 - 19
Sand-silt clay mixed w ith
slightly plastic fines 17 - 20
Clayey gravel, poorly graded
gravel-sand clay 18 - 20
Silty gravel, poorly g raded
gravel-sand silt 19 - 21
Glacial till - very mi xed grained 20 - 21 23

Glacial clay - soft 16 - 19


stiff 20 — 21

Organic clay -
soft slightly organi c 15 — 16
soft very organic 13 - 14

Pumice 10 - 12
21

Positive pore water pressure results from a number of factors,


the most important being static water pressure, seepage of
groundwater or rainfall and seepage from other sources, such as
burst or leaking water supply mains. In some soils, shock or
vibration can cause transient increases in pore pressure. In
low permeability soils, changes in pore water pressure can
result when changes in total stress due to ground loading,
dewatering or excavation are more rapid than the pore water can
flow. These pore pressures dissipate with time, but may need to
be considered in design. Pore water pressures due to static
water pressure and seepage of water are covered in chapter 6.
Negative pore pressures are present in many partially saturated
soils as a result of capillary tension. Capillary tension, and
hence soil suction, may be destroyed by surface infiltration or
seepage, and in general its beneficial effect on the shear
resistance of the soil should not be used in retaining wall
design.
SHEAR STRENGTH

In all earth pressure problems the magnitude of earth pressure


on a particular structure is a function of the shear strength of
the soil. The shear strength is not a unique property of the
material but depends upon the conditions to which the soil is
subjected when it is sheared.
Shear strength is also a function of effective stress and water
content, and is dependent on volume change of the soil. This
gives rise to two separate conditions which determine the type
of analysis to be carried out. These are:
(a) "Short term" (or undrained) which applies when the water
content of the soil cannot change rapidly and hence at the
end of construction excess pore water pressures have not
dissipated. This comonly occurs in saturated soils with
low permeability such as saturated clays. The shear
strength of such a soil does not change when it is sheared
quickly and the undrained shear strength may be used to
calculate earth pressures.

Analysis of "short term" stability may be carried out in


terms of total stress, o, and strength parameter, Su
(undrained shear strength) with ou = 0.

(b) "Long term" (or drained) conditions, apply when the water
content of the soil can change quite rapidly with a conse-
quent change in pore pressure and hence with a change in
shear stress. This generally occurs in cohesionless soils
with a high permeability or when after a long period of
time excess pore water pressures have dissipated in soils
with lower permeability.
In this case it is necessary for earth pressures to be
calculated from shear strengths expressed in terms of
effective stresses 0', and strength parameters c‘ and o‘.
22

If analysis is carried out in terms of effective stresses the


effect of any field pore water pressure must be included in the
analysis.
The shear strength of a soil is proportional to the normal or
confining stress acting on the failure plane. The maximum shear
stress that a sample of soil can sustain under different normal
stresses should be obtained from laboratory tests or site
investigation, see WORKS publication on Site Investigation
(WORKS, 1982). Results of laboratory testing may be plotted to
form a relationship between shear stress at failure and normal
stress on the soil. This relationship forms an envelope which
is commonly termed the strength envelope. The envelope is
generally curved, particularly in the low stress range. But
portions of the curve can be approximated by a straight line
relationship as follows:

s = c + o tan o (in terms of total stress) or


s = c' + 0' tan o‘ (in terms of effective stress)
Where c and ¢ are termed the strength parameters in terms of
total stress and c' and ¢' are termed the effective strength
parameters.
For important walls it is desirable that the design strength
parameters are determined for the range of stress, moisture
content and density that is appropriate to the field situation.
Guidance on determination of strength parameters is given in
WORKS (1982) and Lambe and Whitman (1969).
For less important walls the values given in Table 2 rnay be
used.
23

TABLE 2: REPRESENTATIVE VALUES FOR THE ANGLE OF SHEARING RESISTANCE

[Values obtained mainly from Terzaghi and Peck (1967)]


(c = O in all the cases except clay where c = qu/2)

¢l
¢ (degrees)
Mater ial (degrees) (saturated)

Sandy gravel or rock fill ing 35-45


Sand -
loose, round grains, un iform 28
dense, round grains, un iform 34
loose, angular grains, well graded 33
dense, angular grains, well graded 45
Silt and silty sand -
loose 27-30 20-22
dense 30-35 25-30
Clayey sand 20-25 14-20
Clay, normally loaded or slightly preconsolidated 22-30 0

2.6 BASE FRICTION


Typical values of friction angle (Sb) and adhesion (cb) for
calculating the shearing resistance between a concrete base and
the foundation material are given in Table 3. These values may
be used for low walls in the absence of specific test data. If
a base key is used the failure plane will generally be through
the foundation soil and therefore the shearing resistance is
that of the soil (Sb = ¢' and cb = c‘).
24

TABLE 3: TYPICAL FRICTION ANGLES AND ADHESION


VALUES FOR BASES WITHOUT KEYS
[Values taken from US Department of the Navy (1982)]

Friction Adhesion
Interface Materials Angle (Sb) (kPa)
Degrees

Mass concrete on the following foundation


material -
Clean sound rock 35
Clean gravel, gravel-sand mixtures,
coarse sand 29-31
Clean fine to medium sand, silty medium
to coarse sand, silty or clayey
gravel 24-29
Clean fine sand, silty or clayey fine
to medium sand 19-24
Fine sandy silt, non-plastic silt 17-19
Very stiff and hard residual or
preconsolidated clay 22-26
Medium stiff and stiff clay and silty
clay 17-19

Formed concrete on the following


foundation material -

Clean gravel, gravel-sand mixtures, well


graded rockfill with spalls 22-26
Clean sand, silty sand-gravel mixture,
single size hard rockfill 17-22
Silty sand, gravel or sand mixed with
silt or clay 17
Fine sandy silt, non-plastic silt 14

Soft clay and clayey silt 10 to 35

Stiff and hard clay and clayey silt 35 to 60

2.7 WALL FRICTION


The magnitude and direction of the wall friction developed
depends on the relative movement between the wall and the soil
In the active case, the maximum value of wall friction develops
only when the soil wedge moves significantly downwards relative
to the rear face of the wall. In some cases, wall friction
cannot develop. These include cases where the wall moves down
with the soil, such as a gravity wall on a yielding foundation
or a sheet pile wall with inclined anchors, and cases where the
25

failure surface forms away from the wall, such as in cantilever


and counterfort walls, see section 3.3.6.
The maximum values of wall friction may be taken as follows:
Timber, steel, precast concrete, 8maX_ = Q‘
2

Cast in-situ concrete, 8maX_ = 2g‘


3

In general, the effect of wall friction is to reduce active


pressure. The effect is small and often disregarded. An
exception to this is when large vertical loads are applied to
the top of the wall, an inclined anchor is stressed to an
appreciable load or the wall is founded on compressible soil.
In these cases the wall has to move down, and because the
friction acts on the soil wedges in a downward direction, this
increases the earth pressures acting on the wall. However, this
can be ignored because limiting conditions are considered in the
calculation of overall stability (Padfield and Mair, 1984).

The effect of wall friction on passive pressures is large (see


section 3). However, considerable structural movements may be
necessary to mobilise maximum wall friction, for which the soil
in the passive zone needs to move upwards relative to the
structure. Generally, maximum wall friction is only mobilised
where the wall tends to move downwards, for example, if a wall
is founded on compressible soil, or for sheet piled walls with
inclined tension members. Some guidance on the proportion of
maximuni wall friction which may develop in various cases is
given in Table 4.
26

TABLE 4: INDICATIVE PROPORTIONS OF MAXIMUM WALL FRICTION DEVELOPED


(Granular Soils - Passive Case)(Rowe & Peaker, 1965)

Proportion of
Maximum Wall
Friction Developed
Structure Type
Loose Dense

Gravit y or free standing walls with


horizo ntal movement. Sheet pile walls 0 0.5
bearin g on hard stratum

Sheet walls with freedom to move downwards


under active forces or inclined anchor 1.0 1.0
loads

Walls where passive soil may settle O 0


under external loads

Anchor age blocks, etc. which have freedom


to mov e upwards on mobilisation of passive_ 0 0
pressu re.

Where a wall will be subjected to significant vibration, wall


friction should not be included, if the soil is cohesionless.

2.8 MODULUS OF ELASTICITY AND POISSON‘S RATIO


The relationships between stress and strain in soils are
important in the settlement of soil-supported foundations. They
also determine the change in earth pressure due to the small
movements of retaining walls or other earth supports. These
relationships are complex since they depend on stress, strain,
time, initial degree of saturation and various other factors.
However, it is often convenient to express them in terms of
modulus of elasticity and Poisson's ratio, since for small
stress differences the soil behaviour closely approximates that
for a perfectly elastic, homogeneous material.
The modulus of elasticity of the soil, ES, is important in
problems where displacements are to be calculated. The value is
usually determined from triaxial compression tests, but plate
bearing tests may be used. Seismic methods may be used to check
a larger mass of material. These methods give higher values of
E5 than those obtained from static testing, especially in
27

jointed rock and results are not directly applicable to problems


of static load problems.
For all soils the elastic modulus increases with increasing
consolidation pressure, pc. For loose sand, E5 approximately
equals 100 pc. A range of values for the modulus of elasticity
in compression for common selected soils is given in Table 5.
Poisson's ratio, v, is very important in stress oriented
problems (eg, stresses on retaining walls for no wall movement)
since it controls the relationship between orthogonal stresses.
It may be determined from triaxial tests; however like the
elastic modulus, it is dependent on the confining pressure and
rate of loading amongst other factors. For granular or normally
consolidated materials, v may be estimated from the relation-
ships for at-rest pressure coefficients, see section 3.4.
Representative values are given in Table 6.

TABLE 5: MODULUS OF ELASTICITY FOR SELECTED


SOILS (UNDRAINED COMPRESSION)

[Values Taken from Bowles (1982)]

Soil E5 (MPa)

Very soft clay 15


Soft clay 25
Medium clay 50
Hard clay 100
Sandy clay 250
Silty sand 20
Silt 20
Loose sand 25
Dense sand 80
Loose sand and gravel 145
Dense sand and gravel 190
Loess 60
Sandstone 6,900 20,600
Limestone 13,800 41,300
Basalt 48,200 89,500
28

TABLE 6: TYPICAL VALUES FOR POISSON'S RATIO


[Values Taken from Bowles (1982)]

Soil v

Clay, saturated 0.4-0.5


Clay, unsaturated 0.1-0.3
Sandy clay 0.2-0.3
Silt 0.3-0.35
Sand
dense 0.2-0.4
loose to medium dense
(void ratio 0.4 - 0.7)
coarse 0.15
fine 0.25
0o1“0c4

Loess 0.1 - 0.3

COEFFICIENT OF SUBGRADE REACTION


In the design of footings and wall foundations, the simplified
concept of subgrade reaction is often used to determine founda-
tion settlement. This concept is based on the assumption that
the settlement, p, of any element of a loaded area is entirely
independent of the load on the adjoining elements. It is
further assumed that the ratio

KS=g'

between the foundation pressure q on the element and the


corresponding settlement p is a constant, K5. The foundation
pressure, q, is called the subgrade reaction and the coefficient
K5 is known as the coefficient of subgrade reaction.
Representative values of K5 for foundation design are given in
Table 7. Allowance should be made for errors in this
approximation by applying a suitable factor of safety, see
Terzaghi and Peck (1967).
29

TABLE 7: COEFFICIENT OF SUBGRADE REACTION (VERTICAL)

Ks
Soil Type (kPa/mm)

Dense gravel and gravelly soils


(no clay fines) >80
Dense sand and sandy soils including
clayey sand, clayey gravel 55-80
Silts, clays of low compressibility 25-55

Clays of high compressibility 15-25

Note: For clays K5 may be assumed to vary linearly


with the unconfined compressive strength qu,
from 8 kPa/mm for qu of 100 kPa to 90 kPa/mm
for qu of 380 kPa.

2 10 SNELLING AND SOFTENING OF CLAYS

Some clays, particularly those with high plasticity


(plasticity index exceeding 20) tend to expand in the
presence of water and if restrained by a structure can
develop very high earth pressures exceeding 500 kPa. These
pressures are not related to soil strength, but to the
mineralogy and moisture content of the clay. Swelling
pressures can be estimated from laboratory swell tests, but
at -present such predictions may not be reliable. These
pressures usually only develop in the zone of weathering at
a depth of up to 1 to 1.5 m below ground level. The above
pressures should be considered if cohesive soil is to be
used behind ‘non-yielding‘ walls, but need not be allowed
for in the case of free standing walls where a small yield
can be tolerated.

When a natural deposit of clay or silt is disturbed by an


excavation for a retaining wall the change in stress
conditions and water content may lead to a change in
shearing strength with time. With stiff fissured clays it
has been shown that progressive softening can reduce the
shearing strength to a small fraction of its original value.
This is usually due to water percolating into the fissures
which are open at the time of excavation for the wall
resulting in a reduction in the value of cohesion c'. Earth
pressures should therefore be calculated using a 'softened'
strength to allow for this deterioration. For details. see
Chandler and Skempton (1974), and Cullen and Donald (1971).
30

In fissured clays and clay fill the rate of softening is reduced


by adequate drainage and when the wall is prevented from
yielding progressively. However the latter requirement will
mean that lateral earth pressures higher than active will
result.
2.11 PERMEABILITY
The permeabilities of soils in broad terms are given in Table 8.
The effects of seepage pressures and permeability of the
backfill material is detailed in section 6.

TABLE 8: PERMEABILITIES OF SOILS


[Values Taken from Terzaghi and Peck (1967)]

. Coefficient of Permeability
Soil Type k(m/S)

Clean gravel 10" - 1


Clean sands, clean sand and gravel 10" - 10"
mixtures
Very fine sands, organic and inorganic
silts, mixture of sandy silt and clay, 10" - 10"
glacial till, stratified clay deposits,
etc
Homogeneous clays below zone of
weathering 10"‘ - 10"’

2.12 LIQUEFACTION

Liquefaction is the process'which causes saturated cohesionless


soils to lose strength or stiffness during earthquake ground
motion. The process is associated with densification of soil
grains, with a corresponding build-up in pore water pressure and
hence a reduction in effective stress and shear strength.
Liquefaction of saturated backfill material and/or foundation
soils has been responsible for a large number of documented wall
failures (e.g. Quay wall failure in Puerto Montt during 1960
Chilean earthquake). A
The liquefaction potential of a wall site and the backfill
material used behind the wall must be considered. In the
extreme case liquefaction may lead to an increase in lateral
soil pressures acting on the wall or a decrease in those
resisting failure.
31

The potential for liquefaction in the backfill material is


reduced by providing drainage and using a free draining gravel
sized material which is well compacted in place.
Where site conditions indicate that liquefaction is possible,
steps should be taken to prevent this occurring since it is not
likely to be feasible to design a retaining wall for this
condition. Liquefaction potential may be reduced on site by
either providing drainage or through densification of the
deposit. Removal and replacement of localised deposits may also
be considered.
For walls of low importance the following method of assessing
liquefaction potential is recommended.
Saturated sandy soil layers which are within 9 m of the ground
surface, have a standard penetration test N-value less than 10,
have a coefficient of uniformity less than 6 and also have a
D20-value between 0.04 mm and 0.5 mm, have a high potential for
liquefaction during earthquakes. Saturated sandy soil layers
which have a D20-value between 0.004 m and 0.04 mm or between
0.5 mm and 1.2 mm have a potential for liquefaction during
earthquakes. Very soft/loose and sensitive silts can also
liquefy.
Soils outside this range of sizes or layers deeper than 15 m are
less likely to liquefy. A more rigorous analysis is recommended
for important retaining walls, see National Research Council
(1985).
Section 3

STATIC EARTH PRESSURE


32

SECTION 3 - STATIC EARTH PRESSURE

STATES OF STRESS
The stresses at any point within a soil mass may be represented
on the Mohr coordinate system in terms of shear stress, t, and
effective normal stress, 0', see Scott (1963), Lambe and Whitman
(1979) or Henry (1986), for the plotting of stresses and use of
the system). With this system, the shearing strength of the
soil at various effective normal stresses gives an envelope of
the possible combinations of shear and normal stress. When the
maximum shearing strength is fully mobilised along a surface
within a soil mass, a failure condition known as a state of
plastic (or limiting) equilibrium is reached.
Where the combinations of shear and normal stress within a soil
mass all lie below the limiting envelope, the soil can be
considered to be in a state of elastic equilibrium, see Terzaghi
and Peck (1967). A special condition of elastic equilibrium is
the ‘at-rest‘ state, where the soil is prevented from expanding
or compressing laterally with changes in the vertical stress.
Any lateral strain in the soil alters its horizontal stress
condition. Depending on the strain involved, the final
horizontal stress can lie anywhere between two limiting
(failure) conditions, known as the active and passive failure
states.
AMOUNT AND TYPE OF WALL MOVEMENT
The earth pressure which acts on an earth retaining structure is
strongly dependent on the lateral deformations which occur in
the soil. Hence, unless the deformation conditions can be
estimated with reasonable accuracy, rational prediction of the
magnitude and distribution of earth pressure in the structure is
not possible.
For no movement of a retaining wall system, at-rest earth
pressures (or pressures due to compaction) act on the wall.
When a wall moves outward, the shearing strength of the retained
soil resists the corresponding outward movement of the soil and
reduces the earth pressures on the wall. The earth pressure
calculated for the active state is the absolute minimum value.
When the wall movement is towards the retained soil the shearing
strength of the soil resists the corresponding soil movement and
increases the earth pressure on the wall. The earth pressure
(or resistance) calculated for the passive state is the maximum
value that can be developed.
33

TABLE 9: MOVEMENT OF HALL NECESSARY TO


PRODUCE ACTIVE PRESSURES

SOII Wall Yield

Cohesionless, dense 0.001


Cohesionless, loose 0.001 - I\> I
Clay, firm 0.01 -
Clay, soft 0.02 - CDGQ II QGCDI U‘lI‘\)@ II

where H is the height of the wall

The amount of movement required to produce the active or passive


states in the soil is dependent mainly on the type of backfill
material. Table 9 gives the outward movement of a wall which is
necessary to produce an active state of stress in the retained
soil. The movements required to produce full passive resistance
are considerably larger, especially in cohesionless material
with movements of 0.05H to 0.1H being indicated (Wu, 1975),
where H is the height of the wall. These requirements apply
whether the movement is a lateral translation of the whole wall
or a rotation about the base. The pressure distributions for
full active and passive states are basically triangular for
constantly sloping ground, see section 3.3.
If a wall rotates about its top in the direction away from the
soil, the soil between the wall and the surface of sliding does
not all pass into the active state. The soil near the top of
the wall stays near the at-rest state. This condition arises in
cuts that are braced as excavation proceeds downwards from the
top. The distribution of pressure may be represented by a
trapezium with dimensions which vary according to the soil type.
The amount of wall movement which will take place depends mainly
upon the foundation conditions and the flexibility of the wall.
The designer must ensure that the calculated earth pressures
correspond to the available wall movement. A free standing wall
need only be designed for active earth pressure as far as
stability is concerned since, if it starts to slide or overturn
under higher pressures, the movement will be sufficient to
reduce the pressures to active. However, if it is on a strong
foundation or otherwise fixed so that adequate stability is
provided, the stem may be subject to pressures near those for
the at-rest state. The following pressure coefficients should
be used for rigid foundation conditions unless a more exact
analysis of movements is made:
(a) Counterfort or gravity type walls founded on rock
or piles KO
34

(b) Cantilever walls less than 5 m high founded on


rock or piles 0.5(Ko+KA)
(c) Any wall on soil foundations or cantilever walls
higher than 5 m KA
In situations where soil-structure interaction is significant
(eg bridge abutments) a rigorous analysis of earth pressure
should be carried out by an experienced engineer. Broms and
Ingleson (1971), and Mathewson et al (1980), describe some
solutions for static earth pressures at bridge abutments.
The tilting movement that will result when earth pressures act
on a retaining wall may be estimated by simulating the founda-
tion soil as a series of springs with an appropriate coefficient
of subgrade reaction, K5, see section 2.9. The base rotation,
9b (in radians), is then given by:

ob = 12Ve/KSLB3 for e 5 g
Where V is the vertical component of resultant of loading on
the base
e is the eccentricity of the load on the base
L,B are the length and breadth of the base respectively
3.3 LIMITING EQUILIBRIUM CONDITIONS
3 The Rankine Earth Pressure Theory
Rankine's equations give the earth pressure on a smooth vertical
plane, which is sometimes called the virtual back of the wall.
The earth pressure on the vertical plane acts in a direction
parallel to the ground surface and is directly proportional to
the vertical distance below the ground surface (ie, a triangular
pressure distribution with the resultant acting at 1/3 H).
Rankine's method is theoretically only applicable to retaining
walls when the wall does not interfere with the formation of any
part of the failure wedges that form on either side of the
vertical plane as shown in Figure 2, or where an imposed
boundary produces the conditions of stress that would exist in
the uninterrupted soil wedges. The method assumes that the
earth pressure acts parallel to the backfill slope, w.
Rankine's equations for earth pressures for cohesionless soils
are given in Figures 2 and 3 and for cohesive soils in Figure 4.
Where there is submerged backfill behind the wall or the
possibility of build up of groundwater level, then the
hydrostatic water pressures should be added to the earth
pressures derived from Rankine's equations, see section 6.2 for
details.
Care should be taken in assessing the earth pressure coefficient
in cohesive soils, see section 3.3.4.
35

Rankine's active earth pressure coefficients are presented


graphically in Figure 3.
Passive, earth pressure calculations using Rankine are not
recommended, since the direction of wall friction will be
incorrect and an underestimation of passive resistance will
result.

3.3.2 Coulomb Earth Pressure Theory


Coulomb theory assumes that a wedge of soil bounded by a planar
failure surface slides on the back of the wall. Hence shearing
resistance is mobilised on both the back of the wall and the
failure surface. The resultant pressure can be calculated
directly for a range of wall frictions, slopes of wall and
backfill slopes.
Where the wall friction is at angles other than the backfill
slope angle the equations given in Figure 5 are an approximation
due to the curved nature of the actual failure surface and the
fact that static equilibrium is not always satisfied. The error
is only slightly on the unsafe side for the active case, but is
more unsafe for the passive case when 8 > ¢/3. Hence the charts
for Coulomb active pressure coefficients given in Figures 6 to 9
may be used at all times, but the passive pressures from Figure
5 should only be used when 8 < ¢/3 or when wall friction is
ignored (conservative).
In the active case the soil can slip down along the back of the
wall, causing the resultant earth pressure to be inclined at a
positive angle, 8, to the normal of the wall, see Figure 5.
It is recommended that an angle of wall friction, 6, of +2/3 ¢
be used in the equation for active pressure for concrete walls
which have been cast against formwork.
However, wall friction may not always result in an increase in
wall stability, see section 2.7 for a detailed discussion of
wall friction.
The inclination of the failure surface assumed by the Coulomb
theory is given by the charts in Figure 10.
Coulomb wedge theory may be used for either the long or short
term conditions of drained or undrained soil respectively (see
section 2.5). For drained soils the effective strength para-
meter ¢', should be substituted for the total strength parameter
¢, in the Coulomb equations given in Figure 5.
For undrained soils analysis in terms of total stress, the use
of saturated unit weight Ysat allows for any pore water
pressures. Hence pore water pressure need not be considered
separately.
Linear interpolation may be used to find the earth pressure
coefficient or failure plane angle for intermediate values of ¢.
36

3.3.3 The Trial Wedge Method


Where the ground surface is irregular or where it is constantly
sloping in cohesive soil, a graphical procedure using the
assumption of planar failure surfaces is the simplest approach.
This procedure is known as the trial wedge method, see Figures
11 to 14.
The backfill is divided into wedges by selecting planes through
the heel of the wall. The forces acting on each of these wedges
are combined in a force polygon so that the magnitude of the
resultant earth pressure can be obtained. A force polygon is
constructed even though the forces acting on the wedge are often
not in moment equilibrium. This method is therefore an
approximation with the same assumptions as the equations for
Coulomb's conditions and, for a ground surface with a constant
slope, will give the same result. If the conditions are the
same as those for Rankine's equations, the trial wedge earth
pressures will correspond to these also. The limitations on
wall friction and passive pressures mentioned in the use of the
Rankine and Coulomb equations also apply to the trial wedge
method. The adhesion of the soil to the back of the wall in
cohesive soils is neglected since it increases the tension crack
depth and hence reduces the active pressure.
For the active case the maximum value of the earth pressure
calculated for the various wedges is required. This is obtained
by interpolating between the calculated values. For the passive
case the required minimum value is similarly obtained.
The direction of the resultant earth pressure in the force
polygons should be obtained from the considerations in sections
3.3.1 to 3.3.3. For the cases where this force is assumed to
act parallel to the ground surface, a substitute constant slope
should be used where necessary, as shown in diagram (A) of
Figure 15 for soil both with and without cohesion.

For cohesionless material, Culmann's graphical construction


shown on Figure 12 provides a compact method of plotting the
resultant earth pressures for the various wedges and obtaining
the maximum value with the corresponding failure plane.
For an irregular ground surface the pressure distribution is not
triangular. However, if the ground does not depart
significantly from a plane surface, a linear pressure
distribution may be assumed, and the construction given in
Figure 16 used. A more accurate method is given in Figure 17.
The latter should be used when there are abrupt changes in the
ground surface or there are non-uniform surcharges.
3.3.4 Earth Pressure in Soils with Cohesion
In cohesive soils tension cracks can exist to a depth 20 (see
Figure 4) where,
37

20 = 25' tan (45° + Q‘)


Y 2
In theory this could cause a reduction in earth pressures on the
back of the wall. In practice the open cracks could fill with
water and increase the hydrostatic pressure on the wall.
In addition cohesive soils can have a high clay content, and
some clay minerals may swell resulting in increased earth
pressures.
For these reasons it is recommended that reductions in earth
pressure are ignored when assessing the long-term stability of
walls, unless a more rigorous analysis is used.
3.3.5 Passive Earth Pressures
The shape of the failure surface for passive failure is strongly
curved when wall friction is present.
Rankine's method gives an underestimation of the passive
resistance because it assumes a positive angle of wall friction
equal to the ground slope, when in fact the wall friction angle
would be negative. (Angle of wall friction is measured as
positive in an anti-clockwise direction from the normal to the
wall).
Coulomb's and the trial wedge methods allow the use of the
correct direction and magnitude for the wall friction angle but
assume a planar failure surface. Hence when wall friction is
high (8 > ¢'/3) these methods lead to an overestimation of the
passive resistance (unsafe). This is accentuated further if the
back of the wall has a negative slope. These methods should not
be used when 8 > ¢'/3 and care should be taken to ensure that 8
is not overestimated as the error is on the unsafe side.
Methods using curved failure surfaces, such as log-spiral and
circular, may be used without introduction of significant error.
Caquot and Kerisel (1948), have presented charts for simple
geometries based on a combination of log-spiral and a plane, see
Figures 18 and 19. For more complex geometries, passive
pressure may be calculated using the circular arc method
outlined in the Hong Kong Geoguide 1 (Geotechnical Control
Office, 1982). This method is quite laborious for even
relatively simple conditions.
3.3.6 Influence of Geometrical Shape of the Retaining Structure on
Wall Friction
The geometrical shape of the retaining structure largely
determines which of Rankine's or Coulomb's conditions are
satisfied or most nearly satisfied for a particular soil, see
(B) of Figure 15. Generally Rankine's conditions may be taken
as applying to cantilever and counterfort retaining walls with
heel lengths equal to at least half the wall height. The earth
38

pressure is calculated on the vertical plane through the rear of


the heel which is sometimes referred to as the virtual back of
the wall.

Coulomb's conditions may be applied to gravity type walls and


walls with small heels, since it will usually be found that the
soil slides on the back of the wall.
For further information on the application of Rankine's or
Coulomb's conditions, see Terzaghi and Peck (1967) or Henry
(1986).
3.3.7 Limited Backfill
The limiting equlibrium methods given above assume that the soil
is homogeneous for a sufficient distance behind the wall to
enable an inner failure surface to form in the position where
static equilibrium is satisfied. Where an excavation is made to
accommodate the wall, the undisturbed material may have a
different strength from that of the backfill. If the trial
wedge method is used, the position of two failure planes should
be calculated - one using the properties of the backfill
material and one using the properties of the undisturbed
material. If both fall within the physical limit of the
backfill the critical failure plane is obviously the one
calculated using the backfill properties. Similarly if they
both come within the undisturbed material, the critical one is
that for the undisturbed material properties. Two other
possible situations may however arise - one where critical
failure planes occur in both materials (the one giving the
maximum earth pressure is used), and the other where the failure
plane calculated with the backfill properties would fall within
the undisturbed material and the failure plane for undisturbed
material would fall within the backfill. In the latter case,
which occurs when the undisturbed material has a high strength,
the backfill may be assumed to slide on the physical boundary
between the two materials. The earth pressure equations do not
apply in this case, but the trial wedge method may be used with
the already selected critical failure plane and the backfill
soil properties. The total pressure thus calculated will be
less than the full active value. However the variation of
pressure with depth is not linear, and should be determined by
the procedure given in Figure 17.
The boundary between the two materials should be constructed so
that there is no inherent loss of friction (or cohesion) on the
failure surface. Benching the undisturbed material will ensure
that the failure surface is almost entirely through solid
backfill material.
39

3.4 ELASTIC EQUILIBRIUM CONDITIONS


3.4.1 At-Rest Pressures
The special state of elastic equilibrium known as the at-rest
state is useful as a reference point for calculation of earth
pressures where only small wall movements occur. For the case
of a vertical wall and a horizontal ground surface the
coefficient of at-rest earth pressure, K0, for normally
consolidated materials, may be taken as:
KO = 1 - sin ¢'
where ¢' is the angle of shearing resistance in terms of
effective stress.
This assumes that the material does not have any built-in over-
consolidation stress. For the case of a vertical wall and
uniform backfill with a slope w, the Danish Geotechnical
Institute recommends using K0 (1 + sin w) for the ‘at rest‘
earth pressure coefficient. For other wall angles and backfill
slopes, it may be assumed that K0 varies proportional to KA, the
coefficient of active earth pressure. At-rest earth pressures
may be assumed to increase linearly with depth from zero at the
ground surface for all materials.
The calculation of wall pressure is carried out in the manner
previously described for active and passive pressures.
For gravity type retaining walls the at-rest pressure should be
taken as acting normal to the back of the wall (ie, 6 = o). For
cantilever and counterfort walls it should be calculated on the
vertical plane through the rear of the heel and taken as acting
parallel with the ground surface.
In cohesionless soils, full at-rest pressures will occur only
with the most rigidly supported walls. In highly plastic clays,
pressures approach at-rest may develop unless wall movement can
continue with time.
3.4.2 Over-Consolidation Pressures
Several factors produce a coefficient greater than that given in
section 3.4.1 above. If a braced excavation is constructed in
over-consolidated clay, the built-in over-consolidation produces
lateral pressures in excess of those that would be obtained by
using the existing depth of material. This is particularly
marked at shallow depths below present ground surface, and is
dependent on the degree of over-consolidation. If some wall
movement takes place these high pressures drop rapidly.
Compaction of backfill in a confined wedge behind a restrained
wall also tends to increase lateral pressures. This is a form
of over-consolidation and is discussed in section 3.4.4.
40

3.4.3 Elastic Theory Methods


When the solution of a lateral earth pressure problem requires
the estimate of some deformations or the relationship between
load and deformation, elastic methods of analysis may be
considered. Usually only linear theory is used. Particular
care and judgement is required in order to select appropriate
elastic constants and boundary conditions. One currently
available general computer program using ,the finite element
method of analysis is ICES-STRUDL.
From elastic theory the coefficient of at-rest pressure for a
vertical wall and horizontal ground surface is given by:

KO = 1%; (for plain strain)

Elastic theory may be used to predict the increase in lateral


loading resulting from a vertical load or surcharge, see section
5, or the increase in vertical stress resulting from an increase
in lateral loading, see Clayton and Milititski (1985).
3C4\4 Compaction Induced Earth Pressures
Proper compaction of backfill behind a retaining wall is
necessary in order to prevent excessive settlement and increase
the shearing strength of the backfill. Care should be taken to
ensure that the compaction process does not damage the wall, as
compaction induced pressures vary considerably in magnitude and
distribution and can be much larger than those predicted using
classical earth pressure theories.
Compaction induced earth pressures are dependent on wall
movements. Translations or rotations in the order of H/500,
where H is the height of the wall, may reduce compaction induced
pressures to those of the active state. Therefore free standing
relatively flexible walls such as crib walls or cantilever walls
on soil foundations need only be designed for active pressures.
Free standing flexible walls on rock foundations, or in situa-
tions where structural deformations are limited by compatibility
with other structures or when the risk of structural damage is
unacceptable, may require further consideration. In these cases
wall movement is sufficient to reduce the compaction induced
earth pressure to the active state but a parabolic rather than
triangular earth pressure distribution may result. The centre
of earth pressures of a parabolic distribution may be as high as
H/2 with a corresponding increase in bending moment of 50%.
Therefore for these cases a reasonably conservative basis for
derivation of bending moments, bearing pressures and overturning
stability may be warranted.
The factor of safety against sliding may be based on active
earth pressure.
41

Restrained rigid walls such as foundation walls or counterfort


and gravity walls on rock or pile foundations need to be
designed for earth pressures from compaction induced earth
pressure theory.

Aggour and Brown (1974), give guidance on the formulation of


numerical solutions to compaction problems and include in their
paper graphical solutions which indicate the influence of some
factors affecting residual pressures, eg, backfill geometry,
wall flexibility, end wall restraint.
Broms (1971), has presented a method for the determination of
lateral earth pressures due to compaction against unyielding
structures and proposes the earth pressure distribution shown in
Figure 20(A) for use in design. The associated data relating to
the figure are for a limited range of compaction plant.

Ingold (1979) has presented a simple analytical method which can


be used to give a working approximation of compaction induced
pressures for routine designs. The method results in the
following earth pressure distribution:
(a) Below a critical depth zc there is no reduction in
horizontal stress after removal of compactive force. Ingold
shows that approximate values of zc may be obtained from the
following expression:

zc = KA
/2.2
nv
Where KA is the coefficient of active earth pressure,
Y is the bulk unit weight,
p is the effective surface line loading imposed by
the compactor
(b) The depth, hc, below which active pressure due to the weight
of the overlying soil exceeds the compaction induced
pressure is obtained from:

hc =l_ /ZE
KA HY

The resulting effect of compaction on lateral pressure is shown


in diagrams (a) and (b) of Figure 20(B), and the resulting
pressure distribution for use in design, based on this
simplified theory, is shown in diagram (c) of Figure 20(B).
Ingold's design pressure distribution can be seen to be very
similar to that of Broms shown in Figure 20(A).
Section 4

EARTHQUAKE EARTH PRESSURES

AND DESIGN
42

SECTION 4 - EARTHQUAKE EARTH PRESSURES AND DESIGN

4.1 BACKGROUND
1 Wall Categories
The behaviour of wall structures during earthquakes can be
broadly classified into three categories related to the maximum
strain condition that develops in the soil near the wall. The
soil may remain essentially elastic, respond in a significantly
nonlinear manner or become fully plastic. The rigidity of the
wall and its foundations will have a strong influence on the
type of soil condition that develops.
Flexible structures, such as cantilever walls, displace
sufficiently under gravity backfill loads to produce a fully
plastic strain condition in most soils.
For rigid walls, such as gravity walls, basement walls, and
other walls on rigid foundations, including piles or rock, the
soil behaviour may be essentially elastic under combined
earthquake and gravity loads.
1 Soil-Structure Interaction
Basement walls in buildings and abutment walls that are
monolithic or rigidly connected to bridge superstructures are
often subjected to displacements relative to the soil mass
because of the dynamic displacement response of the structure
during an earthquake. These types of walls may be subjected to
a complex interaction of dynamic soil pressures arising from
both the displacement response of the structure and earthquake
elastic waves in the soil.
It is usual to simplify the complex problem of interaction of
earthquake elastic waves with wall structures by assuming that
the earthquake ground motions are equivalent to dynamic inertia
forces acting in the backfill mass. Dynamic pressures on the
wall can be estimated by analysing the wall and backfill
modelled as an elastic continuum or failure wedge and subjected
to gravity and horizontal body forces.
The dynamic displacement of basement walls in tall buildings is
likely to be dominated by the rnovements of the building but
often basement structures are very rigid and wall displacements
small. Where the backfill is firm soil, small movements of the
wall in a direction towards the retained soil can result in
significant increases in pressures.
43

1 Simplifications for Design


Although mathematical modelling techniques, such as the finite
element method are available to investigate the interaction of
soils and wall structures under seismic loading, because of the
uncertainty existing in many of the input parameters it is
unlikely that precise estimates of the wall response and earth
pressures can be obtained.

Only in rather exceptional circumstances will soils


investigation information be available to provide a good
prediction of the soil behaviour under dynamic loading. The
character of the earthquake motion is often only known in terms
of generalised response spectra, and it is difficult to develop
detailed information on the spatial variation in the ground
displacements and how the ground shaking is modified by the
foundation soils and by interaction with the wall. Thus, unless
the wall is of unusual importance, it is acceptable for design
purposes to use dynamic pressures calculated on the basis of
simplified wall geometries and to use ground accelerations
defined by response spectra and assumed to be uniform within the
extent of the surrounding soil mass. However, it is important
that the simplified methods make valid assumptions regarding the
maximum strains likely to develop in the surrounding soil and
proper account is made of the influence of wall deformations on
pressures.
1 Plasticity Theory and Failure Modes
The Mononobe-Okabe (M0) method has been widely used for
predicting earthquake forces on walls (Seed and Whitman, 1970).
The method is based on an approximate plasticity theory and is
essentially an extension of the well known Coulomb sliding wedge
method for estimating static pressures. The method is based on
the assumption that the wall displacements are sufficiently
large to produce a fully plastic stress state in the soil by
either outward movement (active state), or by movement towards
the backfill (passive state). The assumptions made regarding
wall displacements are frequently satisfactory for free standing
walls but are not appropriate for many types of walls associated
with other structures such as buildings and bridge abutments
where the soil may remain essentially elastic.
It is important that the failure mode and capacity design
principles be considered in the earthquake design of walls. For
walls that are relatively rigid, the initial earthquake
accelerations may induce pressures, corresponding to elastic
soil behaviour, that are significantly greater than the M0
pressures. If the wall has insufficient strength to resist
these elastic pressures, then yielding and outward movement of
the wall may occur with a fully plastic stress condition
developing in the soil. With outward yielding and the onset of
the plastic stress condition, the pressures on the wall will
drop to the M0 values. The outward yielding may result from
either permanent displacements in the soil foundation or from
44

yielding in the structural wall elements. The failure mode will


depend on the wall configuration and the relative capacities
available in each of the potential failure mechanisms.
Although it is common procedure to design building frames for
inelastic behaviour under earthquake loads, it may be
undesirable to design wall structural components for yielding.
In a wall structure, owing to the presence of lateral gravity
pressures, yielding will tend to occur only in a direction away
from the retained soil. In major earthquakes this may result in
large permanent deflections and cracking with loss of
serviceability.

If structural damage is to be avoided, it is necessary to design


for either the maximum peak earthquake-induced pressures
consistent with the type :of soil behaviour expected, or to
detail the wall to displace outwards, where this is possible, by
movement on failure surfaces in the soil (sliding or tilting).
Where soil failure modes are possible, the sliding block analogy
of Newmark (Elms and Richards, 1979), may be used to determine
the approximate magnitude of the outward movement as a function
of the peak design acceleration and the acceleration level
required to initiate failure. Theoretical studies and lnodel
tests on shaking tables have shown that for many types of walls,
it is possible to design for significantly less than the
expected peak ground acceleration without exceeding acceptable
limits of outward movement.
1 Resonance Effects
Most wall/soil systems are relatively rigid and have fundamental
periods of vibration less than 0.5 s. It is usual to design
walls for earthquake peak ground accelerations (or lower than
peak values) on the assumption of relatively rigid behaviour and
low periods of vibration. However, typical earthquake response
spectra are very steep at low periods and any flexibility may
lead to ground motion amplification, particularly close to the
top of the wall. Although amplifications have in fact been
noted in model studies, the results of these investigations have
probably been affected by the presence of rigid boundaries that
reflect and contain the vibrational energy within the model
wall/soil system. In full scale wall structures, there are
seldom boundary effects and the damping will be higher because
of energy losses by elastic wave radiation. Also, where soil
failures develop, it is likely that the soil damping may be
higher than commonly assumed. Thus, it is difficult to draw
firm conclusions on whether resonance effects should be
considered or whether it is satisfactory to design to peak
ground acceleration levels. It is therefore recommended that
resonance effects be neglected in design unless special studies
are undertaken for important walls.
45

DESIGN SEISMIC COEFFICIENTS


For free standing walls it may be assumed that the wall/soil
system has a short fundamental period of vibration and that the
inertia loads can be approximated by using the zero period
(T = 0) ordinate [C(O)] on the design response spectrum (or peak
ground acceleration). Where walls form integral parts of other
structures, such as bridges and buildings, the appropriate
design coefficient for estimating the wall displacements should
be obtained from the periods of vibration of the structure and
the design response spectrum.
Seismic coefficients based on the response spectra given in
DZ 4203: 1989, (Standards Association of New Zealand, 1989), are
recommended for wall design. The horizontal earthquake
coefficient at period, T, in DZ 4203 is C(T) and is given by:
C(T) = co RZ
Where Co is the response function for a 150 year return period
earthquake and has a value of 0.4 for zero period.
Thus C(O) = 0.4 RZ
Z is a zone factor given in Figure 21.
R is a risk factor which may vary from 1.3 to 0.8 as
defined below.
R = 1.0

Major retaining walls supporting important


structures, developed property or services and
where failure would have serious consequences such
as cutting vital comunication services and loss of
life. Walls forming part of the earthquake
resisting structure of bridges, major buildings or
other important structures.
R = 0.8

Walls other than as described for R = 1.0 with


heights greater than 4 m for level backfills, or
3 m with significant backfill slope.
A risk factor greater than 1.0, as defined in DZ 4203 Table
3.2.1, may be used for walls that form part of the earthquake
resisting structure of buildings classified in Categories I to
III.
The Bridge Manual (National Roads Board, 1989) contains similar
seismic loading provisions.
The risk factor for highway bridge abutments should be the same
factor used for the design of the bridge. For important
bridges, risk factors greater than 1.0 may be specified in
design briefs adopted by roading authorities.
46

Walls not included in the above descriptions need not be


specifically designed for earthquake loading.
The DZ 4203 seismic design coefficients are based on a 150 year
return period event. A reduction of the risk factor to 0.8
effectively reduces the design return period to about 100 years.
LOAD COMBINATIONS
Under normal circumstances, when live loads, such as traffic,
are of a transient nature, only the combination of earthquake
pressures with static gravity pressures should be considered.
The static gravity pressure should include water pressures and
surcharge loading, see section 1.4.1.
FACTORS OF SAFETY
Where the design approach is to prevent outward movements that
may develop because of failure in the soil or yielding of the
structure, the following factors of safety for the load
combination of gravity plus earthquake pressures should be used:
Safety factor against overturning: F5 = 1.5
(or gross rotational failure)
Safety factor against sliding: F5 = 1.2
Where outward movements are to be permitted, an outward sliding
mechanism is usually preferred to a rotational failure. It
would then be appropriate to prevent a rotational failure by
using a factor of safety of at least 1.2 against overturning or
gross rotational failure.
The combined gravity and earthquake induced forces to be
considered in a wall stability analysis are shown in Figure 22.
Resisting forces from passive pressure and base friction should
be calculated using the usual assumptions made for the soil
strength parameters with gravity forces acting alone.
Resistance to overturning is usually related to the allowable
bearing pressures under the toe of the wall. Inertia forces
acting on the wall and any soil mass in contact with the wall
but not included in the sliding wedge mass should be considered
in the analysis. The resistance of the passive wedge may be
based on the MO solution for passive failure (see section 4.5.4
and Figure 41).
DYNAMIC FORCES AND PRESSURE DISTRIBUTIONS
Earthquake pressures resulting from both soil inertia loads and
dynamic wall displacements are given in this section for a
number of wall categories defined in terms of wall stiffness and
displacement relative to the soil mass. The selection of the
appropriate pressures for particular wall construction types and
foundation conditions is discussed in section 4.6.
47

4.5.1 Rigid Wall

The earthquake pressures on a smooth perfectly rigid wall from


horizontal inertia loads in the soil are shown in diagram (A) of
Figure 23 (Wood, 1973). An approximate linear pressure
distribution suitable for design purposes is given in diagram
(B) of Figure 23 (Matthewson et al, 1980). The increment of
earthquake force is given approximately by:
APE = c(o) v H’
The point of application of the earthquake force increment is at
approximately 0.6 H above the base (Wood, 1973).
The earthquake induced pressures and forces are dependent on the
soil elastic properties but are not very sensitive over the
normal range of properties for typical soils, see Figure 23.
Therefore, for design purposes, the earthquake pressure
distribution and force on a rigid wall can be assumed to be the
same for all soils including soils with both cohesion and
friction.
For the case of sloping backfill, with inclination angles less
than 20°, the earthquake pressures and forces may be obtained by
using an equivalent wall height as shown in Figure 24. At
backfill slopes greater than 20°, a more detailed analyses is
required. Slope stability may also be critical for this case.
4.5.2 Stiff Wall
A stiff wall is defined here as a wall that moves outward at the
top by 0 to 0.2% H under combined gravity and earthquake
pressures. An approximation for the increment of earthquake
pressure on a wall that displaces 0.2% H at the top is shown in
diagram (A) of Figure 25. The increment of earthquake force for
0.2% top displacement may be taken as:
APE = 0.75 c(o) v H’
The point of application of the earthquake force may be taken as
0.5 H from the base.
The pressure distribution shown in Figure 25 is based on the
analytical work of Wood (1973) and experimental results from
Stevenson (1987).

Pressures and forces on walls that displace less than 0.2% at


the top may be obtained by linear interpolation between the
stiff and rigid wall pressures and forces.
The earthquake pressures on a stiff wall are more sensitive to
the soil stiffness properties than for the rigid wall case.
However, estimating the effects of soil stiffness adds
considerable complexity and it is recomended that these effects
be neglected for design purposes.
48

The effect of sloping backfill can be obtained in a similar


manner to the rigid wall case.
4.5.3 Flexible Wall
Where the outward movement of the top of the wall under gravity
and earthquake pressures exceeds 0.5% H, an active pressure
state may be assumed and the pressures obtained from the Coulomb
sliding wedge theory or the Mononobe-Okabe formulae. In these
methods, an additional force equal to the seismic coefficient
times the weight of the soil wedge is included in the analysis.
The M0 equations were originally derived for a cohesionless soil
for both active and passive conditions and for both vertical and
horizontal earthquake forces. Vertical accelerations do not
produce significant increases in the horizontal pressures and
for most walls vertical earthquake effects may be neglected.
The method has been extended to cohesive frictional soils by
Prakash and Saran (1966). The extended method may be used for
the analysis of saturated clays by carrying out a total stress
analysis assuming ¢ = 0 and c = Su, the undrained shear
strength.
The shape assumed for the M0 earthquake pressure increment is
shown in diagram (B) of Figure 25. The M0 method does not give
detailed information on the shape of the increment of earthquake
pressure and there are conflicting results from some of the
model tests carried out on shaking tables and other simplified
plasticity methods (Athanasiou-Grivas, 1978; Prakash and
Nandakumaran, 1979; Saran and Prakash, 1970). However, finite
element studies (Wood, 1973) and more recent tests on flexible
walls (Stevenson, 1987), have shown that providing the wall is
sufficiently flexible for active conditions to develop, the
increment of earthquake force acts at approximately 0.33H above
the base of the wall. This is the assumption made in the
original development of the theory by Mononobe-Okabe (Mononobe
and Matsuo, 1929).

The M0 equations for a cohesionless soil and active pressure


state are given in Figure 26. Figures 27 to 34 give the
earthquake earth pressure increment where the wall is vertical
(or virtual backface vertical) and Figures 35 to 38 give the
coefficients for a wall sloping at 4V in 1H. For the vertical
wall case, coefficients are given for the wall friction 8 = ¢
and 8 = 2¢/3. The sloping wall coefficients are given for the
8 = 2¢/3 case. When the shape of the wall forces the wall slip
plane to be on a soil interface or virtual plane in the soil
behind the wall, it is usual to assume that the wall friction is
equal to the soil friction angle. When a slip surface can
develop on the back face of the wall it may be assumed that the
wall friction is 2/3 of the soil friction.
The earthquake pressure increment curves become infinitely steep
when the seismic coefficient values reach the limit that causes
49

general failure in the backfill or slope behind the wall. For a


cohesionless soil this occurs when:
C(O) 2 tan (¢ - w)
Where w = the slope angle
Where possible, conditions that lead to general soil failure and
high wall forces during earthquakes should be avoided.
The earthquake increment pressure distributions are obtained
from the plotted coefficients by the following expression:
Ap(Z) = AKAE Y Z

where:
Ap(z) = the earthquake pressure increment at depth z below
the top of the wall.

AKAE = KAE - KA A
KAE = M0 active pressure coefficient
(total gravity + earthquake component)
KA = Active earth pressure coefficient
The earthquake increment of wall force is given by:
APAE = 0.5 AKAE Y H2

A useful approximate expression for estimating the earthquake


force increment for cohesionless soil with a horizontal backfill
surface and friction angle of 30° to 35° is:
APAE = 0.5 c(o) v H’
The above expression gives a good approximation when C(O) is
between 0.2 to 0.3.

Where the soil is cohesive or the ground surface is irregular,


the trial wedge method, as shown in Figure 39, can be used to
estimate the earthquake force increment. There is no available
information on the correct shape of the pressure distribution
for these more complex cases. However, it appears reasonable to
assume that the earthquake force increment acts at 0.33 H above
the base. That is, the same assumption used for cohesionless
soil appears appropriate for all soils and backfill slopes.
The presence of tension cracks in cohesive soil may be ignored
since the lateral compression at the ground surface from the
dynamic inertia forces in the soil offsets the tensile stresses
that develop because of outward yielding.
For walls with stiffnesses intermediate between the stiff and
flexible cases, linear interpolation may be used between the
pressures and forces for the two limiting cases.
50

4.5.4 Displaceable Wall


When it is acceptable for a wall to undergo permanent outward
displacement during earthquake loading, it may be designed for a
threshold acceleration N g less than the peak ground
acceleration of the design earthquake, C(0)g. The threshhold
acceleration is defined as the acceleration level that initiates
permanent movement of the wall. The outward displacement can be
calculated from the Newmark sliding block theory (Elms and
Richards, 1979). The forward displacement of the centre of mass
is given approximately by (Matthewson et al, 1980):

= = c(o N -2
d f3I\(8)oLNl+¢(0) :i
Where V0 = Peak ground velocity.
= 1.3 C(O) m/s

Alternative expressions for outward displacement are given by


Franklin and Chang (1977), Elms and Richards (1979) and Nadim
and Whitman (1985). The above expression is in reasonable
agreement with these alternatives and is considered more
suitable for design applications.
The ratio of d/C(O) obtained from the above expression is
plotted in Figure 40 against the ratio of the threshold
acceleration to the peak acceleration, N/C(O). This plot
enables the outward displacement of the centre of mass to be
estimated for any level of peak ground acceleration.
The earthquake pressure increment on the displaceable wall may
be taken as the M0 values given in section 4.5.3 for flexible
walls.
If it is required to prevent yielding in the wall elements,
capacity design principles should be used. In estimating the
overstrength of the soil failure mode, the estimated values of
soil cohesion, c, and friction, tan ¢, should be increased by a
factor of 1.3 (see section 2.5 for discussion on shear strength
parameters).
In general, the inertia force acting on the wall should be
included in the determination of the critical acceleration to
cause failure. This can be done using results given by Elms and
Richards (1979), or by analysing the stability of the wall in
accordance with section 4.4 (Figure 22). For gravity walls the
wall inertia load has a significant influence on the threshold
acceleration.

Where stability is considered using a virtual back face, (for


example at the heel of a cantilever wall), the weight of the
soil between the virtual back face and the wall needs to be
included with the wall weight to estimate the inertia force. In
this case, the additional soil mass may be quite large and the
behaviour will be similar to a gravity wall.
51

For some cases, the weight of the wall may be small in relation
to the weight of the soil wedge and the wall inertia force may
be neglected.

To estimate the threshold acceleration it is usually necessary


to know the passive resistance for any soil in front of the wall
toe. The earthquake increment of passive pressure and the total
passive pressure coefficient for gravity forces plus soil
inertia forces are plotted in charts A and B of Figure 41. Both
plots were evaluated using the MO equations for a cohesionless
soil, vertical wall face and horizontal backfill. The passive
pressure coefficient is very sensitive to the assumption made
regarding the wall friction and the passive pressure
coefficients shown are for a conservative assumption of zero
wall friction. Note that in the passive pressure case, the
earthquake increment reduces the static gravity load passive
resistance when the soil inertia force is assumed to be acting
in a direction away from the wall. (Negative values are plotted
in Chart A of Figure 41 to indicate this reduction).
4.5.5 Forced Wall
Where the wall is part of a larger structure such as a building
or bridge, it may be forced to vibrate with amplitudes governed
by the inertia loads on the structure. The total earthquake
pressure increment can be estimated by combining the component
of earth pressure due to inertia forces in the soil (usually
based on a rigid wall assumption) with pressures resulting from
the wall displacement amplitudes against the soil.
Figure 42 shows the earth pressure components caused by
rotational and translational displacements of the wall against
the backfill soil. The flexibility of the wall will have an
influence on this component of pressure but a reasonable
estimate can be made for most walls by combining the results
from the simple rotational and translational deformations shown
in Figure 42. For more complex wall geometries, an acceptable
estimate for the pressure against a forced wall may be made by
modelling the soil as a system of linear Winkler springs.
An upper limit to the combined static and forced wall pressure
at any depth is given by the soil passive pressure distribution.
4.5.6 Water Pressures
Increases in pore water pressures resulting from earthquake
inertia effects need to be considered in the analysis of wall
pressures. For some backfill and foundation soils it may also
be necessary to allow for pressure increases due to the effects
of liquefaction of the soil, see section 2.12.
The increase in pressure due to inertia loads on any water can
be taken into account by applying the seismic coefficient to the
total weight of soil and water in the failure wedge.
52

A convenient method is to consider the effective soil stresses


on the wall separately from the water pressures. The
coefficient of earthquake pressure increment (AKAE) can be
obtained from the M0 solution by assuming the location of the
failure plane is unaffected by the position of the water table.
The effective stress earth pressure increment can then be
obtained by using the pressure equations with the soil bulk unit
weight above the water table and the submerged unit weight below
the water table. The earthquake increment from the pore water
pressure can be obtained from the static water pressure
multiplied by the seismic coefficient. The total seismic
increment is then the sum of the effective soil stress increment
and the pore water pressure increment as shown in Figure 43.
The hydrodynamic pressure from any water in front of the wall
(e.g. quay walls) may often act in the same direction as the
earth pressure increment and should be considered in both
stability and wall strength analyses. The critical case for
overall stability of the wall will occur when the hydrodynamic
water pressure reduces the static water pressure in front of the
wall and is in phase with the active earth pressure increment on
the wall. It may also be necessary to consider the case when
the hydrodynamic pressure increases the static water pressure
and is in phase with the passive earth pressure increment. That
is, when the inertia loads in the water and soil are directed
towards the backfill. Although it is unlikely that this
direction will be critical, in some circumstances it may be
necessary to consider the loads from this case in the structural
design of the wall.

Hydrodynamic pressures can be estimated using the Westergaard


(1933) theory. From the solution given by Werner and Sundquist
(1949) for a relatively shallow long reservoir, the dynamic
water pressure force is given by:
PW = Yw H2

Where h = depth of water


The dynamic water pressure force acts at a height of about 0.4 h
above the base.
Further information on the effects of the length of the
reservoir and fluid resonance can be obtained from Werner and
Sundquist (1949) and Chopra (1967).
53

4.6 APPLICATION TO VARIOUS TYPES OF WALLS


4\6\1 Free Standing Walls Founded on Soil
Most types of free standing walls founded on soil are
sufficiently flexible for the flexible wall earthquake pressures
to apply. The maximum permissible displacement should usually
be adopted as the prime criterion for earthquake design. The
failure mode should avoid yielding in the structural elements
wherever practicable.
4I602 Free Standing Walls Founded on Rock or Piles
Earthquake pressures will usually be higher than for the
flexible wall. If yielding in the structural members of this
type of wall is to be avoided, earth pressures and wall inertia
forces should be based on the peak ground accelerations and
account should be taken of the wall stiffness in estimating the
earthquake pressure distribution.
Yielding of the structural elements may be permitted when the
loss of serviceability, or the cost of removing the backfill and
repairing the wall can be justified on economic grounds. Where
significant outward displacement occurs because of yielding, the
displaceable wall theory outlined in section 4.5.4 may be used.
4.6.3 Tied Back Walls
The pressures on tied back walls are dependent on the type of
anchor and flexibility of the wall. Because of the complexity
of the interaction between the tie forces and wall facing
deformations, major walls should be investigated by structural
analysis to estimate the pressure distributions.

If tie backs are restrained by some form of deadman anchor and


the ties are required to remain elastic during earthquake
loading, the peak ground acceleration should be used to
calculate the pressures and forces. In estimating the
earthquake increment of pressure from soil inertia forces, due
allowance should be made for the tie and wall flexibility.
Where the wall top movement meets the flexible wall criterion
given in section 4.5.3 the M0 pressures may be used.
If tie backs are restrained by a nmvable anchor, such as a
friction slab designed to slide while the other structural
components remain elastic, a reduction in the design
acceleration may be made on the basis of the displacement
criterion given in section 4.5.4. For walls of minor
importance, permanent displacement resulting from yielding of
the ties may be acceptable but particular consideration needs to
be given to the effectiveness of the tie corrosion protection
system after yield extensions.
For all types of anchor systems the mode of failure, in the
event of overload, should be by yielding of the ties or passive
54

failure of the anchor rather than failure of the wall face or


connections between the ties and the wall face or anchor. The
probable variation in the soil parameters and frictional
resistance between the wall components and the soil should be
considered in determining the permanent displacement threshold
accelerations and failure modes.
When investigating the stability of a tied wall, the forces on
the face and ties may be estimated using the active wedge
failure criterion. The "passive" failure modes of the toe and
the anchor system should also be considered (Anderson et al,
1983). Failure by a wedge through the anchor and toe, as shown
in Figure 44, or by a slip circle, may be possible under
earthquake loading. In these failure modes, the horizontal
earthquake force corresponding to the peak ground acceleration
should be applied to the wedge or weight of soil within the
circular slip.
4.6.4 BOSEIIIGIIIZ Walls

The earthquake pressures that develop on basement walls will


generally consist of components from the inertia forces in the
soil and pressures resulting from the wall displacement against
the soil.
In most cases the pressures from the soil inertia loads may be
taken as the rigid wall pressures given in section 4.5.1, but
allowance may be made for reductions because of wall
flexibility.
The component of earthquake pressure from the movement of the
wall relative to the soil may be estimated from the forced wall
solutions given in section 4.5.5. Where the basement is founded
on rock or very firm soils, the relative movement of the wall
against the soil may be small and the resulting pressure
component small in relation to the soil inertia force increment
for rigid walls. When piles are used or the structure is
founded on soft soils, the relative movements may be quite large
and the resulting pressure components may dominate the total
earthquake pressures on the wall. A limiting value of full
passive pressure for combined gravity and forced wall components
may occur where the basement walls are used to provide lateral
resistance against the earthquake base shear forces of the
structure supported by the basement. On flexible foundations,
the response of the structure may be affected by the stiffness
of the soil surrounding the basement. In this case, it may be
necessary to investigate the response of the building using
Winkler springs to model the soil. Spring stiffnesses may be
estimated from the typical values of coefficient of subgrade
reaction for vertical loads given in section 2.9 and from the
forced wall solutions given in section 4.5.5. The effect of
non-linear soil behaviour on the pressures on translated and
rotated walls can be obtained from the finite element results of
Wood (1985).
55

Since in general the two earthquake components of pressure (soil


inertia and forced wall) will have different vibrational
frequencies, they may be combined using the square root of the
sum of the squares rule (SRSS). Using this method it is readily
shown that if one component is less than 50% of the other,
neglecting the smaller component reduces the total by less than
12%. Thus, it is helpful to make preliminary estimates of the
components and only carry out detailed analyses to estimate the
smaller component when it is estimated to exceed 50% of the
larger component.
4.6.5 Bridge Abutments
Pressures on bridge abutments are influenced by the earthquake
forces and displacements transmitted to the abutment by the
bridge superstructure. The bearing detail between the abutment
and superstructure will influence these forces and
displacements. If for example, the bearing is a sliding type
with a low coefficient of friction, the abutment may act
essentially as a free standing wall. Another limiting case is a
monolithic abutment where the wall is forced with displacements
that may be governed by the response of the total bridge system
to the earthquake inertia forces.
When evaluating earthquake pressures it is helpful to consider
two categories of abutments. The first type is the case where
the soil pressures make no significant change to the dynamic
response of the bridge. The second case is when the soil
pressures on the abutment have a significant influence on the
dynamic response. In this latter case, the abutment will
generally be monolithic or integral with the superstructure.
Procedures for estimating the abutment pressures for the two
cases are given below.
(i) No Significant Interaction
The forces and displacements acting on the abutments in this
category are shown in diagrams A and B of Figure 45. Two
separate cases are shown. Firstly, where the load from the
superstructure is limited by a sliding or a deformable bearing,
the force imposed by the superstructure will be known. The
second case is where there is a rigid connection to the
superstructure and the analysis of the abutment will need to be
based on an imposed displacement rather than the force
transmitted. The forces and displacements shown in Figure 45
are defined as follows:
APE = Earthquake pressure component from inertia forces in
the soil.
PF = Earthquake pressure component from forcing of the
wall against the soil.
P1 = Inertia load acting on the abutment mass.
56

P5 = Gravity pressure component.


PL = Load from superstructure.
Ax = Displacement of superstructure.
The magnitude and direction to be assumed for some of these
forces depends on whether the wall is being displaced against
the backfill or away from the backfill. The movement against
the backfill is usually the critical case for the backwall
design and movement away from the backfill is usually critical
for the design of the abutment foundations. Particular
consideration of the force component directions may be required
for the design of clearances, joints, bearings and linkages.
For movement away from the backfill, APE may be estimated from
the details given in sections 4.5.1 to 4.5.3 making due
allowance for wall flexibility. When the displacement is
against the backfill, rigid wall pressures may be assumed for
estimating APE.
PE may be estimated from the forced wall solutions given in
section 4.5.5. For translational deformation PE is given by:
PE = 0.6 E5 a Ax
Where a = horizontal width of abutment

For the rigid connection case, Ax is known and PE can be


obtained from the above expression. For the case where the load
from the superstructure is known rather than the displacement,
PE cannot be obtained directly and needs to be evaluated by
analysing the abutment, including both the foundation and wall
soil stiffnesses, loaded by PL + P1. The wall stiffness can be
approximated using the above expression or Winkler springs.
PE always acts in the direction opposite to the movement of the
wall. For abutment movement towards the backfill, PE
effectively increases the static gravity pressure and, the
maximum resultant of PE + P5 is limited by the static passive
pressure force. For abutment movement away from the backfill,
PE reduces the static pressure and the minimum resultant of
P5 - PE is the static active pressure force.
APE and P1 should be assumed to be in phase. PL or Ax may or
may not be in phase with APE and P1. Directions of the forces
should be chosen to produce the most critical loading case.
Because the earthquake pressure components APE and PE are caused
by vibrational effects with different frequencies, they may be
combined for abutment movement against the backfill (critical
case for wall design) using the SRSS method. For movement away
from the backfill, the critical loading on the abutment
foundation will probably occur with APE and PE acting in
opposite directions. The correct method of summing forces will
57

depend on the relative magnitude of the forces and also on


whether the connection is load limiting or rigid. In view of
the complexity, for this case the individual force components
should be combined using the algebraic sum.
(ii) Significant Interaction
When the bridge response is significantly influenced by the
interaction with the abutment soil it is difficult to account
for the dynamic effects in a simple analysis procedure. The
critical loading on the abutments and lateral load resisting
elements can usually be obtained by considering the two cases of
in-phase and out-of-phase earthquake soil inertia pressure
components as shown in diagrams (C) and (O) of Figures 45.
When the dynamic components of earth pressure are out of phase
at either abutment (diagram (C) of Figure 45) it may be assumed
that the structure does not move relative to the foundation and
is subjected to rigid wall pressures on each abutment wall.
That is, the total pressures on the walls are the sum of the at-
rest static pressure and the rigid wall earthquake component
from the soil inertia loads. This assumption may overestimate
the earthquake pressure components on very short bridges where
it is unlikely that out-of-phase accelerations will occur.
Because of the influence of the soil properties and the
frequency content of the incoming waves on the phase
relationships, it is difficult to make more precise predictions
of the effect of the bridge length on the pressures.

When the dynamic components of pressure from the soil inertia


loads are in phase at either abutment (diagram (0) of Figure 45)
the bridge will displace relative to the foundation, generating
forced wall dynamic pressures that are dependent on the overall
displacement response of the bridge. The analysis procedures
for this case are similar to those discussed in section 4.6.4
for basement walls.
Peak ground acceleration should be used to calculate the wall
earthquake pressure component from the soil inertia loads.
These components have upper limits of rigid wall values. On the
abutment moving away from the soil, the component may be reduced
by taking into account the wall flexibility. A lower limit for
this component is the MO earthquake active pressure increment.
The dynamic pressure components on the abutments due to the
relative displacement of the bridge can be estimated by
computing the period of vibration taking into account the
abutment soil stiffness. The displacement response can then be
estimated from the design response spectrum and the overall
ductility factor. For relatively rigid structures, a
satisfactory estimate of the displacement can be obtained by
using the peak ground acceleration to obtain the inertia load on
the bridge.

The minimum limiting value of P5 + PE on the abutment moving


away from the soil is the active static pressure. The maximum
58

limiting value of P5 + PE on the abutment moving towards the


soil is the passive pressure. The total forces for these
limiting cases can be obtained from the MO solutions for KAE and
KpE.
Section 5

EFFECT OF SURCHARGES
59

SECTION 5 - THE EFFECT OF SURCHARGES

UNIFORM SURCHARGES
Uniform surcharge loads may be converted to an equivalent height
of fill and the earth pressures calculated for the corres-
pondingly greater height. The equivaient height is given by:
_ g cos B
he _ Y cos(B - W)
The depth of the tension zone in cohesive material calculated
from the top of the equivalent additional fill. The
distribution of pressure for the greater height is determined
from the procedures given in sections 3 and 4. The total
lateral earth pressure is calculated from the pressure diagram
neglecting the part in tension and/or the part in the height of
fill equivalent to the surcharge.
Concrete buildings may be represented as a uniform surcharge of
10 kPa per storey. Timber buildings may be taken as half the
above.
Traffic loading, when at a greater distance than 2/3 times the
height of the wall from the back face of the wall, may be
represented as a uniform surcharge of 12 kPa.
In some cases a surcharge load could aid stability. The effect
of two loading cases on wall stability are shown in Figure 46.
If the surcharge is not permanent, then loading 2 in Figure 46
should be assumed, for design purposes.
LINE LOADS
Where there is a superimposed line load running a considerable
length parallel to the wall, the wedge method of analysis may be
used, and the weight per unit length of this load can be added
to the weight of the particular trial wedge to which it is
applied, see Figure 47. The increased total earth pressure will
be given from the trial wedge procedure but the line load will
also change the point of application of this total pressure.
The method given in Figure 17 may be used to give the
distribution of pressure.
Alternatively, when the line load is small in comparison with
the active earth pressure, the effect of the line load on its
own can be determined by the method given in section (A) of
Figure 48. This is based on stresses in an elastic medium
modified by experiment. The pressures thus determined are
superimposed on those due to active earth pressure and other
effects.
60

POINT LOADS

Point loads cannot be taken into account by trial wedge pro-


cedures. The method based on Boussinesq's equations given in
section (B) of Figure 48 should be used. A similar method is
given in the Earth Retaining Structures CP2 (Institution of
Structural Engineers, 1951). An alternative semi-empirical
method is given by Henry (1986).
NON-UNIFORM SURCHARGES
Non-uniform surcharge pressures can be incorporated using the
Culmann Line method shown on Figure 47.
Section 6

EFFECTS OF WATER
61

SECTION 6 - EFFECTS OF WATER

GENERAL

The presence of water behind a wall has a marked effect on the


pressures applied to the wall. When the phreatic surface
intersects the wall, a hydrostatic pressure is exerted against
the wall, together with uplift pressures along the base of the
wall. Even when there is no water in direct contact with the
wall, such as when adequate drainage is provided, there is an
increased pressure on the wall due to the increased earth
pressure (section 6.2). The effect of water behind the wall is
significant; the total force may be more than double that
applied for dry backfill. Many recorded wall failures can be
attributed to the presence of water.
The height to which water can rise in the backfill, and the
seepage pressures, are both of prime concern. To determine
these the ground water conditions must be established. These
may be best derived from the observation of ground water
conditions prior to construction. However, possible changes to
the existing groundwater regime due to the construction of the
retaining structure should also be considered.
The effect of leakage from services can be significant. This
leakage contributes substantially to both perched and main
groundwater tables.
Where inadequate drainage is provided behind a retaining
structure, there may be a damming effect which would result in
raising groundwater levels locally and in the general area.
Such a rise may adversely affect the stability of slopes and
retaining walls. Effective drainage measures should always be
provided in such cases. In the absence of such measures or
where drainage may be impractical such as behind sheet retaining
structures, design should allow for appropriate water pressures.
STATIC WATER LEVEL
Where part or all of the soil behind a wall is submerged below a
static water level, the earth pressure is changed due to the
hydrostatic pore pressures set up in the soil. The water itself
also exerts lateral pressure on the wall equal to the depth
below the water table times the unit weight of water, see
section (A) of Figure 49.
When a soil is submerged, its effective unit weight is reduced
to Y‘ = vsat - Yw. The lateral earth pressure should, in this
case, be calculated using Y’ in equations or charts. The method
of analysis is illustrated by the static pressure distributions
on Figure 43. Alternatively, in graphical procedures such as
the trial wedge method, all forces acting on the soil wedge,
including the hydrostatic normal uplift pressure on the failure
plane and the lateral hydrostatic pressure, may be included in
the trial wedge procedure. This is illustrated in Figure 14.
62

In cohesive soils the pore water pressures set up during


construction will override any hydrostatic pore pressure until
dissipation of the excess pore water pressures has occurred, see
section 2.5.
Where tension cracks occur, hydrostatic water pressure should be
included for the full depth of the crack zo, as given in section
3.3.4 or H/2 whichever is less. If however shrinkage cracks are
liable to fonn to a depth greater than that given above, water
pressure should be allowed for the full depth of such shrinkage
cracks. The maximum depth of shrinkage cracks varies with the
type of soil and climate but may be taken as 1.5 m.
Full water pressure must be allowed for below the weep holes or
other drainage outlets.
Static water pressure always acts normal to the surface of the
wall.
SEEPAGE PRESSURE
If the water in the soil voids is flowing, the pore water
pressures will be changed from the hydrostatic values to values
determined by the seepage pattern. These values are then used
in the trial wedge analysis methods to determine wall forces.
For major structures, the pore water pressures under seepage
conditions should be determined by flow net procedures, see
Huntington (1957), Scott (1963), or Terzaghi and Peck (1967).
The pore water pressure normal to the failure surface of active
or passive wedges affects the earth pressure acting on a wall.
The resultant uplift force on the failure surface determined
from a flow net is applied in the force polygon for the soil
wedge together with any hydrostatic water pressure at the wall,
see Figure 14. For an approximate analysis the uplift intensity
may be taken as being equal to the pressure of the vertical
height of water between ground water table level (may be
sloping) and a point directly beneath on the failure surface.
Section (B) of Figure 49 shows a flow net for seepage from the
ground surface behind a wall to the vertical drain. For
cohesionless materials sustained seepage under the conditions
shown could increase the wall forces 20-40% over that for dry
backfill, depending on the backfill shearing strength.
DYNAMIC WATER PRESSURE IN EARTHQUAKES

The dynamic pressure of any water in the backfill should be


taken into account by applying the seismic coefficient to the
weight of water in the failure wedge as well as to the soil.
See section 4.5.6 for detailed description of the recomended
method of analysis, and this is illustrated on Figure 43.
63

DRAINAGE PROVISIONS
Water pressures must be included in the forces acting on the
wall unless adequate drainage is provided. For walls less than
2 m high, drainage material is usually only provided on the back
face of the wall, with weep holes to relieve water pressure, see
Figure 50. In these circumstances it may be desirable or more
economic to design for hydrostatic water pressure.
In general, if a drainage system similar to that shown in Figure
51 is used, water pressures may be neglected both on the wall
itself and on the soil failure plane. Adequate drainage reduces
the rate of softening of clay filling and of stiff-fissured
clays and lessens the likelihood of reductions in the strength
of the foundations, and is therefore very desirable for clay
soils. '
It is worth noting that in cohesionless soils, the active force
on a wall with static water level at the top of the backfill is
approximately double that for a dry backfill. For walls over
6 m high, particular care should be taken to ensure that the
drainage system will control the effects of water according to
the assumptions made in design. Many recorded wall failures
seem to be the result of inadequate drainage. Water should
preferably be prevented from entering the backfill from the
surface, otherwise any resulting seepage pressures must be
allowed for in design.
For a drain to be effective it must be able to carry the design
flow of water without backing up or blocking. The rate of
seepage into the drainage material and flow rate that the drain
can accommodate depends on the permeability of the drainage
material, thickness of the drain and hydraulic gradient in the
drain. Using a flow net sketch the flow into the drain may be
estimated and the required cross sectional area (A) of the drain
found using Darcy's law:
A = qd
PEI
where Kd = permeability of drainage material

i = hydraulic gradient
qd = the flow rate through drain
Cedergren (1977), gives methods for constructing flow nets and
applications to drainage problems which may provide guidance for
situations where drainage is important.
As a very general guide drainage material should have a
permeability at least 100 times that of the material it is meant
to drain. If this is achieved, pore water pressures due to
seepage will be minimised at the boundary and the soil mass will
drain as though it had a free boundary. Permeabilities of
granular (drainage) materials are given in Figure 52.
64

6.6 FILTER REQUIREMENTS


6.6.1 General
To prevent blockage, a drain must be protected from infiltration
of fine particles by an adequate filter. Filters should be more
permeable than the protected soil.
6.6.2 Graded Filters

The filter principle must be used when seepage from fine grained
to coarse grained drainage materials has the potential to move
fines and block the drain.
The following particle size ratios should generally be provided:

D150 < 5 5 < D150 < 40


985E T ’ DIEF '
Where D15¢ = size at which 15% by weight of the coarse material
is finer
015E = size at which 15% by weight of the fine material
is finer
D35E = size at which 85% by weight of the fine material
is finer
For clay soils D15; size should not be less than 0.2 mm.
The filter material must have sufficient permeability so that
the seepage can pass through to the drainage material or drain.

The following grading is the finest acceptable for any filter


material regardless of the material that is being protected:
Sieve Size Percent Passing

100
92
|dl\>-b 533 74
um 50
um 25
}I
—l(*)ml U'ICD$l-*LA)\l @@@@O3U'| pm 8
75 um O
Material surrounding a perforated subsoil drain pipe must have a
D35 size greater than the diameter of the pipe perforations.
6.7 GEOTEXTILES

In some cases, it may be possible to use man-made fibrous woven


and non-woven fabrics, known as geotextiles, to protect the
drainage facilities.
As yet, there is little experience in New Zealand with the long
term performance of fabric filters for permanent drainage
measures. Consequently, it is recommended that they should only
65

be used in low risk situations, and where failure could not be


expected to occur even if total blockage of the fabric occurred.
It is also recommended that they should only be used in
locations where they can be replaced if found to be defective
after a period in operation.
There are objections to the use of some of these materials, such
as serious deterioration on exposure to sunlight and ultra-
violet light, clogging due to movement of fines, reduction in
permeability due to compression, constructional difficulties and
geotextiles forming planes of weakness in the works. If these
objections are overcome by attention to design, construction and
quality control, then the availability of geotextiles provides
new opportunities for innovative filter/drain design and
construction.
Fabric filters should be properly designed to be in filter
relationship with the surrounding soil. Care must be taken to
select a geotextile which is appropriate to the grading of the
soil it is intended to protect and has adequate drainage
capacity for the particular application. A summary of design
criteria for fabric filters is given in the books by Rankilor
(1981), John (1987), and Koerner (1987).
Available literature suggests that fabrics with an equivalent
opening size of less than 150 um (or an open area of less than
4%) and the thicker non-woven fabrics, may be more prone to
clogging than other varieties. The use of these types should
therefore be avoided unless the satisfactory performance of the
particular soil/fabric/drainage-medium system has been
demonstrated by permeability test. On the other hand, some of
the very thin fabric varieties exhibit quite large visible gaps
caused by uneven distribution of fibres, and the use of such
defective materials should also be avoided.
During construction, stringent measures are required to ensure
that the manufacturer's instructions concerning storage and
handling are strictly followed, and that storage, placement and
backfilling of fabrics are carefully controlled to avoid
excessive exposure to ultra-violet light, mechanical damage and
ineffective overlapping. It is prudent to use two layers of
fabric as a precaution against impairment of the filter function
by mechanical damage during placement.
Section 7

STABILITY OF RETAINING WALLS


66

SECTION 7 - STABILITY OF RETAINING WALLS

GENERAL
The stability of a free standing retaining structure and the
soil contained by it, is determined by computing factors of
safety or ‘stability factors‘ which may be defined in general
terms as:
F _ Moments or forces aiding stability
5 ' Moments or forces causing instability
Factors of safety should be calculated for the following
separate modes of failure:
(a) Sliding of the wall outwards from the retained soil.
(b) Overturning of the retaining wall about its toe.
(c) Foundation bearing failure.

(d) Slip circle failure in the surrounding soil.


The forces that produce overturning and sliding also produce the
foundation bearing pressures and therefore (a) and (b) above are
interrelated with foundation bearing failure in most soils.
Generally bearing capacity is critical for counterfort walls on
soil foundations. Overturning stability is often critical for
walls on strong foundation materials such as rock or when the
base of the wall is small eg, crib walls.
In general, to limit settlement and tilting of walls on soil
materials, the resultant of the loading on the base should be
within the middle third for static loading and within the middle
half for earthquake loading. For rock foundation material, the
resultant should be within the middle half of the base for both
static and earthquake loading.
When calculating overall stability of the wall the lateral earth
pressure is calculated to the bottom of the blinding layer, or
in the case of a base with a key, to the bottom of the key as
shown in Figure 1.
The vertical component (if any) of the resultant earth pressure
is added to the weight of the wall system when computing
stability factors.
If the passive resistance of the soil in front of a wall is
included in calculations for stability, either the top 0.5 m of
the soil should be neglected, or only 2/3 of the calculated
passive resistance should be used.
Stability criteria for free standing retaining walls are
sumarised in Figure 53.
67

7.2 SLIDING STABILITY


7.2.1 Base without a Key
Sliding occurs along the underside of the base. The factor of
safety against sliding F5 (sliding) may be calculated from:

F5 (sliding) = c'b B + (Wt + P'V) tan 8'b + Pp


PH
Where Wt total weight of wall
vertical component of active earth pressure
effective vertical component of PA = Pv - u B
effective angle of base friction
fiO'7'U effective adhesion at base
force due to passive earth pressure
I'DU'U'<< horizontal force component of active earth pressure
average pore water pressure beneath the base
TC'---"U
U'U breadth of base

The F5 (sliding) must be at least 1.5 for normal loading and at


least 1.2 for earthquake loading. The equation is based on the
full passive pressure Pp being developed and implies that wall
movements in the order of those suggested in section 3.2 are
acceptable.
For cases where the wall movements are unacceptable or the
development of full passive pressures is unreliable the design
should take a more conservative approach. It is standard
practice in some codes to allow for only one half of the full
passive resistance or to increase the minimum factor of safety
to greater than 2.0 with full passive resistance.

7.2.2 Base with a Key


If the passive resistance of the soil in front of the wall is
neglected, the critical surface for sliding is usually the
horizontal plane through the bottom of the key. If passive
resistance is to be included, the critical failure surface may
be along a plane from the bottom of the key to the toe (reduced
passive resistance). The weight of soil in front of the key and
below the base down to the critical failure surface should be
included in the total weight, see Figure 1.
The factor of safety against sliding should be as in section
7.2.1 above. The angle of base friction may be taken equal to
the ¢' of the foundation soil since the failure surface is
almost completely in the soil.
68

7.3 OVERTURNING STABILITY


7.3.1 Gravity and Counterfort Retaining Walls
Moments calculated about the bottom of the front of the toe must
give an overturning factor of safety:
. 2 Moments resistin o erturn'
F5 (overturnlng) = Z Moments causinggovgrturnihgq
of at least 2.0 for normal loading and of at least 1.5 for
earthquake loading, see Figure 53.
For semigravity and counterfort walls, only the overturning
factor of safety of the wall as a whole is significant. For
cribwalls and other walls for which the base and upper portion
are usually separate units, the factor of safety of the upper
portions should also be calculated.
7I302 Sheet Walls
with sheet retaining walls the value of factor of safety can be
very sensitive on how the factor of safety is defined. Two
types of definition are commonly used, ‘Load Factors‘ eg, the
method described in section 7.3.1 or ‘Strength Factors‘ which
are determined as follows:
Fs = Available Soil Strength
Mobilised Soil Strength
Factors of safety based on loads are more comonly used because
they are generally easier to calculate. However there are a
number of possible definitions of resisting and driving moments
and some of these result in inappropriate design solutions.
The method described by Burland et al (1981), is simple to use
and results in appropriate design solutions for a range of soil
conditions. The method assumes that the resisting moments are
represented by the ‘net passive pressure‘ ie, the passive
pressure minus the active pressure generated below excavation
level. The method is illustrated in summary form on Figure 54
using the case of cohesionless soil with no water pressures, for
both free cantilever and anchored sheet walls. The method is
applicable to soils with cohesion and friction (c - ¢ soils),
for undrained conditions and where surcharge or water pressures
exist. These cases are described in detail in Burland et al
(1981).
Factors of safety of between 1.5 and 2.0 are generally
considered appropriate for gravity loads and between 1.25 and
1.5 with earthquake loads. The lower values can be used in
granular soils where material strengths and water pressures are
accurately known and the higher values should be used in
cohesive materials or where there is uncertainty in the value of
material strength or water pressure.
69

7.4 FOUNDATION BEARING PRESSURES


4 Vertical Central Loads
The ultimate bearing capacity of the foundation soil on which an
earth retaining structure rests should generally be determined
from a theoretical analysis of the foundation using
representative soil properties obtained as described in section
2. The applied loading should provide a factor of safety of 3.0
against ultimate bearing failure for static loading and a factor
of safety of 2.0 for earthquake loading. Saturated sand or silt
which may be susceptible to liquefaction (see section 2.12) is
not recommended as a foundation material.
The recommended methods of calculating the bearing capacity
applies to both earthquake loading and static loading. In
general no consideration need be made for the cyclic effects of
dynamic load or the dynamic properties of the soil. However, in
special problem soils, eg, collapsible or liquifiable soils
special consideration is required.
Other factors which may influence the bearing capacity are the
foundation depth, soil compressibility, scale effects and non-
homogeneous soil conditions. These are discussed by Vesic
(1975).
The general expression for ultimate bearing capacity of a
horizontal infinite length strip foundation on level ground
subjected to vertical central loads is:
qu]t = gt = C NC + Y D Nq + % Y B NY

The bearing capacity factors NC, Nq, NY, are dimensionless


functions of the angle of shearing resistance, ¢, of the soil.
A number of different methods for calculating N5, Nq and NY are
presented in the literature. One of these methods, Brinch
Hansen (1970), as revised in Danish Geotechnical Institute
(1978), and presented in Tomlinson (1987), is considered to
produce answers which compare well with experimental results.
Using this method, the bearing capacity factors are calculated
as follows:
Nq = tan’ (45° + ¢/2).e"ta"¢
NC = (Nq ' 1)COt¢

NY = 1.5 (Nq - 1) tan¢


These bearing capacity factors can also be obtained from Figure
55.
Where appropriate, these bearing capacity factors are then
modified by empirical factors to account for different
foundation conditions, so that the formula can be extended to:
70

qu]t = C N¢.S¢.dc.ic.b¢.gc ‘I’ Y D Nqu5q.dq¢Iqobqugq

+ 92 Y B

where sc,sq,sy - shape factors to account for the shape of the


foundation;
dc, dq, dy - depth factors to account for embedment depth
to the base of the foundation;
i5, iq, iy - inclination factors to account for the
inclined loads on the foundation;
bc, bq, by - base inclination factors to account for the
inclined base of the foundation;
g¢, gq, gy - ground inclination factors to account for the
inclination of the ground.
The loading arrangements for Brinch Hansen's formula as
presented in Tomlinson (1987) is shown on Figure 56. The shape
factors sc, s and sy for centrally applied vertical loading can
be obtained grom the diagram (A) based on Tomlinson (1987),
shown on Figure 57. Approximate values can be obtained from the
table shown on the same Figure. Where the loading is inclined,
Brinch Hansen (1970) gives shape factors for bearing capacity in
each direction (parallel to L or B) and these are also presented
on Figure 57.
Values of the depth factor d¢ can be obtained from Figure 58.
The value of depth factor dq can be derived from dc and Nq as
shown on this figure. Depth factor dy can always be taken as
unity. Use of depth factors assume that the shear strength of
the soil above foundation level is not less than the soils below
the foundation. Therefore where soft or loose soils exist above
foundation level, or where there is a possibility that soil in
front of the wall could be disturbed, eroded or removed, the
depth factors should not be used. Inclined loading must be
considered from two directions, ie, in the direction of
effective breadth B‘ and effective length L‘ of the foundation,
see Section (B) on Figure 56. The inclination factors ic, iq
and iy can be obtained from Figure 59.
Base inclination factors bq and by can be derived as in Brinch
Hansen (1970) from the following for an inclination of the base
of n as shown on diagram (A) of Figure 56. These are extracted
from Bowles (1982).
bc = 1 - q“/147°
bq = by = e‘ Zfltan P where q is in radians
Where the ground surface slopes away from the foundation at an
inclination of B as shown on Figure 56, the ground inclination
factors gc, gq and gy can be calculated and used as proposed in
Danish Geotechnical Institute (1978):
71

gc = e-2Btan¢

gy = gq = 1 - Sin 28 where B is in radians

Foundations constructed on soils of relatively high permeability


require the analysis of bearing capacity to be determined in
terms of effective stress, see section 2.5. Under these
conditions the contribution to the bearing capacity of the
cohesive terms is in general very small and may be neglected.
For foundations constructed on saturated clayey soils of low
permeability, the short term stability is often critical and in
this case the problem is usually analysed using undrained
strength (¢ = o) in terms of total stress, see section 2.5.
Brinch Hansen (1970) gives a separate solution for the ultimate
bearing capacity for the undrained case. Undrained bearing
capacity is also appropriate for the consideration of earthquake
loading.
For this special case with ¢ = 0, the bearing capacity can be
calculated using the equations given below:
qu'|t = Su + Sc’: 'I' dC* " Ic* " bc* - gc*] "

Where s¢*, d¢*, ic*, b¢* and g¢* are the factors to take account
of the shape of the footing, the depth of the base, the
inclination of the load, the inclination of the base and the
inclination of the ground respectively. The factors can be
derived as follows:
s¢* = o.z B/L
ac* = 0.4 D/B if o 5 B
= 0.4 tan-1 D/B if o > B

i¢* = 0.5 - 0.5 / H


B‘L‘Su
bc* = n°/147° (for n = 0, b¢* = 0)

gc* = 6°/147° (for B = 0, gc* = O)

The bearing capacity factors of Figure 46 have been determined


on the assumption that the foundation material is reasonably
incompressible, so that failure would occur by general shearing.
For compressible materials, failure occurs by local or punching
failure. For these materials Terzaghi (1943), recommended that
the value of cohesion used should be reduced to 2 c‘/3, and the
angle of shearing resistance to tan'1 ((2 tan o‘)/3). More
detailed recomendations for preventing local failure are given
in Vesic (1975), and Peck et al (1974).
72

7.4.2 Eccentric Loads


If the load on the foundation is eccentric this can sub-
stantially reduce the bearing capacity. To allow for this the
base width, B, is reduced to an effective width B‘ given by:

B‘ = B - 2e
where e is the load eccentricity
For a footing eccentrically loaded in two directions the
effective dimensions of the base become such that the centre of
an area A‘ coincides with the vertical component of the applied
load V.
A‘ = B‘ x L‘
where L‘ = L — 2e;
B‘ = B - 28b

where e2 is eccentricity in the longitudinal direction and eb is


eccentricity in the transverse direction.
L‘ and B‘ replace L and B in all equations.
The factor of safety is given by:

F5 (bearing) = C _l ('0'

21°
where q = %T for a rectangular footing
V
or 9 = BT for a continuous strip footing
When a foundation carries an eccentric inclined load, an
estimate of the ultimate bearing capacity lnay be obtained by
combining the methods given in sections 7.4.1 and 7.4.2.
7.4.3 Miscellaneous Foundation Conditions
In addition to the foundation conditions specified above,
bearing capacity can be influenced by shallow water tables,
layering of soils, and the proximity of slopes. Methods are
available to calculate bearing capacity in these circumstances
and reference should be made to Bowles (1982).
7.4.4 Foundations on Rock
Foundations on continuous rock seldom present problems since the
rock is stronger than most foundation materials. Structural
defects and discontinuities, or the compressibility of the rock
mass below the foundation, usually control the allowable bearing
pressure.
73

Where discontinuity controlled failure mechanisms are possible,


joint surveys should be carried out in the excavation and
adjacent slopes.
The compressibility of the rock mass below foundation level
depends on the frequency of joints and on the amount and type of
infilling of these joints in the zone of influence of the
foundation. RQD (rock quality designation) is defined as:
Cum. Length of Intact core 2 100 mm in length
RQD (%) = 100 x Length of core run
In unweathered rocks, R00 indicates the joint intensity, whereas
in weathered rock it gives a measure of the amount of compres-
sible material but no indication of the infill compressibility.
Where only tight clean joints are present, the correlation
between ROD and allowable bearing pressure proposed by Peck et
al (1974), given in Table 10, may be used.

TABLE 10: ALLOWABLE BEARING PRESSURE ON JOINTEO ROCK


(Peck et al, 1974)

RQD Allowable Pressure


(%) (MPA)

100 30
90 20
75 12
50 6
25 3
0 1

Note:
(1) Use allowable pressure or unconfined compressive strength of
intact rock, whichever is less.
(2) ROD is for rock in the zone of influence of the foundation.
For infilled joints deformation will be larger, and estimates of
the joint infill compressibility may be required. The effect of
joint infilling on allowable bearing pressure for a limited
range of joint spacing and thickness is given in the Canadian
Foundation Manual (Canadian Geotechnical Society, 1978).
74

SLOPE FAILURE IN SURROUNDING SOIL


The overall stability of the ground surrounding the retaining
wall should be investigated, and calculations should be carried
out on the full range of potential failure surfaces to ensure
that an adequate factor of safety against overall slope failure
is maintained. The calculations should include the influence of
the surcharge from the wall on the slope.
For walls higher than 9 m, slip circle failure in the soil
containing the wall should be investigated. The slip circle
safety factor:

F5 (slip circle) = L-Ii__N'“


ta? ¢' T c'l

Where N is the normal stress acting on the failure surface


¢‘ is the effective angle of shearing resistance
c‘ is the cohesion in terms of effective stress
l is the length of the failure surface
T is the tangential force along the failure surface
u is the pore water pressure
should be at least 1.5 for static loading and at least 1.3 for
earthquake loading. An effective stress analysis using
appropriate pore water pressures is recommended.
Earthquake loading should be allowed for by applying an
equivalent static horizontal force to the soil mass as described
in section 4.
A computer program, ICES-LEASE, is currently available for this
type of analysis and has the capability for including the
horizontal force from the earthquake loading.
Section 8

STRUCTURAL DESIGN
75

SECTION 8 - STRUCTURAL DESIGN

8.1 GENERAL
8.1.1 Codes
Reinforced concrete design shall be in accordance with
NZS 3101:1982 ‘Code of Practice for the Design of Concrete
Structures‘ (Standards Association of New Zealand, 1982).
8.1.2 Limit State Design
When limit state design methods are used for proportioning a
structural section, the design loads shall be computed so that
the capacity of the section shall not be less than:
U = 1.35 (DL + 1.35 EP + N); or
U = 1.00 (kDL + EQ + 1.35 W))

Where DL = dead load of the structural element


EP = static earth pressure acting on the element
(including the effects of any surcharge loads)
E0 = earthquake earth pressure acting on the element
U = ultimate load
W = hydrostatic water pressure
k = 1.3 or 0.8 whichever is more severe, to allow for
vertical acceleration
when strength limit state design is used, a serviceability check
on crack widths at working loads shall be made to ensure that
the limits given in clause 4.4.2.3 of NZS 3101 are not exceeded.
Ferguson (1958), gives an example of limit state design of a
cantilever retaining wall.
8.1.3 Cover to Reinforcement
Particular attention shall be given to the cover of
reinforcement in both the design detailing and during
construction.
8.1.4 Selection of Wall Type
For walls up to 7.5 m high where crib walling is not suitable, a
cantilever wall will usually be found to be the most economical.
For higher walls an investigation should be made for the
relative economies of using a counterfort or cantilever wall.
This should take into account unit costs for formwork,
reinforcing steel, and concrete, and not just "all in“ cost per
cubic metre of final wall. Counterfort walls should have
approximately a 9 m bay length (varied to suit architectural
finish, etc) with three counterforts per bay. The position of
76

the counterforts is obtained by considering the stresses in the


stem.
8.2 TOE DESIGN
For
Length of toe > 1
Effective depth at face of support ’
design as a cantilever in accordance with NZS 3101. Shear may
be taken at ‘d‘ out from face of support.
For
Length of toe < 1
Effective depth at face of support ‘ ’
design as a bracket in accordance with section 7.3.13 of
NZS 3101.
8.3 STEM DESIGN
8.3.1 Stem Loading
For the stem design in cantilever and counterfort walls, the
earth pressure acting on the vertical plane through the rear of
the heel is projected onto the stem.
For stem design of counterfort walls the earth pressures
resulting from backfill compaction shall be included in the
design loadings.
8.3.2 Lower Section of Counterfort Wall Stem
The bottom L5/2 of the stem is to be reinforced for vertical
spanning action in addition to horizontal spanning.
Bending moments per unit height of stem may be assumed as:
WL52

-ve M = 14 (horizontal steel)

V/LS2

+ve M = 22 (horizontal steel)

WLS2

-ve M = 25 (vertical steel)


77

where w is the lateral design pressure at the level being


considered, and
L5 is the length of the~retaining wall stem.
8.3.3 Horizontal Moments in Counterfort Wall Stem
Bending moments in the top part of the stem may be calculated
from:
WLS2

-ve M = 11 (horizontal steel)

WLS2

+ve M = 16 (horizontal steel)

Use continuous horizontal steel in both faces. Horizontal BM


variations with height should be catered for by varying the
reinforcement spacing in preference to changing the bar sizes.
When calculating the bending moments for the stem, the span
should be taken as the clear span between counterforts (L5).
8.4 HEEL SLAB DESIGN
8.4.1 Loading
The design loading on the heel slab is shown in Figure 60. The
foundation bearing pressures may be calculated“ by using the
theory of subgrade reaction, see section 2.9. For a rigid base
slab this theory gives bearing pressures which vary linearly
across the base width.
The pressures for use in structural design are not the same as
those used to check the factor of safety against ultimate
bearing failure (section 7.4). They are normally taken as the
bearing pressures at working loads, ie,
(a) If the resultant cuts the base within the middle third the
toe and heel pressures for structural design may be
calculated from:
p = V/BL 1 6 Ve/BZL
Where V = the vertical component of the resultant loading
on the base
B = the base width
L = the length of wall for which the resultant earth
pressure is calculated (usually unity)
78

(b) If the resultant lies outside the middle third:


2 V
Pmax = 3 (B/2 - e) L

8.4.2 Heel Slabs for Counterfort Walls


The heel slab for counterfort walls should be designed as a slab
spanning in two directions if a key is included at the rear.
The design bending moments may be obtained from tables in Bowles
(1982).
Alternatively, the heel slab can be divided into four or five
strips, of approximate width 1 to 1.5 m spanning between
counterforts. The outermost strip including the key can be
designed as an L-beam for bending, its breadth equal to the
strip width. The width of the key strip resisting shear should
be assumed as the maximum width of the key plus half the
thickness of the heel slab. Bending moments may be calculated
as in section 8.3.3.

Each strip should be designed for the average load occurring.


The critical section is at the face of the counterforts. This
shear will usually govern the heel thickness.
The heel slab should also be considered as strips spanning at
right angles to that mentioned above, ie, between stem line and
key strip. Simple assumptions can be made as to end fixity of
these strips and an approximate amount of reinforcing provided.
8.5 COUNTERFORT DESIGN
Vertical steel in the counterfort is required to carry the net
tensile load from each strip of the heel slab. The main moment
reinforcement for the wall is usually concentrated at the back
of the counterfort. Where it joins the heel slab, the above
steel should be considered as taking only that load occurring on
the outermost strip incorporating the key, as defined in section
8.4.2 above.
Horizontal steel in the counterfort is required to carry the net
load on each horizontal strip of stem.

Cut-off positions for the main tensile steel in the counterforts


is shown in Figure 61.
8.6 KEY DESIGN
In general the depth to width ratio of the key should be approx-
imately one. It is difficult to predict the force that will act
on the key. An approximate design horizontal load on the key
is:
79

= Z horizontal loads causing sliding


- 0.4 x total vertical loads above blinding layer
This load acts at 1/3 key height from bottom of key. Design the
key as a bracket - refer to section 8.2 above. Note some
stresses are carried from the key into the bottom of the heel
slab, and will call for some reinforcement in that area.
CONTROL OF CRACKING
To minimise cracking in the retaining structure:
(=1) Provide shrinkage and temperature reinforcement equal to
0.25% of the gross concrete area as a minimum in both
directions in all members.
In the stem:
2/3 of this steel to be on the outside face
1/3 of this steel to be on the earth face.

(b) Specify that the temperature during concrete placing and


curing is to be kept as low as practical especially in the
summer period.
(B) Specify successive bay construction.

(<1) Specify early curing for the purpose of cooling so as to


minimise the heat rise.
(E) Place the steel in suitable bar sizes to limit crack width
to 0.25 mm (see code requirements).

(T) Added protection can be given by painting the earth face


with say two coats of "Mulseal" or “Flintcoat", see Evans
and Hughes (1968).
Section 9

SPECIAL PROVISIONS FOR CRIB WALLS


80

SECTION 9 - SPECIAL PROVISIONS FOR CRIB WALLS

GENERAL
A considerable amount of literature is available from Cribwall
Unit Manufacturers (eg, Hume, ICB, Cement Products) and also
Portland Cement Association on the design of crib walls.
However, care must be exercised in the interpretation of this
data. Crib walls must be checked for stability in accordance
with section 7. Figures 62 and 63 may be used as an aid in
determining the maximum height for different wall thicknesses.
The crib units and wall construction should be in accordance
with the current WORKS standard specification for this work, see
CD 209:1988 (WORKS, 1988).
DESIGN LOADING
The pressures acting on a typical crib wall are shown in Figure
64. These pressures are calculated by the methods of section 3.
Cribwalls are often used for low walls of low importance,
therefore earthquake loading will usually not be applied. If
earthquake loading needs to be included the methods of section 4
should be used and the crib wall should be considered as a
dynamic active wall.
FOUNDATION DEPTH
The minimum depth of foundation shall be as shown in Figure 64
which includes a continuous concrete foundation slab. A minimum
slab thickness of 150 mm reinforced with one layer of 665 mesh
is recommended to prevent differential settlement of the wall
structure. The consequences of such settlement are described in
Tschebotarioff (1965).
DRAINAGE

A continuous 150 mm diameter (minimum) subsoil drain should be


provided at the rear of the foundation slab, to ensure a dry
foundation. This should be provided for all heights of crib
wall.
Adequate drainage of the whole crib structure is essential.
Many of the failures in crib walls have occurred because
material of low permeability was used" as backfill - thus
developing high static or seepage water pressures. A free
draining backfill should always be used if possible, otherwise
the effect of water should be allowed for.
Unless effectively drained over the full height, crib walls
should be designed to resist lateral hydrostatic pressures in
addition to earth pressures.
81

MULTIPLE DEPTH WALLS


Walls of more than single depth should be checked at the changes
from single to double and double to triple depth to satisfy the
following stability criteria:
(a) For normal conditions resultant to be within middle third
of base. This will ensure that no part of the wall
structure is in tension.
(b) For earthquake conditions resultant to be within the
section of the wall.

(c) The appropriate overturning factor of safety must also be


met at these sections.
WALLS CURVED IN PLAN
Crib walls with a convex front face are much more susceptible to
damage by transverse deformations than are concave walls, see
Tschebotarioff (1965).
Section 10

SHEET RETAINING STRUCTURES


e2 _

SECTION 10 — SHEET RETAINING STRUCTURES

10.1 GENERAL
Walls which have uniform cross-section with depth are considered
in this chapter. These include flexible sheet structures, such
as sheet piled and soldier piled walls, and more rigid walls,
including diaphragm and caisson walls.
The earth pressure which acts on an earth supporting structure
is strongly dependent on the amount of lateral deformation which
occurs in the soil. For flexible sheet walls, the determination
of deformations, and hence the earth pressures, is not simple,
because the yield of one part of a flexible wall throws pressure
on to the more rigid parts. Hence, the pressures in the
vicinity of the supports are higher than in the unsupported
areas, and the loads on individual supports vary depending on
the stiffness characteristics of the supports themselves.
Therefore it may be necessary to design a stiff wall to resist
at rest pressures.
Deformation of the ground adjacent to excavations may cause
breakage of water carrying services. In situations where large
flows may result, the prudent designer will allow for the water
table being at the ground surface when calculating loads to be
retained.
10.2 STRUTTED EXCAVATIONS
Strutted sheet piling is often used to provide temporary support
for the sides of deep excavations. The sheet piles are usually
driven first with support struts being installed as the
excavation proceeds. The final deformations of the wall are
highly dependent on the construction sequence and detailing.
The CIRIA Report 97 on Trenching Practice (Irvine and Smith,
1983) gives guidance to safe practice in the design and use of
temporary support for trenches not deeper than 6 m.
Failure of a strutted wall often results from the initial
failure of one of the struts, resulting in the progressive
failure of the whole system. The forces in identical struts in
any particular support system may differ widely because they
depend on such factors as the way in which the struts are
preloaded and the time between excavation and installation of
struts. Loads in similar struts in any set of observations have
been found to vary from the average value by up to i 60% (Lambe
et al, 1970).
Since failure of strutted cuts often occurs by structural
failure, particular attention should be paid to the structural
detailing of the internal strutting. Guidance on the structural
design of such walls, together with typical details of
83

connections and strutting systems, are given by Goldberg et al


(1975). Struts must be sufficient for all stages of
construction.
The distribution of pressure on a strutted excavation is com-
plex, and it is normal to use a pressure envelope covering the
normal range pressure distributions. The envelopes (Figure 65)
given by Peck (1969), and the Japan Society of Civil Engineers
(1977), together with loadings from groundwater and surcharge,
should be used to determine strut loads for all internally
strutted excavations. In assessing loading from groundwater,
the effect of accidental breakage of water carrying services
should be considered.
The load carried by each internal strut is estimated by assuming
that the sheet pile is simply supported between struts, and that
a reaction below the base of the excavation exists. This
reaction is provided by the passive resistance of the soil
beneath the cut.
The depth of penetration of the wall below the base of the
excavation should be sufficient to provide this reaction.
Since the wall moves towards the excavation, it may be assumed
that active and passive pressures develop against the wall below
the excavation level, and horizontal equilibrium may be used to
determine the depth of penetration. The passive resistance
should be factored by 2.0.
For soft clays, negligible passive resistances develop, and the
lower section of the wall must be designed as a cantilever, and
the bending moment and deflection must be checked.
The maximum bending moment at, or below, the lowest strut should
be checked against overstressing of the wall.

Instability of the base of an excavation can occur due to shear


failure in soft to firm clays (known as base heave). In
granular materials, piping or heave associated with groundwater
flow can occur.
The factor of safety with respect to shear failure is given by:
F5 =

YH + q

where the terms are defined in Figure 66. Where F5 is less than
2 substantial deformations may occur with consequent loss of
ground, and the probability of failure exists. Figure 66 also
shows a chart from Janbu et al (1956), from which the Stability
Number (Nb) can be obtained.
Where soft clay extends to considerable depth below the
excavation, the effect of increased sheeting stiffness, or
84

depth, is minimal. However, driving the sheeting into a hard


stratum before commencing the excavation can appreciably reduce
the deformations.
Control of the groundwater may be necessary to prevent piping or
heave associated with groundwater flow. This may be achieved by
wellpoint dewatering for large excavations. For further details
on dewatering, see Cedergren (1975).
It may be necessary to carefully consider the settlements caused
by the excavation in the vicinity which may cause significant
damage to nearby structures and services. Figure 67 gives an
indication of the settlement which could result from strutted
excavations, but the settlements are highly dependent on the
construction sequence and detailing.
10.3 ANCHORED FLEXIBLE WALLS
10.3.1 Walls Anchored Near the Top
The deformation of an anchored sheet pile wall depends on the
relative stiffness of the pile/soil system. For a relatively
rigid system, such as a heavy pile section in a loose sand, the
earth pressure distribution corresponds closely to the
triangular active and passive conditions. The toe of the pile
is assumed pinned, and the free earth support design method as
outlined by Clayton and Milititski (1986), Cornfield (1975), or
Teng (1962) is appropriate.
As the stiffness of the system decreases the pressure distri-
bution alters in such a way as to reduce the bending moment in
the pile. As a consequence, the sheet pile section used may be
reduced as compared with an infinitely stiff wall. Rowe's
Theory of Moment Reduction (1952, 1955, 1957) takes this effect
into account; it is summarised by Clayton and Milititski (1986),
and Teng (1962).
10.3.2 Multiple Anchored Walls
The multiple-anchored system of wall support results in the
retaining structure being progressively fixed. Consequently,
the lateral deformations are limited to such an extent that
failure within the retained soil is unlikely. The earth
pressure which finally acts on the wall depends on the relative
stiffness of the wall to the soil, the anchor spacing, the
anchor yield and the prestress locked into the anchors at
installation.
The earth pressure distribution has been shown to be similar to
that obtained for internally braced excavations. A rectangular
pressure envelope similar to that adopted by Peck as shown in
Figure 65, is appropriate. The earth pressure coefficient may
be taken as KA. However, it is comon to use a value between
KA and Kb, such as (KA + Kb)/2, in an attempt to control
surface movements.
85

Successful designs have been made using triangular pressure


distributions with earth pressure coefficients varying between
KA and Kb. However, because of the mechanism involved, the
rectangular distribution is considered more appropriate (Hanna,
1982). Anchor loads may be checked using both distributions,
and the worst case taken.
The determination of vertical and horizontal spacing of anchors
using the procedure for internal strut spacing gives acceptable
results. Another approach is the semi-empirical design method
of James and Jack (1974), which simulates the field construction
procedure using triangular pressure distributions. This method
allows determination of the depth of penetration required, and
results correspond well to field and laboratory tests.
10.3.3 Effects of Anchor Inclination
Anchors are usually inclined downwards, transmitting the
vertical component of the anchor force into the anchored member.
This force should be considered in design, together with the
weight of the member itself (White, 1974).
A number of cases have been recorded where soldier piles have
failed in end bearing due to the vertical component of the
anchor force.
10.4 CANTILEVERED WALLS
Relatively rigid cantilevered caisson walls rely entirely on the
development of passive resistance in front of the wall for their
stability. As a consequence, considerable movement must occur
before equilibrium is reached, and deep penetration is required.
The deflection at the top of the wall may be the governing
criterion. Such walls should not normally be used as permanent
structures to retain a height of more than 5 m unless canti-
levered from rock.
Installation of a drainage and filter medium behind the wall may
be difficult and so full hydrostatic pressure may have to be
considered for the design.
86

REFERENCES

Agour, M S and C B Brown (1974). The Prediction of Earth Pressure on


Retaining Walls Due to Compaction. Geotechnique, Vol 24, pp 489-502.
Anderson, W F, Hanna, T H, and Abdel-Malek, M N (1983). Overall Stability
of Anchored Retaining Walls. Proc. ASCE Jnl. of Geotechnical Engineering,
Vol. 109, No. 11, November.
Athanasiou-Grivas, D, (1978). Reliability of Retaining Structures During
Earthquakes, VI Symposium on Earthquake Engineering, University of
Roorkee, Vol. I.
Bowles, J E (1982). Foundation Analysis and Design. Third Edition.
McGraw-Hill International Book Co., Singapore.
Brinch Hansen, J (1970). A Revised and Extended Formula for Bearing
Capacity. The Danish Geotechnical Institute, Bulletin No 28.
Broms, B (1971). Lateral Earth Pressures Due to Compaction of
Cohesionless Soils. Proc. 4th Budapest Conference on Soil Mechanics and
Foundation Engineering.
Broms, B B and I Ingelson (1971). Earth Pressure Against the Abutments of
a Rigid Frame Bridge. Geotechnique Vol 21, No 1.
Burland, J B, Potts, D M, Walsh, N MI(1981). The Overall Stability of
Free and Propped Cantilever Retaining Walls. Ground Engineering, July,
pp 28-38. t
Canadian Geotechnical Society (1978). Canadian Foundation Engineering
Manual, Part 4. Canadian Geotechnical Society, Ottawa. 68 p.

Caqout and Kerisel, J. (1948). Tables for the Calculation of Passive


Pressures, Active Pressure and Bearing Pressure of Foundations.
(Translated from the French by M A Bec London) Gauthier-Villars, Paris.
Cedegren, H R (1975). Drainage and Dewatering. Foundation Engineering
Handbook. Edited by Winterkorn, H F and Fang, H Y. p 221-243.
Van Nostrand, Reinhold, New York.

Cedegren, H R (1977). Seepage, Drainage and Flow Nets, Second Ed. John
Wiley & Sons, New York. pp 534.
Chandler, R J, and Skempton, A W (1974). The Design of Permanent Cutting
Slopes in Stiff Fissured Clays. Geotechnique 24. No. 4. pp 457-466.
Chopra, A K (1967). Hydrodynamic Pressures on Dams During Earthquakes.
Jnl. of Engineering Mechanics Division, ASCE, Vol. 93, EM 6, pp 205-223.

Clayton, C R I and J Milititsky (1986). Earth Pressure and Earth


Retaining Structures. Surrey University Press.
87

Cornfield, G M (1975). Sheet Pile Structures. Foundation Engineering


Handbook. Edited by Winterkorn, H.F. and Fang, H.Y. Van Nostrand
Reinhold Company, New York, pp 418-444.
Cullen, R M and I B Donald (1971). Residual Strength Determination in
Direct Shear. Proc 1st Australian-New Zealand Conference on Geomechanics.
Danish Geotechnical Institute (1978). Code of Practice for Foundation
Engineering. Bulletin No. 32. Copenhagen. p 52.
Elms, D G, and Richards, R (1979). Seismic Design of Gravity Retaining
Walls, Bulletin of NZ National Society for Earthquake Engineering, Vol.12,
No.2.
Evans, E P and Hughes, B P (1968). Shrinkage and Thermal Cracking in a
Reinforced Concrete Retaining Wall. Proc. Institution of Civil
Engineers, Volume 39.
Ferguson, P M (1958). Reinforced Concrete Fundamentals. 2nd Edition.
John Wiley and Sons.
Franklin, A G and Chang, F K (1977). Earthquake Resistance of Rockfill
Dams : Report 5 : Permanent Displacements of Earth Embankments by Newmark
Sliding Block Analysis. Misc. Paper S-71-17, Soils and Pavements
Laboratory, US Army Engineers Waterways Experiment Station, Vicksburg,
Miss.
Geotechnical Control Office (1982). Geoguide 1: Guide to Retaining Wall
Design. Hong Kong Government Printer.
Goldberg, D T, Jaworski, W E and Gordon, M D (1976). Concepts for
Improved Lateral Support Systems. US Federal Highway Administration,
Report FHWA-RD-75-131, Washington. p.40. (National Technical Information
Service No. PB257210).
Hanna, T H (1982). Foundations in Tension - Ground Anchors - Trans Tech
Publications, Series on Rock and Soil Mechanics, Vol. 6, pp 573, McGraw
Hill, USA.
Henry, F D C (1986). The Design and Construction of Engineering
Foundations. Second Edition. London p 1 - 1090.

Ingold, T S (1979). The Effects of Compaction on Retaining Walls.


Geotechnique Vol 29, pp 265-284.
Institution of Structural Engineers (1951). Earth Retaining Structures.
Civil Engineering Code of Practice No 2. Prepared by Civil Engineering
Codes of Practice Joint Committee.
Irvine, D J and Smith, R J H (1983). Trenching Practice. Construction
Industry Research and Information Association, CIRIA Report 97, London, 62
pl
88

James, E L and Jack, B J (1974). A Design Study of Diaphragm Walls,


Proceedings of the Conference on Diaphragm Walls and Anchorages, London,
pp 41-49.
Janbu, N, Bjerrum, L and Kjaerrusli, B (1956). Veiledning red losning av
foundermenterringsoppgaker (Soil Mechanics applied to some engineering
problems). Norwegian Geotechnical Institution, Publication No. 16, 93 p.
Japan Society of Civil Engineers (1977). Guide to Tunnelling by Cut and
Cover Method. Japan Society of Civil Engineers, Tokyo, 203p.
John N W M, Geotextiles, Blackie USA, 1987.
Koerner R M (1987). Designing with Geosynthetics.
Lambe, T W and R V Whitman (1979). Soil Mechanics. SI Version. Wiley.
Lambe, T W, wolfskill, L A and Wong, I H (1970). Measured Performance of
Braced Excavations. Journal of the Soil Mechanics and Foundations
Division, American Society of Civil Engineers, Vol. 96, pp 817-836.
Mathewson, M B, Wood, J H, Berrill, J B (1980). Seismic Design of
Bridges, Section 9 : Earth Retaining Structures. Bull NZ Nat Soc
Earthquake Engineering, Vol 13, No 3, pp 280-293.
Mononobe, N and Matsuo, H (1929). On the Determination of Earth Pressures
During Earthquakes. Proc. World Engineering Conference. Vol. 9.
Nadim, F and Whitman, R V (1985). Seismically Induced Movement of
Retaining Walls, Norwegian Geotechnical Institute, Publication No. 155,
Oslo.
National Research Council (1985). Committee on Earthquake Engineering.
Liquefaction of Soils During Earthquakes. National Academy Press,
Washington DC.
National Roads Board (1989). Bridge Manual. Prepared by Consultancy
Services Division, Works and Development Services Corporation (NZ) Ltd,
for National Roads Board, Wellington, New Zealand (Preliminary).
O'Rourke, T D, Cording, E J, and ‘Boscardin, M (1976). The Ground
Movements Related to Braced Excavation and their Influence on Adjacent
Buildings. US Department of Transportation. Report No. DOT-TST 76,
T—23, 123 p.
Padfield, C J and Mair, R J (1984). Design of Retaining Walls Embedded in
Stiff Clay. Construction Industry Research and Information Association.
Report 104. London. 146 p.
Peck, R B (1969). Deep Excavations and Tunnelling in Soft Ground. 7th
International Conference on Soil Mechanics and Foundation Engineering,
Mexico City, State-of-the-art Volume, pp 225-290.
Peck, R B, Hanson, W E, and Thornburn, T H (1974). Foundation
Engineering. 2nd Edition, Wiley, New York, p 514.
89

Prakash, S and Saran S (1966). Static and Dynamic Earth Pressures Behind
Retaining Walls. Proc 3rd Symposium on Earthquake Engineering, University
of Roorkee, India, Vol 1, pp 277-288.
Prakash, S and Nandakumaran, P (1979). Earth Pressures During
Earthquakes. Proc. second US National Conference on Earthquake
Engineering, Stanford University.
Rankilor, P R. (1981). Membranes in Ground Engineering, Wiley,
Chichester, 1981.

Rowe, P W and Peaker, K (1965). Passive Earth Pressure Measurements.


Geotechnique, Vol. 15. pp 57-78. London.
Saran, S and Prakash, A (1970). Seismic Pressure Distribution in Earth
Retaining Walls. Proc. Third European Symposium on Earthquake
Engineering, Sofia.
Scott, R F (1963). Principles of Soil Mechanics. Addison-Wesley
Publishing Co.
Seed, H B, and Whitman, R V (1970). Design of Earth Retaining Structures
for Dynamic Loads. ASCE Speciality Conference, Lateral Stresses in the
Ground and Design of Earth Retaining Structures.
Standards Association of New Zealand (1982). NZS 3101 - Code of Practice
for the Design of Concrete Structures.
Standards Association of New Zealand (1989). General Structural Design
and Design Loadings for Buildings. 2/DZ 4203, Wellington - Second Draft.
Stevenson, R B (1987). Dynamic Sand Pressures Against a Flexible
Retaining Wall. Report No. 5-87/8. Central Laboratories, Ministry of
Works and Development, Lower Hutt.
Teng, W C (1962). Foundation Design, Prentice-Hall.
Terzaghi, K (1943). Theoretical Soil Mechanics. Wiley, New York.
pp 129-130.
Terzaghi, K and R B Peck (1967). Soil Mechanics in Engineering Practice.
2nd Edition. John Wiley and Sons.
Tomlinson, M J (1987). Foundation Design and Construction, 5th Edn.
Pitman, London.
Tschebotarioff, G P (1965). Analysis of a High Crib Wall Failure. Proc
6th International Conference on Soil Mechanics and Foundation Engineering.
US Department of the Navy (1971). Design Manual — Soil Mechanics,
Foundations, and Earth Structures. Navfac DM-7. -
90

Vesic, A S (1975). Bearing Capacity of Shallow Foundations. Foundation


Engineering Handbook, edited by Winterkorn, H F and Fang, H Y, van
Nostrand Reinhold Co, New York, pp 402-417.
Werner, P W and Sundquist, K J (1949). On Hydrodynamic Earthquake
Effects, Transactions, American Geophysical Union, Vol. 30, No. 5,
October.
Westergaard, ll M (1933). Water Pressures on Dams During Earthquakes.
Transactions, ASCE, Vol. 98, pp 418-433.
White, R E (1974). Anchored Walls Adjacent to Vertical Rock Cuts.
Proceedings of the Conference on Diaphragm Walls and Anchorages, London,
pp 181-188.
Wood, J H (1973). Earthquake Induced Soil Pressures on Structures.
Report No. EERL 73-05, Earthquake Engineering Research Laboratory,
California Institute of Technology, Pasadena, California.
Wood, J H (1985). Earthquake Pressures on Monolithic Bridge Abutment
Walls, Report No. M1.85/3, Central Laboratories, Lower Hutt.
WORKS (1982). Site Investigation (Subsurface). COP 813/B Works and
Development Services Corporation (NZ) Ltd, Wellington. 84 p (currently
being revised).
WORKS (1988). Specification for the Manufacture and Erection of Concrete
Cribwalling. CD 209:1988. Works and Development Services Corporation
(NZ) Ltd, Wellington. 10p.
Wu, T H (1975). Retaining Walls, Foundation Engineering Handbook, Edited
by Winterkorn and Fang, Van Nostrand, Reinhold Co, New York.
1mE8E_
__A E

IfyQV_n__WSW
\ COWQCQ<EO
M__M_

Qo
Mmmmhgwwmba I4
_ "WW
SIoc_wyyy°E_un_y_ m'T4%
23
ZOQZ
|_<Q_n_:
EcQ:_ n_
_n_ _>w_<n_
o
mjg@Z_Z<FmE

\_\
CmmW®U E“\

“Wmflaw bf
“Em\T_
U_O®+D:WmQWDCMDL
®ELJ_N
Wy®lmOWD+U3©L®®U<LU>£
_£+O U
bssgq
_Uflw_O+®L@UWmCm£mzkW_H
+3m_@LOUW_2C+®3 £+L®©
*U®+UOWD
®£FQ>2
mm
LLOW
E®+W
®£+
OD_m
WC®UWQLQ
MM
\
Q;
__3\
U_DO3
Um+L®+@3
QOn0+®>M£
LDW®WO+_WUL®QU:_OCm
_ 3_O_E2_>
xOUm_°UD#:>

>_w/Ea____
x

*0LOECmCOm+M_DU>_+©mU_WmD@+m __“

%$P@Z

_
RAN KINE EARTH PRESSURE
COH ESIONLESS SOIL P P 1

CONSTANT BACKFILL SLOPE


4 f

A‘ +0-*°
1 7 I
‘-_;.,’-
4» .
£1 Z p=KAYz
.i‘. ‘I
~-:
.., 1‘.
,3

. PA
l H (DO

,90°—cl>
Al

l“|'I H
..’ PP¥m:"
- . ..:.L;’.'1_y
“A '
?5~
\,\ _< \'\ .
?';Z1“.'I;':.='-.‘~‘ 5,-. -; ' .'~. ».
A
FAILURE PLANES FOR P
RANKlNE'S ACTIVE STATE
PRESSURE ON VERT
PLANE A-A‘

The following equations require that the earth pressure act s at the slope of
the backfill.

ACT IVE PRESSURE

_ H2
PA
_ COS w (cos w — /cosz w — co )
KA cos w + /cosz m - co wmMN -age

“A = 45° + ¢/2 - %(e—m) where sin e = §%%-%-wrfh o< e <90 O


_.O - -450+
F0 P/2
I P T KA ' 1 + sin ¢ “A '
PAS SIVE PRESSURE
_
PP ' KP Y 2
Di

_ COS w (cos w + ¢cos2 m - cosz E)


KP cos w - /cosz m — cosz ¢

“P = 45° - P/2 + A (e—w)


Not e the angle between the failure plane s for the passive pressure case IS
90° + ¢_

For w ='0 Kb =l_+_$ll_3‘l


1 _ Sin ¢ up = 45 P_ P/2

I FIGURE 2.
RANKINE ACTIVE EARTH PRESSURE COEFFICIENTS
FOR COHESIONLESS SOIL WITH UNIFORM SLOPING
(C=0)
BACKFILL.

PRESSURES "ON A VERTICAL PLANE.

O~90'
E
I I

Ls I __2, ::||'
. . "='f-I--I
III.-I
I

O-80 I.-I.--I
I‘-II-II
‘- \- . . , -, ..
-
.-,'_- J ".¢~_'.-
'-.-.\.'.‘
.‘ ‘E
_____._9
_
8°.§ I I.-I I-I

I'_f_f§ l_'I III T’ l:

o-roe P. = Y H2 K.
KA K = Cos m cos I0 - ,/ (cos Z w '_ cos
A
' 2 11>)
cos w + ¢(cos7 w - cos? ¢)
P II Patti
. I)'IiI'jjI ,;.,.- 1?"
‘..'I'IlTL I
->-
-v
-—.
.y__ _.__,__.- .-L
4-
0'60 I: Q
: ' -- .-
.. _..._y._ __ .. ...,.._~-

.
l
..
l
.l.
*4-
-El
-. l:: -Il lizflgl l :-I
IIII
IIIIII
;."‘"lIl - ...

.... .. __.. .. . ... . .....- ..,... .._,,-»

"_. .:i1;t:t:*'ii.' : _ it 1
BEBE!‘ ._y.I . . . _
0'50
E Ii."
i"O!‘i
'“j.I'.5 i_‘_].If'_:_'_y .',.f' ‘I_I I I I *:q_f.:-.: =
iIIIII,
J§..f; .,§.Il .'
_.__.. ._ _ _._.t.i_._ -

:
E I .
-III-‘iii
EE|E§§i-. ----
. . .
I , -. -.- .__ -- --.._ ._ . _ I . E.
I. ._.......
~<I>I1Ij1
_ ,_._. _,..
.J,._.

I
'----4‘ ----,--
._ ._ ___ _. __

I '
-.---_
1
I _. .__ _.. . .. . _ ._ ,__ _.._._.

O-4O
Ii: 5 I 3"-;.'”'..i;;;"" 'i1i"";; ;_;.;;"1; 1 T.'_Ilj.. 1.1;-
IEE E " ,'3O ‘ _ _:i.1;;.":t.'::‘-:1"iii" in-:
P ‘ .. P I ‘ .1. :5: 1:. I
ii§iE§iiE‘I‘5 iii ) - .-. )._ _ _._.
__..___....__,,__,.> ._ L 5§E.,
__._.

0'30
is=s::::e;; - . - _ ;
:II::II II ii _'
- J I ii; -;13f;;ii1‘i;" "" EII

-:==-i g i »
.isIaii.ie ====i= lIIII.
IIIEIIII l=
I

'- IIIII=
IxI

__
I" II 44 I1'

i

I
IF
1)
._l_.
IIIfI

l
O-20
II
IIIIIIIIIII llIIII:21‘I I II IP‘IIl::IIIl'I I I .I1 :BII I I IILI- I I I IlIE::IIII
EEEII I I I I :
0° 5° lO° l5° 20° 525° 30‘
BACKFILL SLOPE h)° y-__>__.)
FIGURE 3
RANWNE EARTH PRESSURE
SOIL wrrn COHESION
HORIZONTAL enoump SURFAQE

Iv Ten§on zone
1
“I neglected
N O
.;§'_..__.,___-—-_.-.------- Z
I H
IA ' KAY (Z - _z<>)
I
45 9/2'45-‘ii/2 H
PA
:.:. -'1." -
§%m;%w&fi¢qfl&%<gmmfi§5;
'::‘:':-'1'-§"»‘,"J':.
_y
(
3 KAV(H-29
A
FAll..URE PLANES FOR PRESSURE °|§'LXE';T'§f‘A-.
RANKlNE'S ACTIVE STATE

Wafer pressure should also be added on ALA

ACTIVE PRESSURE

PA = %KA Y (H - ZO)2

: l — sin Q 2
KA 1+ Sin¢ Z9 : ‘V9 TEN (450 +

unll pressure al deplh Z below Top of wall, p = KA y(z,— go)

PASSIVE PRESSURE
Pp = %Kp Y H2 + 2cH/RE

K : i + sin Q
P l — sin ¢

unll pressure al depTh.z below Top of wall, p = Kp Y Z]+ 2c/E5

NOTE: The angle belween lhe failure planes for The passive case is 900 + ¢.

‘FIGURE 4-
COULOMB EARTH PRESSURE '
COHESIONLESS SOIL
CONSTANT BACKFILL SLOPE ,
assumed failure plane
I ‘ya + we / f l r alure
/‘_____aciua ‘I sur f ace

= l<A’Yzcos P

- ______._.
.-;.

fl' 5

cos I5
I; .- . ',-___-,_ i / °‘A , P‘
\#l>- I
FAILURE WEDGE FOR
ACTWE STATE ACTIVE PRESSURE ow
‘BACK OF WALL.
The following equafions give only an 'approximafe' solufion for fhe earfh
pressure when sfaflc equilibrium ls nof fully saflsfied. The deparfure
from an 'exacf' solufion is usually very small for The acfive pressure case
buf passive resisfance may be dangerously overesflmafed. ‘

ACTIVE PRESSURE
Hz cos2(¢-B)
PA = KA Y 7?‘ KA = cosz B cos (dffiiil + /éin( +6) Sin‘ ~m)i2
cos(6+B) cos(w—8) _

cof (aA-ml = -Tan (¢+6+B-w) + sec (¢+6+B-w) / fig: Egfig 25$ §¢fg§

If The pressure surface AB ls projecfed on To a verfical plane, The pressure


per unif of verfical dlsfance af a verflcal depfh, Z below The fop of fhe
wall is p‘ = KA Y2 .

for 6_= w and B = O, KA E Rankine's value.

PASSIVE PRESSURE

_ H2
PP‘KpY3'

cos2( +3) _
Kp = 2 I _ 2 Ii’) (Q-6) si i'l (Q+ Lu]
) 2
cos Bcos (6+B)[l coS(6+B) COS(w_B)

for 6 = w and B = O, Kp E Rankine's value

I FIGURE 5
COULOMB ACTIVE EARTH PRESSURE COEFFICIENTS
FOR COHESIONLESS SOIL WITH UNIFORM
SLOPING BACKFILL.

¢= 25° ‘(s=%g¢)
-1----_--.1---I-n
V

I §i%%i»<f“f"li;z;1$Wfif' : : : : gig
0'9 I 2 E‘: ,,i--:Eé§:=I
3; fig PA8 ‘ /2 W“ :5. II p
:II "II
Ii=:.
III
T=!!IIII ,
H
;q
- v.
fiE; IIIIIII
_ IIIIIIII
I I I::
I 0-8 - ;;5; .g;SEE y, IIIIIII| ail III‘
'5:::::a:' |lIlIli§-, .i:'E
‘ 1

7
%§ fiéi
i;¥E*Efi£E
¢~ .::::;i:
Iiiiiiii A "'
i, :::
‘ I I 4
k IEEEI
IHIII I

I-= 'A I
In. ' r: :§:I
0'7 '5-I: '
- IIIII
i‘i@:::'ii‘ I
I':" .i I
. i:I ' I I
II zIIId :'
IE5:-l I KI I IIIIIII
“A IIIIIIIIII II
IIII
=F5ii5*= ‘=l ::-: II

I CI6 léi:i!i?' E IIET

-:::::
I-Bil : we
::: ::: “IIII'
El ‘ -
\IH
IIIIII
III I I II
III I
j!!IltTIIII
IIII:
1"-zE"i lggi .::i
isswisi
IIEIIII gig] :::I

I Cys IIIII I
--::-..
III
OIIII» I I-I I I I.I :i:: ‘lllli. I I I.‘I ,¢:::a::r ::' -::-
I
:55
' IIII IIEIIIZI VI IIII

-::. ,@» f ~ ' =


fie 45:2"
1:: -:E:;a 2:’ :--:
- I_
:I IIII I IEI I
IIIIIIIIIIIII I I I I.I ‘\: : 55 III:I ‘IIEIII:
. -II-
- I *III
I.I
é: "” :III

0'4 IIIIZ IIII


L. IIIII
L‘:I I’ y '
_:::: 4 'LI'A\II I ......\|i;i|i| i|E.| |\ ' ?§g;5§'§§§$§§i::::-
-Eases
.$Hm“IIgiggéigi
5q. ,_I I I
II‘gags
IRIIII IIII§I::
‘IIIIIII

_ ifi-==ii-
fi fi fii E
lF"nfinui. i lu
_ 'IIIi:::§::L:IIi
\
5" ':IlI:E:

'ri' §I-EIEI =%§?5 ; ~ 5fi


4

_.' ‘ ‘E
‘i
:~ _ - V I‘:i:.. .I:::I::

_~A igEiE";ifii ! - A qr .fifiE ‘:I‘I


l
II II

O_2 ~;: 7, _¢7_,, I p_ g I I I:::iiii:,p


-20 -IO

; _ Bf‘°“T%ESF?FTEESE,,?_’,E 7' 1 all FEIEGURE


COULOMB ACTIVE EARTH PRESSURE COEFFICIENTS
FOR COI-IESIONLESS SOIL WITH UNIFORM
SLOPING BACKFILL. ::::::5! 4
¢= 50° (s=@3¢)
Q--1-i-in-—-Z-I

5aaa~s;la;l=
.+Boii¢mwF’“””qfl§m§r’
lg i*°’° ::: II'uI
IV
A“%KA -=2LE

Ii .
.__~
-,',|
';,'

a_
.‘..'~
-;‘>.‘;
:_. _n
5~.‘._~.I:_-_=-..:.-:=f:1-,-5.
__ ,-f.~_;‘;1-_'-‘;;;l.:_-'_'--1
—~— 55555555S ‘55iiiiill.
<J
-U0.,
H

55555555
Q-_

OJ
i.:::;z:
I Ell‘!

-I-'55::
:II:I:I:’
" :'.":::l

5"'5»555:5"-I I
I:'l‘:I III

555'555:' 5:55:55’
I I =i.-1: ‘IIIIIIIII i I:i pi
:::: ‘:::,’ 5' T I
ll.
II 4 I-55 II‘
I I I§ IIII7 _
III H ‘§F‘: II‘..

. :%:5
.::I M
III
II
.5.:::55: iI‘-E55
E1; Ei-sah
L!-

:5=E§§2iA
Enxe-:i : 55555555
III“:
:i‘~!I!:=i
IIIIIIIII:II
I IIIIIEI
iIII:'II El
III ,i;s-$2 ‘I-i‘=:
K:-I-L
IIIL I:
I

§"-*'.E" 1a.i5'a.5='f=‘=
.
. HI II II_ V
I B.II
I
liilunu

I:
I
kI‘:Ii-
V
IBIIII
I IIII

2&5“ ‘:5é§5.
5 :-
4 4

:==:g55E:‘
_4III
5.-5..-gg.-=»
- .:‘ . "7
::=.E::..;;, "II
2555555 A
T .-5'5::'.:5
...::-:::! I
I".-‘y ll
4
E
‘=.-.'i1'=*Ii“ 5.5.5
_4II
.4‘

.i
5. 4 - _l,| I:

E55!
i5‘-
55 .ii=i
EiiHE?F ~
"' IIIIIII “I_
III
;:::!!!
I -ilgjiaj
LE
III

-20
.=. A0 5.0 IQ
, _

20
» 3
BACKF LL SLOPE CL)° ‘ HGURE 7
_J- _ ~:::_:__

COULOMB ACTIVE EARTH PRESSURE COEFFICIENTS


FOR COHESIONLESS SOIL WITH UNIFORM
SLOPING BACKFILL.

: :55 O (8::
(P I 2’3(I)i
09 - c_444_-nu-44-44

»I I ::::::
::::
MO 5'l
I
C) G)
- '/2 KATHE-
5
H
5555
‘II
. J E‘ 2% »
°7 :‘;:" >:. '; Ii
<s-— — ~ — %> I 4.//j;R
,1 _ =,_Kefi*£; ;5>w~_--‘-
“0i ”;"

0-6 I 5 5 5 :-15
I
"'-';.. _
L_
IIIII -I I I _ VI
KA IIIII -Ii
IIII I
III ;§5 5F.5 5 5 5:. ,: III
I I I I .I
IM
II

III
EH‘ .: .::.:: :: :
O-5
IIIIIII I:I I I 5?
::::II:: lI 'I':1+
I:HI:II:I:IIIII
. ‘:::,
III
IIIIIIII IIII
-I IIIIIII
‘:::I4 A
IH

IIIII
III , II:II J

OY4 :: ..:i:5: IIIIEIIII II -: :- : - IIIIIIIII

EE5-555;;-;::::.::
55"55"f» 555::
VIII I IL..
.:55 l
III::IIii

'555:::: IIII %5
: : 5 : a : ~ . :55‘
i:::5
To-3
[
'
.=I

IiI_ ‘Ea-II
5:55:
:-I-4i
IIIIEEII
1I54IIIIII! IEEIIII -:§i
‘ :
. : -
::§:II:II1I:I
: ‘um
' : : . : . ' _ lI
5 : : 5 : i : :
IflIIIIII!l
-7
: :
I I I g!iu‘E!}EI
5 I
I

Il
I s‘I I lI I I: EI I I I.II
IRII
I
g!*‘!li§l5l l\ ‘:: i !§§E:%@Jq|:E!I§lg%g!§é:5 EE- EEII
I
L

Iiiif '.'|5l'iH"Ii'I
7
IIII!I! IIIII
I I!II:4I IIIIII!: -

: III InII:III
| _.IIIIII 5-
I, 5II!II::n
!=nIIIIIE.III:II!I
IIIIIIII
I!! I:--
IIEEIII
44 IIIIIII
41 -W . :
.4 '
5 : : 5 . :
4

5 . : : P
‘-. :A: 5 5T 5 5: :5:5 5.
-:--IIIII 7
‘' II:III\II
I BHI:I :I=R§iI I IIIIIIIII

0-"E2 ::: I I! i
4
4

55- ::5::5
III II :IIIII
I I II

IIIIIII il:::::::=‘I
II:::=nlII
III!- I--'
-I---III
4

1 :I!!=iII :.::§%i§
4

I‘- -=: ia~


.4 4
I
4
ii-II::=i __ 44

i5555::::
I
1

II
I:III-
O-I
-20 -d0 ,0 IO 2O 30 4O
BACKFHJ_ SLOPE
W w° I
I FIGURE e
COULOMB ACTIVE EARTH PRESSURE COEFf'lC\ENTS
FOR CO}-lESlONLESS_ sou WITH UNIFORM
SLOPING BACKFILL. -gig:
i=-=§
¢=I4O° I<8=%¢)
0 8 I‘-—"""‘ . . H,’/qmflm€/1;////flflwfif’ I II I
::-!:::’-
II ' II
i :I. I: III II I

.___;_§L'*; ° Q
... ‘E .
III I. _


g:
"%
1;
A" 2 A
8 :=::a=a IIIIII Ir Illfl
III III I III:II I
O 6
I
11
-..-
'#~
.
H iii:s:=. ;a*:ii .,
a
~
7 =r —
-'_l -; >,'; vu-
' ';~,; -_.1_|. .
. m-M Hwwf '
i
1""
:11“-'2 Pi" ".11
II ii;-l_’3:L\L-Li» L’,/’f, /cifEfjiT._l?IQ *
iii:
\ .......: :2 :. F
aiiaeigi rs E“‘ ' 5.5;’ "F
KA » ‘ =§;» i =,=s-‘=1
§§::.
: : : -: :!iiE-
: A. Izlglllfll
4 IIIIIIII
‘ IIIIIIII
II
II , :
§E§!:ii:::=':
§§§ii"'-
A
: 5i"FiI£f
l IIIIII
A
II
7. l II ‘:4 5
5 4 ' II
‘ 4

II IIII5
‘II III
II AI‘...7

:nIIIIIIg.-'
IIII!I!-
III II
III F5
IIIEIII m
.

I '*-II‘!
==§~~.
IiII
IIII
I
B 4
‘I
Iglz-III I! IE‘:-
I IIe.II
7 :II !
-IIIII II _ _‘flI.I.-
F
III I
IIII I W55-"
. e.‘‘i :a.:sa=%:~: : §: :i “'i‘:iE'*I‘i="'Z-E‘~J'i“ -:§'-= ‘-i=i.'5'§%i $=i.': .~‘:.:‘.i: :.:,§=!:"-::i5':"*~i= *
III
IIIIIIII
I-E: III III:
iII Qll!
..
IIIIIIIII
III‘I
!III!=nIIiiilllliii III
I_ =
In
':::::I':HEEIIIIIII
IIllnlll
I!::=il
IIIIIIIEE
_,
' ti‘-:::§i
I iig:
.5
I I.I!Ii
.41
7"
‘I
IIIII
IIII
\IIII

iiiiiilii II!!=!=-
'i!
fifiiiiiii i".: : : -H:. ': ::-
'iii=iiii
E:::E==ii
:IIIu!!!!
: :-:..'\.
I--._
- I

IIIIIIII
%=*=i=.i. :':§i:i:'
::::::::. :iiiiiiiI =ai=:5§i:'
sia :::ii5 :::::: :=' niifi .=i ii.-:"In :,=i5=.a:"§i=‘g?'§%!fi:§§‘*
-IIII
III‘
II

:':i“.':iii!
"i‘: ::II

-20 -10 " Z3:I §i: .:a:§:L-: : 5C3: I


BACKFILL SLOPE %0)° F;§]RE- 9
QOULOMB FAILURE PLANE FOR ACTIVE PRESSURE
COHESIONLESS sou. WITH UNIFORM SLOPING
BACKHLL.
61:2/3 ¢ ‘

=IiiiiF"iiiii=iii :iiiiiii“ifiiii
Iv ==#:§E:::::::§i§:::::i:55 ::::i:::i :::' . :-:::.
"T<>~ §:.5;;_g.§:-:.:::.:,:::::::. -::::5:::5:::::.,.. ....:i::::
I
A 0°,
iiiiiiiii =
::::-:5 :-.::-5;? ::::::::- -
_5° !::::i=!IIII
; _ - -ii :::::::- ::::::::: IIII IIIII I IIIIIIIII

.;w i....::=;i..
_- ’ _gIIi
-l5° ""!!!!§§lIi'i‘E:n -I ..
~ =~~I.. :::::::: :a:i:'
-. I~

S35 IHEEIEIH IIII "II


,

- .- -EIIIIIIIIIQEIIE
-IIIIIIIIIIi=:!IIII _‘ _ <§\\ IIIII IIIIIII
\§§QIIIIIIIIIIIIII:-
I II
"
::E:::E::iiii=iiii:E:§§= IIHIql!_‘ ‘ T “§“gIIiIi:.IIIIIII
IIIIIII
M -:.-5.: ..2:?"-"A’-':*g::?"~:2,.1?.~§‘ " " ' ';;;;j;.,
III
%§’é'fl'
:i
--::i
%%@%fi %fi&~§F“1
:II
II I|:l'jnIIHIII'
m¢,lIUl IU:
._'I f§gg?,iI I I I I
1
-""' ‘WW5: -I"'-1:" I-I-:~\QiI": :::‘-15::

IIIIIIII i-ii 1». = 4Q °» I..-


] "‘i'Ii::::::g§i§ :-:.-::. : . ::ii:: :'£55,
i:i:' ...‘... ..:::::.
»=-» ZFE-.'=55..i.=ii i ' as%si==
-200 IE!:I IIII I IIIII IIIII II II IIIIII
III
I
Q0 _ Iii! -- IIIIII IIIIIIII IIIIIII
~ _ 'i::*"I 1 - _ 5i.:::::5§::.-:a::. :'.::::.
iii
1.115;‘ : IIII
II
0- "'iiIIIi_.
_FP 5 E‘
=__!IIII!!
_ ,_I!!::::::-
I_¢°'
_
' "'5'“.
5.:
..
:IIIIIIiIIII:Ii It-._
I__-__ _‘
IIIIIII III.‘-:::_I II!!!II
-':-:r:':5.:::'II‘'-F.::;11fi": g;'sisa=-jg”EFF"§...::'"I-‘~Hr;

A
-
_
_ .
"
:
A ..
: . -
__
.

-:1
. - _
IIIIII IIIIII
IIIIII IIIIIIIII

“'I'" 1.P-" _ P IEIIIIIIIIIIIIIIIIEI


II:
I ‘I
=:!I!Ii_ Ini£i:i::sig:§I
PI 1
‘I
I::::: IIIIIII
:-5:::
.'?.=rifeE=
h
:. IIII:IIIi
IIIIII III
IIIIIIII I
Ifl'Jh 1m:
-. Q‘ —.%==5'=:::::--ihluggg;u=::::=i=i-::§~=":::=:==$!::-:: '-
—i ____ I-._ __. _n.i ‘.\_\

.:"--::'E=::::-="§§.:::-'§§i-::-i.==E-':!=ii*-§1=E'§§::i:'
IIIIIIIIIIIIIIIIIII IIIIII IIIIIIIIIIIIIIIIIII§I§§§§JIII
'

t
BACKFILL SLOPE 0)"
FIGURE IO
I

TRIAL WEDGE METHOD


COHESIONLESS SOIL
IRREGULAR GROUND SURFACE .
4
2 3 /
I

\ W
\
\
P \ ‘\-
-\

"E7\ , 0'0 4* AI°P_7 A_ ,,_>_;,


-e-711
PA \ Cb 0 ‘w‘;;\II
g RA
RA \\\\ W
\-

RPIS
A ‘~\_RA
S
FORCE TRIANGLE
FW°E"B‘§ES F’2i,T"A€T?$E ACTIVE u=uu_ LINES)
AN D PASSIVE STATES
PASSIVE ioorfsm

NOTES

1. The laferal earTh pressure is obfained by selecTing a number of


Trial failure planes and defermining corresponding values of PA (or
Pp). For The acfive pressure case, The maximum value of PA is
required and for The passive case, The minimum Pp is required. These
limifing values are obfained by inTerpoIaTing befween The values for
The wedges selecfed. '

2. Culmann's consfrucfion (figure I2) may be used To defermine The


maximum value of PA and crifical failure plane for cohesionless soils.

3. Laferal earTh pressure may be calculafed on any surface, or plane


Through The soil.

4. See clauses 3.3.1 To 3.3.5 for The direcTion of The earTh pressure.

5. See figure I6 for The poinf of applicafion of PA.

6. The Trial wedge mefhod may also be used for a level or consfanfly
sloping ground surface, in which case iT should yield The same resuIT
as Thaf given by Rankine's or Coulomb's equafions, whichever is
applicable.

IFIGURE ll. L
P

TRIAL WEDGE METHOD


COHESIONLESS SOIL
CULMANN'S CONSTRUCTION
(FOR STATIC EARTH
PRESSURE ONLY)
Failure Plane 2 3 ' 4 A

Pressure D 1 A ‘:\¢_$T\ G
\ /S u rfo ce C "V\\\@\ C_€\\\:\§»

B I V /L

\/
\ - ‘g’, W
\ V 4

\‘ /\ H I W2
W
W3

PROCEDURE

l. Draw line A-6 aT an angle of ¢O To The horizonfal for acTive pressure.

2. Draw Trial wedges ABCDI, ABCD2, efc. — a minimum of four will usually
suffice. ‘

3. Calculafe The weighTs of The wedges - say wl, w , eTc., and ploT These
To a suifable scale on A-G, each measured from A.

4. Through wl, w2, eTc., draw lines am an angle E, (see TexT for direcfion
of PA and hence 5), To inTersecT A-I, A-2, eTc., af H, J, eTc.

5. Draw a curve Through A, H, J, eTc.

6. PA is obfained by drawing a Tangenf To The curve, parallel To A—G To


Touch aT T. PA is The line W—T, To The same scale as wl, efc.

7. The failure plane is The line Through A and T.

‘FIGURE’ I2
TRIAL WEDGE METHOD SOIL WITH
COHESION IRREGULAR GROUND SURFACE

surface on which
pressure is
calculated. \ I
O

I __§m r——fi¥ NI
\ .|\>__.__ /) I /”/_;7
/ \ /
,w.
...~ /'_'
__--- \
\
1
/ /
>1 -'1‘ / / / ,’ Depih Of
// // “L/ // Terésion Zone ¢
4 / Q
H // ///// //c’xL 2° =7 fan(45°+-5-I
_\
3
-.1 .

SIIEPA ///’
\
//
\\ '_
\C\\\\\\\
\ RA) L
~ 4

3 O 74 ‘U

TRIAL WEDGES FOR


ACTIVE PRESSURE

/V
PA
FORCE POLYGON FOR
I
*
\

R TYPICAL wanes. 2

|
COMBINATION or
W FORCE POLYGONS
TO OBTAIN MAX.PA
NOTES Y/Ext '
l. The above example shows ‘Rankine's condiTions buf The same principle
applies for Coulomb's condiTions. ( Adhesion on The back of The wall is
ignored).

2. For direcTion of PA see figure I5 (Rankine's condiTions) or figure I6


(Coulomb's condiTions).

3. See figure I6 for poinT of applicafion.

T4. See figure I7 for resulfanf pressure diagram.

5. The Trial wedge mefhod may be used for a level or consfanfly sloping
ground surface.

‘FIGURE I3 A
TRIAL WEDGE METHOD LAYERED SOIL
AND POREWATER PRESSURES

B TRIAL a
TRIAL WEEGE v
WEDGE -’-"""_TT”§:
I X» LAYER ‘ ' PA!
' LAYER 2 .. ':.'Z-'
,'.~ ' ®

' /
, /
@
P P
// P/' A24’ W
/
/
_/Z
/ /
/

7zwm<";/ WMPA vi‘ °“


TRIAL FAILURE wspces PRESSURE DISTRIBUTION
on I-\—B

PM
I

Q! Q Q|.

\
\
\ W1
/ R|\
/ \

lw' /’
//4
\
\‘
PA! ’ °"R| TRIAL WEDGE I c|l

Q11: I /A X X
I W // A \ Q11: . PA2 PW PM
v l IIl/ Q
i \ IE2

WII2
—->l<—x
7'7
‘ X ///°" R‘l\\ wnl
I \\R
\3

/‘ I |/ Q) \ \ wmz.
Al '.<'.," / U‘ RI
U
I. ® /// Ul 2
A czla czyz
c|I
/
2”’ N LAYER I ON LAYER2
PAQTPW/.//\ 0

7H'\WX\\__' U2 Ra
TRIAL WEDGE I;

NOTE. U, and U2 are the resultant porewater pressures y


on the failure wedges. I1
L‘ i 2H i’_

/I C

A‘ /

I}
*f=?,, Xi
E v
("I
PROCEDURE

1. Draw a line from The po inf where the ground


surface infersecrs the back of The wall (B) To a
<-...’;
;% painf on The ground sur face located af a disfance
H‘ ' ~ equal To 2H' from B.

2, The pressure on A—A' ma y be assumed to acr


1 I
parallel with 1-his line
1
, .
I , l _ . l_..___
_.,._.
L 1 .- »
1,1
'I=.-r.1
A

APPROXIMATE METHOD FOR DI RECTION


OF RANKINE EARTH PRES SURE
4/

4
OUTER FAILURE SURFACE

INNER FAILURE
SURFACE

PA
F0155 WITH WALLi'i_
l +5
VERTICAL VIRTUAL BACK
OF WALL

v .— '1,‘
A
I ) RANKINE <1]
5 =9)
i
B __

MOVES WITH WALL


\ SLOPING VIRTUAL BACK
OF WALL

\ >“' NOTE FOR Acflvé CONOITIONS

= 1121; > ..\ 4 I (I) If line AB do es not Intersect the wall, Rankine's
1-, '-. .- conditions ap ply.
vs v-7'
A If Ilne A8 do es Intersect the wall, Coulomb's
||) COULOMB conditions ap ply.
8 = p
(2) 7 =I(90 -0' )" HI: - u) where sln €='ii—'l£
sin 9'
B_) GEOMETRIC couomons FOR RANKINE AN D
COULOMB METHODS

I FIGURE I5 L
\\
\\
It;:3
_COmWQ_ _O
E>V320C22:
ON UwB_mw
__/+0E8
“E8_o_ct;F_?8£QC:
oi66
mgmzmi
_82_8H"ix_<.UQCEBO+LIN+53
U\/®®LJ_
0+
‘X
_‘___§Ug
M>C®CU_Q
_WmLD_®o_w<_X
X|<
m\_
l__
6&0;
'HCWfl_®_'WC_OAOWU_®L£oU<L_O“:<ou
m_ (_Q 365
2_o_W
_©+O+

2+
Ea
2;
E____p_m
C®®3@933
®LJ_
gt‘
0%;
‘I+g
<+53£m3OL£
2'_®__8gX
mm
A2
UCQ
2:
M380
Q9:
@591
U_N
_N
flgw
£m:ttw
Om
U6
2+
_£O_§
CU®_m%
<n_
O]_‘
_OQ
®CO_mI
UOE(_L£J_‘W ®_+__
®m_
'\_ ‘_m‘_m®U®(_ ‘n©_ ‘_ U_mU
to
m>_____u<
mmn:
2__8;U®HWHQE
9:
U5N
m_CML263
mwmm
@:89%
W_LO:UW®_®>MULQ 9|#_8_h__
A
<_8_t£9_9_£
_<m<
*0_m_U
J5+
QC:
9;
®m_U®3
O‘®>_'U®_mL®+E

NM
£g_OL_£
LI96Wm>WtCL<n_
g$WOQ
U®fl__X
w_c_J
&\
_l
E
\
l
\v

\
\\\\\\“

_ +58_N
H(
\\\M\ U225Q
‘L3%/vq
%‘\
/*0
\gt;B
\\ad
M___m_O
68;
\l\l“|_
_
_<
“U8
_ ‘A2
6mg
®U©t_ _W

\\\\\\
\

4
m
<_ H
gag?
co$\
_ EH§_fi_

w___H
\\

_<\\

IO“:
_U_UI_D_WWU””_<FmW_&
\
\\
__< \Aj <n_

m
‘\
W
rm__°_m€_2Q
POINT OF APPLICATION
OF RESULTANT PRESSURE
AND PRESSURE DISTRIBUTION

surcharge

sf: I Ts 2°

- I
B ;_-= .---- '-""' “T-
4 | /J““’ Tin’ B
,/ /
4_ ///
//I/// ///
// ‘ya
K |
/ 3

’ // hi P —a
3 / I
§a$%a%':3 // 2/ ,"".\5 P“ 1”‘ M PA
I1» I F;-P
l _ I
Y h,
A_W [_
A A
TRIAL WEDGES PRESSURE ON A-B

Use when The ground surface is very irregular or when a non-uniform


surcharge is carried.

PROCEDURE

1. Subdivide The line A-4 inTo abouT 4 equal parTs hl (below The depTh 20
of Tension cracking).

2. CompuTe The acTive earTh pressures P1, P , P3, eTc., as if each of The
poinTs I, 2, 3, eTc., were The base of The wall. The Trial wedge
meThod is used for each compuTaTi0n.

3. DeTermine The pressure disTribuTion by working down from poinT 4. A


linear variaTion of pressure may be assumed beTween The poinTs where
pressure has been calculaTed.

4. DeTermine The elevaTi0n of The cenTroid of The pressure diagram, y.


This is The approximaTe elevaTion of The poinT of applicaTion of The
resulTanT earTh pressure, PA.

Nb. WATER Foizces MU5'I' be C.ON€>lDEl2ED SEPARATELY.

IFIGURE I7
T\5“\31fiy~0.?i -0.6 _0.5 -0.4 -0.3 -0.2 -0.l _
REDLICTION FACTOR, R,
IO
15
.-22@..L- _9§Z.._;9q§"_.9294L912
.901 ;.93a .907 .881
I; 898 . 881
830 .803
OI KP I-TOR VAHl()LIS-1
29 .939 i. 901 ._;§92.";8 24 787 752 *.?16
RATIOS OI-T 5/(1) 25
3O ‘]87§T. Bil
860 7 59 .574
<i>:>_ -1620
;§§Z5-'a
Q@§f
35 752 603 .-H7
/i-‘I-__—i-pk

ill) (‘:82 512 i?§_¢;2l2_ ’

ll"lll@ A’:
45 .718 .600 .500 .KTZ'f3s9 . 276 .22I I74
P : L_ I
9°-OI I _ I I w/o'I.1.o k)Ifl:0D-L
I
333 I I Z
I A
T we §oa——
u/I0-» 6j ~ 7%
60.0 I - l‘ * '

. -... I
“’ld:+O.Z
50.0w _ ‘ V -‘

It
I‘)/D =0
__ _ ~w /-- FAILURE _
30.0
' . ‘
.\/{>3-I / SURFACE
En};
‘i;I
ET;
.\

20-0 \ ‘win =-0.2


3 \1-— LO RITHMIC
__I
\ 5 ii‘I'-
4»: __§'*'(<I: I\ \'t ‘J/49: -O-L

_. PASSILE PRESSURE ;_
\ V .
FT ‘Q;K‘!
R: ‘K
$0.0Q0
8.6 "-'3 =k;YH'/2 z R. =F;ws6 ; B =Pp§in6 ‘

T
7.0
5.0
" NOTE : CURVES SHOWN ARE FOR 5/I! ‘\§§\_\ r
Y
4 V
4)/0 : -O-5
EXAMPLE: 6 = 25'; we = -0.2
5.0 F_-
VEPRESSURE
PASS
CENTOF
COEFF
L.O KP=R[Ky FOR 5I6=-1 I
‘s \‘&\\
' ~ T\
R:O.711 T
3.0 IK7 FOR 6/6:-1
U/0 : -0.8
Ky=O.7I1I 3-62=

\ \ \‘mH. IM
2.0

.1\\\\§\\.\ \ \\\\\\55».\ ~\ _.\u\\ \,\ \ .


; A.”/0=-0.9
_ .___. _.. .._lu_i —»r'-"\"__I:.__ __. I

1.0
1 ||\ I I= Pas Z0
0.9
0.8
[17
0.5
I KZ W anIlluIIllal
III!WWWIrm w
1. In
I:. I-no =-1-0
I
' ‘u=0
5:6
Jul/0=v1.0
II/7

m =MP/ \.I¢flf\I/71'!y/mu!Ill/IIIIII/ll‘IIFM/IIIflllW‘T!
>
0.5

0.4

0. 3
I.
E“
4‘g.
I/1'
_l\
I/Z‘

URE
FACE
Y
@=§Q8

-——— LOGARITHMIC SPIRAL "‘


IIIW”/I :0Q5

- 5 -~—— Q =xAy H —~ Mg"


-J _....“T | '”iT"@i
0.2
ACTIVE PRESSURE I”I
PCOEFFCENTOF
ACTVE
RES UREKA

0.1 E~. --.ff


§\U‘ n
C78 L.-I

QQ W
E\9

E \ '5
O 10 20 3) LO l\ U1

ANGLE OF SHEARING RESlSTANCE,p', DEGREES (CaquoI a. Kerisel, may

EARTH PRESSURE CUEFFICENTS-SLOPING anuunn FIGURE I6


-0.7 -0.6 -0.5 ~0.4‘-0.3 ~o.'2 -0.1 0.0
IO .978 .962 _.9l&6 .929 .912 . 898 .881 8611
REDUCTION FACTOR . R. I5 .961 .934 .907 . 851:
. 88I . 830 . 803 775
20 .939 .901 . 862 . 8214 . 787 . 752 .7I6 678
OF Kp FOR VARIOLJS ’ 25
.912 . 860 .808 . 759 . 7II . 666 .620 574
RATIOS OF <»./0 30 ,.s78 .BlI ‘.746 . 686 . 627 S74 .520 . .467
V 35 .836 .752 .674 .603 .536 . 475 .LlI7 362
40 $.78; .682 .592 . 512 . 439 7. 375 .3I6 262
1&5 I.7I8 .600 . 500 . 4111 . 339 . g"/6 .221 I74

51 I I I I ..
‘IL

III:
9 ....-

8 I~__1I‘
I
”%;I3:*"' Y
, 4- FAILURE
SURFACE
72.

5
_
Ag; -
In '

@.=~<.v~
'
-..
II’,/'3.
/
/

’\L00ARnHM|c I
SPIRAL - .1!) F
I” .
I

I I I ' _
PASSIVE PRESSURE
' .
5 "T 1; =K|,YH'I2;R,=PPcos$; P,=F;,sine r'v
K,
PvsRES URE *—- NOTE: cuavss, SHOWN ARE §.. ‘II I

FOR we = -1

PASS
ENTOF
L — EXAMPLE L @:30';#:-10; $l¢:-.5
__I(P:RII(p FOR 5/0:-II
_R:0-B11
___lKyFOR 6/o=-1)=a.2 '
_ Se A
Y Y
I . 1 >9
o

4/.,. _.
K|,:0.8‘I1 1 8.2 = B-65

C
COEFF

‘\w.\‘.
\ ‘w
.i
I lir'i|'
Z
k"
1.s'.o/i\
PASSIVE
K.§. .K w ,_.
‘.

-—- ZONE
//-_ FAILURE SURFACE
_/

I
I


\.\\
. - 4
1‘:
L
5
‘Y

~ . I
\'\

.1
la.
ZPgm 31"“8/ ¥ LOGARITHMIC SPIRAL
6A:KAY H

ACTIVE PRESSURE
KA .
.Q
_ __ r;=i<_AxH'/2;a.=P,coss; P1-=P@sin6
0A
E.
UR
ESS M we»
A Mil _ I -
—< ~_ . E.
7-. /
._ .__._

'lI/ll WI/IM ".


> _
0-5 I
LII
_ ACTIVE UH!‘
<

EPR
II I I ~—s/¢>=oII3“3°
/:5"
__

MT’ ' " €


__ OJ‘ ZONE ---—5I¢:I _ '

ET AIIHI
TV
0.3
M Hfli 1}n:‘15'

‘HMQ E
OFAC
0.2
IJ_ ‘__> $/6-0

~——- ‘ 212: 2.)/»=@


ENT

FFC
COE
I I I‘ 8. F*~
I
3
9 n| (D_~
g‘¢\

I I I I I
iii

I
10ANGLE OF SHEZAJRING RESISTANCE, ¢ , DEGREE? l CaquLoSt a Kerisel,19L8)
I
8%/ 1-
;‘I—
91¢-1);.--15‘
W I’; 0 _

EARTH PRESSURE BflEFFlClENTS—SLDP|NG -WALL FIGURE I5


q : 57 RN/rn LATERAL EARTH PRESSURE The kNIm2
I ‘IO 2O 30 LO 50 CRITICAL OEFTHS AND EARTH PRESSURE VALUES
' _I LI | I I I I I .

0I-2 F?
CRITICAL MAXIMUM HORIZONTAL
COMPACTING MACHINE DEPTH EARTH PRESSURE
Pan 1¢I"'l °'hoMA)$.IkNIm2I I
\\‘< ‘rho = Kodv 1o.z1 suoorn WHEEL ROLLER 0.56 20.0
S \
In
\ *3.3i VIBRATORY ROLLER I O-52 19.0 I

H’
1.5
*1./.1 VIBRATORY ROLLER 0.35 12.5 F
Z,
PT
1.oo_ kg VIBRATORY PLATE 1s.o
OE
2.0
IIIIIIHIIIIII‘. ‘T10:

IEARTH PRESSURE
C OMPACTOR
0.1.5

DUE TO WEIGHT OF ‘I20 kg VIBRATORY PLATE 0.32 11.5


I BACKFILL I ' COMPACTOR
‘ho nAx_= 20 kn/mj I
NOTE, DIAGRAM DRAWN FOR 10.21 SMOOTH WHEEL ROLLER
on FILL, ab’ = so‘, a’ =10kNIm3
* EFFECTIVE WEIGHT OF VIBRATORY ROLLERS ASSUMED TO BE
TWICE TOTAL STATIC WEIGHT.

(Al commcnou AGAINST UNYIELDING WALLS IBROMS,l971)


u

:_i’ __ __ ...
N -\‘ Z_ I ‘ .
_ C " \ /
\ [T .._..' - ZPKA
\ O"hm= -Ta‘,-i \ \.//\ >
dI""Y"i'
/ /*4 VALUES FOR
LLEYEL
___,¢r =_._A
| I
_ \LL \l > SUCCESSIVE
\/ -
P 711 I / 7<\ coumzrsn LAYERS
__ _.-.0‘), = “~31
A />1 ='nn~¢'mn
‘-—- RESULTANT PRESSURE
DISTRIBUTION
1% ~< /.\ >—Locus oF°"hm
I
|
ii I / >_>-¢Ihm:F--_;T-Q (Z>Z¢I

—.
____€*§$
_L';7Z<\__ FOR Z>h¢°'h=KAIZI
DEPTHE
LQIELIZELUWFNALFILL f
DEPTHBELOWFINALFL
L.. _. Ifl /<\r _ I
HORIIONTAL EARTH PRESSURE HORIZONTAL EARTH PRESSURE
Id) SHOWS INFLUENCE OF COMPACTING Ibl SHOWS INFLUENCE OF SUCCESSIVELY
SURFACE LAYER OF FILL WHICH WAS COMPACTING LAYERS o|= son. escmums |
PLACED WITHGIT CQIPACTIUI. AT BASE QF WALL.

uIi<\%‘?

*1 6-Ilm - MAXIMUM VALUE OF HORIZONTAL STRESS SUSTAINED


AFTER comrmcrxou. ,

KAI-1-z\:%
Z

\ WHERE P : EQUIVALENT LINE LOAD ME TO


ROLLER. FOR VIBRATORY ROLLERS
éhm=F%—' I In CALCULATE p USING AN '
-I-%—%i-—-> W? EQUIVALENT WEIGHT EQUAL T0
l he =""i OEAOWEIGHT CF ROLLER PLUS
‘-__.._.._L KA CENTRIFUGAL FORCE INDUCED
DEPTHBELOW
F
LEVEL
LL BY ROLLER VIBRATING
MECHANISM.
°"h=K Th
I I
HORIZONTAL EARTH PRESSURE
lcI sHows PROPOSED DESIGN PRESSURE
DlAGRAM_ I
(B) COMPACTION PRESSURES -I DESIGN DATA IlNGOLD,l979I

EARTH PRESSURE DUE TU UUMPACTIUN FIGURE '20


Sloping Backfill
SOIL FRICTION, ¢= 30°- WALL FRICTION, s=¢
4
' I
35 .- .-_
. -. - _,

3
@259 __-.-

9
O

0)
\J

O
\O

A
KAE/C
O

4*- ‘
J_

_‘_
I I I I I
O 0.2 O4

EARTHQUAKE ACCELERATION COEFF. C(O) .

Sloping B0cI<fIII
ED] so||_ FRICTION, ¢= 35°; WALL FRICTION, s=5z§
I
I
._ .- _.

-i

qkjoo

4?
I9
o
{'3 \O
AKAE/C(O)

/2 l
64
TI I I I I I
O 0.2 O4

EARTHQUAKE ACCELERATION COEFF. C(O)

MONONOBE -OKABE ACTIVE EARTHQUAKE


EARTH PRESSURE INCREMENTS FOR I FIGURE 29
VERTICAL WALLS
/ //K\@\\Y//
/
/
4-—fuIIure pIc1ne
c I0) ww PM /
OI‘ NI/I/W

‘ /
WW /

-——:>—-- I / j°<AE
PPE
B
Sb

where, N.g = Threshold uccelerufion for


oufwurd movemenf

STABILITY ANALYSIS OF RETAINING WALL FIGURE 22


UNDER EARTHQUAKE LOADING
0'0

I
\ ()4 _
’“ 03 2 - dep’rh below
Poissons rufio, 1/'=0'2 0-A -5 reIuIned ground
0-6 - surfI1ce.Isee B
below).
o-e ;-

I
1.0 I. I I I L
0'0 0'25 0'50 0'75 I00 I25 1'50

BIL)
umzw

A. EurIh Pressure DisTribu’rion on Smoofh Rigid Wall


for differenf Poisson's Ratios of the Soil.

P(2)=I~5C(0Ib’H
2
Apos
H
APOE = XH2

-58H

P(z)=O-SCIOIKHZ

B. Approximufe, Design E[1I"‘Ih Pressure Disfribufion.

EARTHQUAKE INCREMENT OF PRESSURE ON


RIGID RETAINING WALL I HGURE 23
rnenf
heghf

eurfhquake
Effecfve for DFESSUFE
ITCFG

EQUIVALENT WALL HEIGHT FOR ESTIMATING


EARTHQUAKE EARTH PRESSURES FOR WALL WITH FIGURE 24
SLOPING BACKFILL
H APE APE: O-7‘5CI0)b’H2
H
2
PIz)=O-75C(OI2SH

A. Sfiff Wall

H AP APAE = %AKAE‘6H2
H AE Where AKAE = KAE_KA

T
P(ZI=(KAE_KAIXI"I

B. Flexible Wall (from Mononobe-Okabe Theory)

EARTHQUAKE INCREMENT OF EARTH PRESSURE FIGURE 25


STIFF AND FLEXIBLE WALLS
MONONOBE - OKABE EARTHQUAKE
EARTH PRESSURE I ,
COHESIONLESS SOIL
CONSTANT BACKFILL SLOPE

5 +0?’
O =P,I+AFIE

\i\N\
\\\
2

I2’ rs;

44

-\ ‘B+s

FAILURE PLANE FOR ACTIVE EARTHQUAKE


EARTHQUAKE LOADING PRESSURE ON A-B

(%IéSee secfion 1+5 for POIFII’)


ACTIVE PRESSURE Of APDIiC@ii°"-
PAE = % KAI: Y H2
' cos2(<I>-B-6) i Z

+
6 = Tan"1 C(O) WHERE Cl0)= The design seismic coefficienf from Secfion ii-2.

coT (oIAE—w) = -Tan (¢+<5+B—w) + sec (¢+<S+B—w)\/E32 _

NOTES _

1. The above equaTions are based on a resoIuTion of The forces acTing


on a wedge of soil. The effecT of an earThquake is represenTed
by a sTaTic horizonTaI force equal To The design seismic
coeTficienT Times The weighT of The wedge. ‘
2. Where The earThquaI<e earTh pressure is caIcuIaTed on a ver“I‘icaI
plane Through The rear of The heel, B is zero and 6 is equal To
4. For The deTerminaTion of The poinT of appl icaTion of PAE, The ToTaI
acTive earThquake pressure is divided inTo Two componenTs, PA
(from sTaTic loading) and The dynamic incr-emenT, APAE = PAE — PA.
PA and APAE are applied af I/3H up The wall. This assumes a flexible wall and
sufficiem‘ movemeni To give acfive pressures. I

5. WHEIZE THE EARTHQUAKE EARTH PREE>5uIZ.=_-_ I5 CAL..C.UL-ATEQ


DIRECTLY ON THE WALL, THE. Ar~J¢II_E. CF‘ WAL.I._
FRICTION MAY E>E. TAKEN A5 5=-93¢ 2,6
Horizontol Backfill
WALL FRICTION, s =¢
2.4 C I
2.3 T I
2.2
2.1
2
1 9
1 8 T
1 7
1.6
/\

\./ 1.5
1.4

my_
I 3

if /
I

0.9
0.8
\\\6*
TI\
07 I:
06 ‘Q
05
O4']
0
'
~»T'T

\g
°\ 0.2
\
s_ L_.

EARTHQUAKE ACCELERATION COEFF.


as

C(O)
04
im_

MONONOBE—OKABE ACTIVE EARTHQUAKE


EARTH PRESSURE INCREMENTS FOR FIGURE 27
VERTICAL WALLS
Sloping Bockfill
4 SOIL FRICTION, ¢= 20°; WALL FRICTION, s= I25


55

-1-.-.

3 _ _
0

25*
‘ \?
1

(0) Q

AKAE/C
Z
1.5 7 2 / '
6
I

O5

O ‘I I I I I I
Q 0.2 O4-

EARTHQUAKE ACCELERATION COEFF. C(O)

Sloping Backfill
4 E
SOIL FRICTION, ¢ = 25°; WALL FRICTION, E:
I
I6
I I
__.

35 -1-Z-.

(UQEOQ
3 -1?‘
O

9
25

Q
C(O)
2

AI<AE/

L5’ {/ 6
I

O5

O TI I I I I I
Q 0.2 Q4

EARTHQUAKE ACCELERATION COEFF. C(O)

MONONOBE- OKABE ACTIVE EARTHQUAKE


EARTH PRESSURE INCREMENTS FOR I FIGURE 28
VERTICAL WALLS
Sloping Backfill
soIL FRICTION, ¢= 30°- WALL FRICTION, s=¢
4
I I
35 .- .-_
. -. - _,

3
<o.85° __-.-

I?
O

0)
\J

O
\O

A
KAE/C
O

4*- ‘
J_

_I_
I I I I I
O 0.2 O4

EARTHQUAKE ACCELERATION COEFF. C(O) .

Sloping Bockfill
ED] soIL FRICTION, ¢= 35°; WALL FRICTION, s=5z§
I
I
._ .- _.

-i

Fdajoo

Ii
I9
o
{'3 \O
AKAE/C(O)

/2 l
64
TI I I I I I
O 0.2 O4

EARTHQUAKE ACCELERATION COEFF. C(O)

MONONOBE -OKABE ACTIVE EARTHQUAKE


EARTH PRESSURE INCREMENTS FOR I FIGURE 29
VERTICAL WALLS
Sloping Backfill I
SOIL FRICTION, §2§= 40°; WALL FRICTION, 5=¢
4
I I
I I
35

.5’
2.5 G _ U 7
0)
\./
Q7’
2 O

LO D .,
AKAE/c x‘? \o
1.5

I
\ II Lee)‘
i
05- T
0 ‘I I I I I I
O 0.2 O4

EARTHQUAKE ACCELERATION COEFF. C(O)

MONONOBE—OKABE ACTIVE EARTHQUAKE


EARTH PRESSURE INCREMENTS FOR I FIGURE 30
VERTICAL WALLS
I'IOl’|ZOI’ILCII Backfill
WALL FRICTION s= Z/2>¢5
2.4
2.3
2.2
2.1
- _.,-._.-_ _.
2
1.9
1.8
1.7
1.6
/\

I.5
1.4
1.3
AKAE/C(O 1.2
I I T / /
I _ // /
0.9
0.8
O.7
0.6
0.5
/’ ///4’-
7 _
_‘J—:
i
_
ti
lIiII\
L
O4 A_LI4I_I
_L_
_

EARTHQUAKE ACCELERATION COEI-‘F C(O)

MONONOBE—OKABE ACTIVE EARTHQUAKE


EARTH PRESSURE INCREMENTS FOR FIGURE 31
VERTICAL WALLS
Slo Pl'n 9 Backfill
SOIL FRICTION, = 20°; WALL FRICTION, s= =/5¢
4
I I ¢
I
-_-—
35

_-_. -_.
3 _

O ¢--

3’
2.5 A G A -—-¢@_-
O
0
\
(0)
2 _ ii-

AKAE/C
1.5 -
O
O
4

O5

91 I I I
O 0.2 O4-

EARTHQUAKE ACCELERATION COEFF. C(O)

Sloping Backfill
SOIL FRICTION, ¢*= 25°; WALL FRICTION, s= '-/2565
4 I fi
-i-_~

35

-~_.
3 0 4--_-_
_ -_.~_-.

/I9
2.5
9" E?

9
_.__
‘<7
2 __
/C(O
I O
\o
AKAE

0
O
137
1 _ ___/ J7 ‘ AI

O5

0 W I I I I
O 0.2 O4

EARTHQUAKE ACCELERATION COEFF. C(O)

MONONOBE—OKABE ACTIVE EARTHQUAKE


EARTH PRESSURE INCREMENTS FOR I=IGuRE 32
VERTICAL WALLS I
Sloping Backfill
4 I
SOIL FRICTION, ¢= 30°;
I
WALL FRICTTON, 5 =1’/5
I
95
.---

3.5
-—-_.- -_.- _

- s-_

3 O

25
<o.eS
—_

I5
2
<0 I5
AKAE/c(o)
I.5 F

0.5 W

O T I I I I I
Q 0.2 Q4

EARTHQUAKE ACCELERATION COEFF. C(O)

@ Sloping Backfill
soIL FRICTION, ;z4= 35°; WALL FRICTION, a= Z/392’
4 I I

¢- _-_
._ _-_.
35
- I- I-.

3 _ ._ .-_i.
._- _. _. -

_._€-.

25
@.__.,O° ___1.

/\

\/ A7’
2
'65 I\
<5
AI0<AE/c
1.5 T

i _ i W I_ ‘ O‘2I
I 7 _l__ _ __I_
f

_ r I
O5" I I

O TI IT I I I I
O 0.2 O 4-

EARTHQUAKE ACCELERATION COEFF. C(O)

MONONOBE-OKABE PASSIVE EARTHQUAKE


EARTH PRESSURE INCREMENTS FOR FIGURE 33
VERTICAL WALLS
Sloping Backfill
SOIL FRICTION, ¢= 40°; WALL FRICTION, s= 1/3,¢
4-
I I I
I
55 I
I -_. - ¢_-. -i_._
3 .
T, §—*_ _-.1- -_

2.5 I 0
(0)
2 £0 I?’ I5
O

AKAE/c
\‘° 4
1.5 e A, I ,

I _I ‘I, I I I Q; I II \O O:
_I_
I iI
O5 I' r

O ‘I I I I I I
O 0.2 O4

EARTHQUAKE ACCELERATION COEFF. C(O)

MONONOBE—OKABE ACTIVE EARTHQUAKE I


EARTH PRESSURE INoREIvIENTs FOR’ FIGURE 34
VERTICAL WALLS I I .
Horizontal Backfill, Woll Siope = —14°
WALL FRICTION

T 5
21/

AKAE/C(O)
,
"'L4»
/K /

\\ \
',,4 / ‘47
\
_ --1)<>/1
\ K
15 A

__
_
_
‘A:
41>

_,_ ~T— _>

__|_ 1
‘ é¢_;°;I

‘if
4.-

I T
__‘T__

T I I I
O 0.2 0.4

EARTHQUAKE ACCELERATION COEFF. C(O)

MONONOBE—OKABE ACTIVE EARTHQUAKE


EARTH PRESSURE INCREMENTS FOR I FIGURE 35
NON—VERTlCAL WALLS
Slopin BocI<I‘II, WoII Slo e = —I4°
QSOIL FRICTION, ¢= 20°; WALL FRETION, S=7-/5Q/
4 I

I I I

as
I
3 L

25
<0:5.
O
9 I
(O)
2 I
AKA:-:/c 6
I5

1 I "

O5

9 ‘I I I I I I
O 0.2 0.4-

EARTHQUAKE ACCELERATION COEFF. c(o)


Sloping E3ocI<I‘II, \/VoII Slope = —I4°
SOIL FRICTION, ,d= 25°; WALL FRICTION, s=l/3,325
I I

__-
I
I
-
Q-4-F~

25A 41:26, .-_»

I
go) I
I

2
\O
AI<AI—:/c(o)

O5 ‘

O "I I I I I I
O 0.2 O 4

EARTHQUAKE ACCELERATION COEFF. C(O)

MONONOBE—OKABE ACTIVE EARTHQUAKE


EARTH PRESSURE INCREMENTS FOR FIGURE 36
NON— VERTICAL WALLS
Sloping Bocdll, I/\/oll Slope = —l4°
SOIL FRICTION, ;z$= 30°; WALL FRICTION, s=’-/355
4 I

55 ‘

__-1‘-o'- p

3* I

25 , I
(0)
90 ’/
'9
AKAE/C
2

1.5 - I a
1 _ /) 1/// IE},
T I I
O5 "
A I

O TI I I I I I
O 0.2 0.4-

EARTI-IQUAKE ACCELERATION COEFFA C(O)

[E Sloping Bockfll, \/Voll Slope = —l4°


SOIL FRICTION, ;2§= 35°; WALL FRICTION, 5==/M5
4- I I I
l I
I ‘ I
3 5 I ‘ I I

3 __.-@-_i -_ I I
I l
I
25 I I
:9
5’
’ 1I
2 I

AKAE/c(o)
'3’
I .5 , ,

I’ Z '_l O» J
»@" ‘
I

05 1 I T

O "I I I I I I
O 0.2 O4

EARTHQUAKE ACCELERATION COEFF. C(O)

MONONOBE -OKABE ACTIVE EARTHQUAKE


EARTH PRESSURE INCREMENTS FOR FIGURE 37
NON-VERTICAL WALLS
Sloping Bockfll, Woll Slope = —I4°
SOIL FRICTION, }2I= 4-0°; WALL FRICTION, S= Z/5}-Zf
~ I I I
4 _

35 I I I I I

3 I j I

25 I ' I—

2 I A ‘I
E/C(O) l
0 0°) O I I O

AKA 5 /"70 W =19 \‘>


I. *1 ’ A
I I I9

~/,;// I/ K
I I/C5’

II
_.I
' -‘
Q5 :::‘, 1%; L F ‘

O ‘I I I I I I
O 0.2 0.4

EARTHQUAKE ACCELERATION com. c(o)

MONONOBE-OKABE ACTIVE EARTHQUAKE


EARTH PRESSURE INCREMENTS FOR FIGURE 38
NON— VERTICAL WALLS
SOIL WITH OR WITHOUT COHESION
IRREGULAR GROUND SURFACE
2 3 4 5mm
////7T“Yfl@\ I \\\
I
4 I

CIOIW /
.' II W

541"? 5/’, / 5
I
. X I I
v‘ -

il.§-{Is R
. ‘P
\
FORCES ACTING ow I I4
TRIAL WEDGES FOR I I
EARTHQUAKE A Mox.PAE
LOADING.
PAE I 3
FORCE POLYGON I
FOR TYPICAL
R W WEDGE \ 2

l
/ F1 § ii COMBINATION OF
<-~ FORCE POLYGONS
vfl To OBTAIN MAX.
NOTES 4 PNE-
1. The above example is drawn for Rankine's condiTions bu+ The principle
applies also for Coulomb's condiTions.

2. For direcTion of PAE see Figure I5 (Rankine's condiTions) or Figure I6


(Coulomb's condiTions).

3. For cons+rucTion of pressure diagram and poinT of applicaTion of


resuITanT see Figure 1?, also secflon lh5.3 for cohesive soils.

4. For cohesionIess soil The vecTor c.l is omiTTed from The force
polygons.
TRIAL WEDGE METHOD I=oR EARTHQUAKE
I=oRoE ow WALL I FIGURE 39
I7 .
16
I5 I I
I4 '
13 —
SS
I2
metr 11 I
/'\ ‘I
09 — '
oa I
ENT/C(O
07 ‘
os
05 I
DSPLACEM
04 I

Si
O1
Y ‘I__ A

O "I I I I F I ’
O2 O4 0.6 0.8

RESISTANCE FACTOR I‘/C(O)

NOTE : N = Threshold Acceleration Factor

DISPLACEABLE WALL — WALL DISPLACIEMENTS I HGURE 40


FROM SLIDING BLOCK THEORY
Earthquake Pressure Increment
HORIZONTAL BACKFILL; WALL FRICTION = O
—1
1.1
1.2
I.3
1
1.4' _
-1.5
\/‘E
_ 250 \
1.6 _I__ *\

1.7 _ 50-= E
—-1.8
T7 + X T
C(O)
E/
-1.9
__2 ‘EIL t S; t.‘ + E

§\
2.1 I I I
A
KP
—2.2
40° I K“;\:
-2.3
7*
I Q
-2.4 _ 45° , A '
. T LL __ _,_
-2.6
, ' 11
2.8
2.9
_31‘ I 7 I I I

O 0.2 0.4

EARTHQUAKE ACCELERATION COEFF. c(o)

Totclf Pressure Coefficient


HORIZONTAL BACKFILL; WALL FRICTION = Q
s\

5.5 2 ‘Q’: 45°

5 __ __ \

4.5 , _L— C I L
‘F 10°
4 7 i7<;_‘?

I‘
3.5 _ E‘
L _ I “Y
‘J
AKPE

3 * _L Q

2.5 _ T5°°
» — _E_ §

T a5~ ‘E
$1
2 _1_ A *—‘~*

1.5
2°“ \\X \

I ‘I I | I I I
O 0.2 0.4

EARTHQUAKE ACCELERATION COEFF. C(O)

MONONOBE — OKABE PASSIVE EARTHQUAKE


EARTH PRESSURE FOR VERTICAL WALLS I FIGURE 41
ROTATED WALL
H 0 __
If p(z) =K|,.15.z
II P|=e TTTTTT T-
H 9 / Z
,1; IJIz)=O-8ES.9('l—H)
1
1
Z

TRANSLATED WALL
bl:-I
0
\
\\
___.T_T \\mi1><. p(z): Kp.2S. 2
H I \

|i Q \\
_-11 3 _ B.
|JI2)=’I-2 ES;
I H
Z

FORCED WALL EARTH PRESSURE DISTRIBUTIONS I FIGURE 42


N Egm
8
EEM
EmEm5
m_L263
_>E_®tEm8EUmE_w_

TQEOU
@QCUE
_n___\
_ _ _ _OH____3x
_ _V_ E
___tEVO
85%
JmWLm_E3
wWQL>O_tmm_QtmH LkkomQLDWWQLQ

W
M
o__ I.iYV‘‘

M
w5
EU;
B53
BB DE
4‘_ M
+P _
<V__p
w<V_<\mN
mu
___
C
&_
___H_T_
wnu
______
P<4 QLDWWQLQ

SPm
Ii.6“

mm
W_E

EI“K_'
%
M
__4_
>

AM mm AW Dm
SWT T N‘ O
Cl B U R E Aw OQ
Acfive failure wedge surface

Q5 From
= / £6 MO
Equafions

CI°IW E _ 'fie anchor

// WI Q
/
/ Assumed passive
/ I failure surface

77Z<>’7Z\Vfi“"/ Rm
P—_"/
PE
From
MO Equafions

FAILURE SURFACES FOR TIED BACK WALL -

UNDER EARTHQUAKE LOADING I FIGURE 44


'1’A><
&

APE PL ' APE


P; pl; PF P2»
P5 Ps

LOAD LIMITING CONNECTION RIGID CONNECTION TO BRIDGE


TO BRIDGE SUPERSTRUCTURE SUPERSTRUETURE

APE I APE

Ps I Ps

\9/74\\%‘\

ovum: PRESSURES our OF PHASE

APE; pi APE; P Z

’T='/ ‘LTTPSTI F
Psz
A /

IE ovum: PRESSURES IN PHASE

EARTHQUAKE AND GRAVITY FQRCES 0N I


I
BRIDGE ABUTMENTS HGURE 45 I.
I m3_%@_h__
I

V65Vgm|ll
I|_<D_E_m” _>

toto|_| _| <> > ”@Z_WM_gm|


O“EDIW<WMM_|mE<_¢_|Q_O<_U:P_EO_95EU

E‘_“pm
#__‘H_?_“A'_____\_I_A_1‘}__v _’_ |_ ‘_H \_V“‘$' \_"____
____“
U:_'
‘____‘
w_’__‘
_\
__\
__“__
____"U."‘F_h_____
“__"_\“ 2_|_HM‘
___'
___’“‘\4.:_(______l
C_‘;_“J______?_‘ _“___H
|__4‘_\_____ _‘__‘(_’_I \__”“__N
_“__l‘_; "___ \__U.)_“__”___l_
__ l_

9/9_/_n_9<_O__
I5]_N
_’
m_OE<IUKDW$_IOh_ ZD

$5
fly__
5 uh__’___JHI
U__¢_5 "___J
I_V‘ Vl_,_7''“_'u__“
___'“R
__~ ”_'_
II‘
‘___

_23
$96
m2_E@O”_h<_IU”Z_D3Q

D24
I_ ZM$_MOEO_"|_Z m_|E _<>>

K
:_|_ m/iw
’_ “___ __> '_ __

23mwmqlogm
H__%O_h__
I LHE
ZU23N_J<$_O)JVED“O_Z
WI
MZ____
O<O___
OM09‘
OP
MT;
J<_I2mM__jWT;
__w6__/m_| m<”<D_O;Q>Z_MOON;
>_DZ_

_(_gfigm
_ZOIOZ”’_O<Th_2m>J_DZ_U| Q
M
1m_ImPUE>E<w_;5Om<En_‘ JUM<>I_DM_$
FWOUM_<IK_"N\
n0’_ O___MDO_ME_Dw_mEmE<Mn_
m_w5<Z<
Hgg
W4
I8
QJ05
”MOON;
_mMDDJU23
<QZ<_ _U|®Z__D<QJ
_<_ZE__ _

_ _ ¢ xm2Mm<Qa figIEEOZ_,_a:|___3S
NMw2q_NZ:
+\_\ Q24
NZ:_Q<O__

MQQM;
ha
m_mfiO23m_Z_L_
Z>_aO_w|D_wmi| <_N:UZ<_fi _

.'\\
/ ’ \V
I“
QAW
$\

Dm_2q_ZJG
_|_
\VK5 _\V

xh
Q
Q‘
u“ Q

x
\\\\\\\

h_O_n_
<

i
O‘ ~|_ L‘ I 1 I I I T
/ I \I, , ‘I
/74 ’1, I mesa-4
0'2? ‘m= 3/ ‘I\“ \I~\\ 7
V/‘ / R \
I-_u I I I

z,H \\ \\ \ I I AI A
W I\ \I<’m‘=O'gI I /

n: <9+- 3 =10 7I"'\ ‘ -3 ‘I7;/


Im5(D'4' m=-I Q A
F
O
//ZR‘//14*?‘ O -5
Q)E
‘r \\\ \ \§
Q ow
75” ' m R
LUE
VA
9 9°
/ IM*\ 6()H
56I4 I
\
[I//
I
m FBQQ R
-vs-sew
-so-54H
11 \<§Qk
999? \]U§U\-#
52H
48PI \>\
\ ooc_> ma» -46-48H
K
O O2 04- 0&5
II
O8
III
Q -5 LO L5
VALUE or= pQ Ii) VA LUE OF r>Q<iI5>
Q1. - QP
I=mH Q|_ F Ix=rnH IQP
F ' ' _ or m g 0.4
_]: n&oms$7

H O.20n E
pQ(6I:) II
Z-=nH (0.16 + n2)2 H4 Po
H JO H -Po
R PQ = Q.55QL _I
A‘-RA
' _ ‘e For m > 0.4
PRESSURES FROM For m g 0.4
A _I_-I_ _ I.28m2I'1

(E _ O.28n2
_ O.64QL “Q Op) ' (0.16 + n2)3
'% "<m2+ n ‘U O
€ ,, Q0
Q1. WALL For m > 0.4
777<w77&T“ \pI)
40
I H2 I.77m2n2
I PQ(6;) = (m2 + n2)3
_X =fl'1|..|
RESULTANT
Pa =KA Q1. SECTION A-A p'Q= pQ cosz (I.IO€)

RESULTANT FORCE PRESSURES FROM POHVT


FROM LJNE LOAD QL LOAD Qp

(approx. method for low‘ retaining walls)


LINELOAD ' I ITETIPOINT LOAD
LATERAL PRESSURE DISTRIBUTION ON WALL DUE
TO POINT AND LINE LOADS
IHGURE 48
_
wQ33>gwm A
O___ A w kw(6 6
oom
oomv
H
*°om‘"9
1‘ E;v_o‘M
mmgzx
egQMDM?
__
O
O__
N__
®__®__¢__
ON
_<QX;&+<&VO_P<M
jEh_OmM§5_/q_m1>5_ZEh_OmIZWQEOK
52; w”_'<5_OZm__D>>|Owm_l'h_n<_M_M>M'>M_>m$_<n$__
D265
MOU_k<FmE
.mm______<_$ '

Aw
mE\_/_VEm_g_W
":_£\
N__\ NA
AV
mmMon3““_OM_
ZO:_U_m|_h<_ZIM_;ZO_Z<
WUIjgOL _2m_5Z_
LOm
6
$324
wDOE<>I8
H
m__‘W\
Vd

VIN

©‘O NO 4:
V_
/M
U‘ H M

___‘
II‘
3:
M1§
_<n_
<n_
C20;
OS
ll‘n_
N3;
U
%.
\}k\_mvQ$
EMA:
'_kn};
MlN“
|
_ _HE|_;_<h_>U_<v_"
%_mgODO__v>$n_ >_fin_ >_

E
OwOQO
8QOm
wI”QwZ__DO|Wm__<__InQ_
RI
_ *”
N
'WNm_>_‘||||¢
<V_u<n_(
*NHf
6
_ _ ___1"_l _.) _l H‘-
Dh_OZ
_ _ _ _ $ O_ _ <E<>
WJO<3
Z<
__O
1 I\:“Hm NJ02<<8

IW
8 Q0{
W
‘ _'___u___H_H__H___v_v_‘
L _ H2 m__H5M___g

Q‘
Xv
M
Q0_
é

E
moQMEqUQlmGZOwwQ
ELMZOzKDgWmaWgEQ
wZP4gwJ50
O_><$On'_Z<&>_

924 m_W24:
2M3“_O>3<8 GZEEj_E

Z_<IQ
O-5»-I LAYER OF FILTER MATERIAL
O2. CIEOTEXTILE DRAINAGE MEMEIZANE.
"1 ~ s ///®’//®’//
,4
I
O-0 M
. ._\ 1

75-.... MIN. DIA. wI-:EI= ~S


.v

"- ‘ I
I‘ ‘IIIIIIII-‘H
~ I‘
HOLES AT 5... 5 "‘::Il‘l‘!|‘
CENTRES.
- I‘
_) L‘, '
1'.
II
I
I
II
II.I'H|I
1
wI-IERE ‘A’ IS LESS THAN
I *..
A
I
IIIII ‘\I‘)
'A' 0-em, FILTER MATERIAL
I . -.110
»--.-
‘H- .1-
I,‘ '1
53% ‘III
‘".'I»' I‘»’~I CAN BE OMITTED
l
I
3:
I I .1 :;:~.~. -,-.-.,I I I‘!!!-
¢
4_ r
,~__ I\.I'..'
r '1.’ >~ x >- “AA:
I 1:‘,- T .‘-H.
":¥x>'>I*x"'**‘**\<*xxxSx‘xx
\‘ \" wt.»
. X 2 xxxx\<"‘X>\*xxx:\;\
I'.\ ' -v pf xxxxxx Xxlkxx.
'.I;' ‘I x xxx Xxx '
B I .1 J “XXX )(x)\xxxXXgx,\xK
x x)\¥x;\x,.""><s-A
_ .
, a
'I r* xx;.;_,\xx%xx
AX>lX\;\x**3XK
\
-. .1.
u If
X
x X
xxx): xxxxx kg‘
A * IMPERVIOUS BACKFILL
$1 . KKKKJ‘ BELOW WEEPHOLE
HESSIAN BAGS (EXISTING sou.)
FILLED WITH ‘- _’..__,,-_.".'. __I¢,.__
O-O?»-R5 OF '~_'~
-‘
‘19
>3»:
X,2.-';f£_Y
-~'aw)!
‘:->Iy)‘,
X ':._:|_
‘, 1.xy
‘v-y.,‘1,.-':_-'31-»
>.._-.,- .-;="._'._"Z._'-__I.-
X
K.x)‘
:0-".
--‘,1
.,__..
COARSE AGGREGATE -. _\:\:-_.,'
1:“-‘_.'
'_:“_'.
“'I‘.‘:.'_..,_:» ;-_
=._- -1'I.‘.1'.Ir. x)‘,.>~x\-(xxx:‘Err
,.',_'_-
nxxxx*x"x
._rI Y
"X
Y
xp;;
Xa

DRAINAGE SYSTEM FOR LOW WALLS

O-Sm NOMINAL LAYER OF


FILTER MATERIAL. .OIZ GEOTEXTILE
DRAINAQE MEMBRANE
BACKFILL
MATERIAL ‘\\v,~.

ORIGINAL
/'."I\"\\._‘ GROUND

////'
I‘ ‘I ‘I

/“I\I\I)l
‘IIIII\l1
II In III\§

I
| 61",‘;-' '
!I\|,‘II‘
CONSTRUCTION

O/4 " DRAINAGE MATERIAL


ad‘ H
“»>?.~is»‘To_"‘
.I“
__ __ Iso...... MIN. DIA. sue>soII. I
# i # § j -—- ii‘ PE.

DRAINAGE OUTLET THROUGH ' T‘


GAP IN BASE SLAB CONCRETE BASE SLAB
CRIBWALL DRAINAGE [_i_**F|GURE
50
W
mmz_Q'<é\w_2‘m_ §i_wnxm_|zOm_U
®2OPm_Z_ Q_2<$EE_‘O¢m;s_ |_ UE mE8_“
_m_ _\ _
mm;EWP_<_l _|2:hO_*Z“W_O
_O<
NOwz_Uo_| ®<wO_”_mZ_<E
Etg
ZO_ _OD”_ 'wZO :30OF
Uta
mam
zQ______
s6mI2_Qmwx|0__DOE_50“_2E8M30: QZDOIO_3/_ 0_m_o
z22_/_<:_m n|$_mOm 5mm
m3_ _“o_m>V
mA
8w¢E_\
{Q0_
____Vk
_m_0< /__:<m|_<o_” H_":<_$ _ _ \.mn|% \_ _\nl>_1]8OJ%_
_ %5$:F_
\_ _
o2_gm_m
M55200
50_xo_~_n_“_<

jg_WWA~A7'”_ _:>
\ MGM“
‘I_]_<_”:_m_l _<§
____:__
___:__
’_ __J
__‘
__ ________
_
__
__
____
_
______“_
____
J
:____
______
_ ___<mw_"__“‘lg\hz_m\O_oM_____
_ _“___"___
_€_____
___’_
I
_ m_MOI‘
w|Qm_2|_‘Z_<>>H<_¥%U<m ®|_ S_wn_
_fl
_______
__
H

“15>H6£_80wmmzEo“_”w_ ,_3ogOE_U5m
________ qw Q2mw‘_qM_______
248uqm_‘m am ‘_mm_>mu:_’
__‘_ ____
____
____
_
________
___ _ ____
:_ _ “_ _ _ flh________
_%“"p_________
_%_J_______
,“fl_4__'P‘_nw”_m p_ _u_ "
___‘_____ _ “ ______
_

I _I_'
___
My
__‘
’_°_l,
_

\\\
IL,_ may
__Q___
H_°$_
_'W___'__pc_°_"“n'_%U_,W_HA1__H£___m__
’w:::_m‘_H___5"A_‘:I__~LW°__'_Im.I"_i_°___‘_'__’.___‘I_H‘_i,__W_°__~'’m_yI_"m___“_‘I;OI_'_:-__
V6 :_ ~_____m_
_' ‘
\

‘O
\_‘’__
_'__
7
_|___
gv
)_____‘_I‘
_r"
Q‘Z_
~___“__
_4_
\I“;
_‘U
__
___U0Ei._NUo7<m_ _
9}”P"
___‘__
_

I“
_
_____

""_‘'
WOI
Mm_O<_“EP4;_OU_‘ |_®O _I__
_ '_____
‘_
‘ll '
I O
__ wl_oO5In_M_>m‘_3m_\| ta|_/; EOm_; ‘_om_< ___
___
UL_
_
‘___‘_____
__U_“ H___
_'
_:_I
AZEIQ___|_<>>V
___‘l
AWO<2_ /#5I ‘:“__
_ I<JDZ<Z@‘_'_ “_v_Ug _l
_“__H
"_ -
‘_

OZMP_WDIZO_ZJOmO
_
-it-.
I00 IIII Illl . " llll II ' Illl ll * IIII
IIIIIIIIII-“IIIIIIIIII llll ll IIIII
IIIII IIII LI llll Illl llll ll IIII
:::::|.'.::|:III:|::::|:::..
9° IIIII lll l*\\&\\ llll
l llll llll ll:::
.... .... .
llll nll \\\ Illl l'Illl
nu-Ill--II~I\“u|nl.n-u‘l.unllulllm
|llllI|IllI—\\\\\1llll lll II::
80 ulll-lll-—I\\I\\ulll l II
ulll-lll-—u\\\.\Illl
|llllIlllI—l\\\\II1_ll
III
lll
IIII
IIII :5 _ —____-iiliii:
lllllnlll-_l\\I.u.ll I|ll IIII III
lllll-lll-—lI\\_\'ml lll IIII l
70 ullll-lll-—-l1III1I\Il III nll
Illll-lrll-—I.l'nli\I II llll Ii
lllll-lll-—-nI1Iv_ImII lll Illl III
IIII-—IIII— —-unulu ll IIII III
GHT
WE IIIII I'§I—ITA\‘lI\\k\\\ I“ IIII _
>-
60 IIII—:lIlI._ZIlUAI‘\IIl‘\‘LIFQI
IIII__Il'I_I-_I:II\‘\‘L\"II:_
IIII—_II—Z-II 'I\\_\‘\\_\_\Y
IIII
II
4.I,_
ii
CD ..-----..l-2-‘.\‘.-\“‘l"- IIII IIIII —-I—r-<—I-~—
IIII-—II—I-—II\IIN“\‘“N- IIII III . !._-
IC=l=.\'lI __ _
50 ‘...-2-H‘--—..‘“““_.\‘Q-2-...----.!--—i I I I I I I I I I I I I I I I I I I I I
IIII_—II'_ZIII
I . IIII I IIIINK III. III I III
IIII_—II_$IjIIlII\‘\IL‘-IL IIII III -III-
IIII_—Il-_!-IIIIIITUIINX-—IIII-_IlI_I—-IIII
IIIII-IIII-—IIIII_—IUIIKI IIII III IIII
4-O IIII_—II_IIfIIIIl\_k‘N-SKYK‘ III. II _ II.“
IIII__II-ZI_IIII_l1IIL‘lI“ IIII III V IIII
I----—---—--....-A-—‘-N-L L L IIII II I,_ IIIII
IIII-_I-III:—IIII“—IflIfll\‘XQ IIII III IIII
IIII_—III$f-IIIIIN—NIN$I1IIII——II_I—-gg::
30 ulll-llIl-—1llll
lllll ll
‘IIIl'I‘?\‘I_IIII.Ill..:
IIIII I Il‘Il In MIII ll IlIII
PERCENT
F
NER lllll ll IIII IIIIII "lll ll IIII
ull. ll Illll lI_ll\ ‘lll II. IIII
llll ll IIII lll. lfll IRIS. llIl llll
ulll
20 lllll ll Illll I\ll'1.‘llnll1-llll-—1|lll
II— IIII ‘I K‘ L ‘ Illllmllu-—i|lll
IIII IIII III '9‘.-mB"@‘Yl;1“‘-IIY;_I\‘I|TIl‘|‘--—III-
IEII III IIII L_.4_-INEI \"l§£.II ‘*’\.."!."-W III-
ll IIIl II-_‘ . . ‘II
‘IS ._ ._ IIIII
IO ll llll llIIlI I lull lain llll
::lI Illl llu_-‘—lwllmlm-llsgn.—1::r|l
llll llll.—_1|llal:llllImu|1
l ll IIII lllIl—$z|lll.lml_llIs_§—
-
lll
I ll
IIII
llll
IIIl\1iiIIIL.—h.‘I—§‘
lll ' Ii: balls-

co g 2 2 2
W

E
00- II80 iiso <3 8 8 Q (D co 4- I0 <\-I
Q. 0'
GRAIN SIZE MILLIMETERS E55:
0'08 IIIIQ.95
COEFFICIENT OF PERMEABILITY
-4 FOR CLEAN COARSE-GRAINED
'0 Y
‘ EFFECT OF FINES ON PERMEABILITY
’ DRAINAGE MATERIAL.
I I t I CURVE k, m/sec
Io-5. A 3-mo"
z-9 X10‘!
. TYPE OF FINES MIXED WITH
lo_6I I COARSE GRAINED MATERIAL -. I 2; 8 »<IO'Z'
TY,
k,
m/sec 6110"‘
» sILICA FINES
I I /‘LIMESTONE, FINES h . I'xIQ'4
‘IxIO"z
lo-‘I I . \ _
3-2>I.IO'3
T X 3'5x‘IO'5
PERMEAB
L I I COARSE SILT ‘ I-I4:1O____
IO‘-”
IO-6 I’ —*-4
:6(O03\l0‘IUI.pgI|\>- ‘I >~1O'4
1 sILT
I I

lo-9 ‘ . I _.

I ‘ .
PERMEABILITY 0F
‘ CLAY I DRAINAGE MATERIALS
COEFF
CEOF
NT lo-IQ I; I r__I I T O I

O 5 IO I5 ’ 20 25
PERCENT BY WEIGHT PASSING _
N°- 20° s'EVE (OfferU5.O6pqrTmenToI'-II-\EI\Iavg,I5&Z)IFIGURE 52
TYPE
OF LOAD DIAGRAM STABILITY CRITERIA
WALL

I
'SLID ING

s = <w+ + PQ) Ian sg + age


. . S + P
GRAVITY Fs (sliding) = P
Pl-I
-log
‘_

PP
4——I' nu HEEL Z 1.5 (s+aIic loading)
E;q“§:)‘ or g 1.2 learlhquake loading)
TOE [inn BASE
l I4
d
‘ OVERTURNING

Momenls aboul The Toe of The base


I

F5 (overlurning) = !i_?Q:lLE
I’ PHb
2 2.0 (slalic loading)
SEMI — “ l OF > l.5 (earlhquake loading)
GRAVITY Also check overlurning al selecled
hori zonlal planes up ihe wall for
5 ‘igii b
grav ily Iype walls.

E
‘, REINFORClNG~ BEAR ING PRESSURE

Poinl where RW inlersecls base,


from loe,
,; I
VERTICAL a ' d :
w+ a + P;+ - PHb
STEM -0 '
W-|-+P(/

WI BSSLI ming Pp = O
CANTI ~ For soil foundalion maierial, d
LEVER -I should be wilhin middle Third of
pm ll
‘[20
1/
_ 1 HEEL lhe base (sialic loading) or
PP A5 SLAB \ middle half (earlhquake loading).
l‘ nun 3 For a rock foundaiion, d should be
“Bu BASE o|=
FOOTING wilhin middle half for bolh sialic
SOIL PRESSURE. and earlhquake loading.
F5 (bearing) g 3.0 Islalic loading)
l J I or g 2.0 (earlhquake loading).
-

See seclion 7.4 for calculalion of

\=Hi
facTor of safely for bearing.

\i'*‘
7

W1-= Iolal weigh+ of The wall in-


COUNTER cluding soil on Toe plus soil
- FORT. I above heel (for caniilever and
counlerforl wal ls only)
PP A l Pv verlical compone ni of PA
PH = horizonlal compo nenl of PA
IE§::III" ‘ Rw = resullani of WT and PA
FU = .EFFE.¢Ti\/E. VERT ICAL csnnnouemrr
OF PA = Pv - U5

STABILITY CRITERIA FOR RETAINING wAI_|_s IFIGURE 55 I


E Q- Concenlraled force
1h represeniing passive force
I V acling on lhe back of
1 |-I The wall for a shorl
dislance above loe.
P
A d Oblain penelralion from
Pp l
, _ la
P .1 =P .1 I
Passive Aclive Q P b A Q
Pressure Pressure

2 Faclor of Safely
Z P
., d=
E A‘ F5 = ii»;
P I
"' %\ PA1‘ la1+ FATIOZ
lb.-1]
P» E 2 E
PA2
Q ICI2 l
‘Q1 = 1/2(i<,,-i<A).x. <12. 1,,“
Ne’r resisling Nel aciivaling I/zKA,X.h2.lu1+KA.25.h.d. [C12
pressure pressure
where PPN is lhe nexl passive
pressure on The wall IIEI passive
FREE CANTILEVER WALL pressure-.ac’rive pressure due Io
soil below excavalion level.

A [o
E
-
-
Faclor of Safely
£1 h
== FPN-lpn g
Ibn E pA1'lu1+ pA2' [[12
hi \ I = 1/2 (.l<p—KA).2$. dz. 1;,"
10v2
=
-7-
___- PM d 1/2K[1.3.h2.IQ‘|+ KQ.2$.h.d. luz A
"— I
Nel" resisling Ne’r aclivaling
DFESSUFE DPESSUPE

ANCHORED /PROPPED WAL L

after Burland et al (1981)


SHEET WALL STABILITY- COHESIONLESS SOILS FIGU RE 54
250
20()
160 l

I20
IOC
BO
as <2
1-

-I5O
-

3O
:-

2O
I6
12
N}
N5
A9
oclyfocflns IO
8
cap
ng 6 \

Bear 4
3

I I [Gs I I I I I
O 5 " 1O 15 20 25 3O 35 4
Angw ofsheafingreswlance, ¢ (degrees)

I BEARING CAPACITY FACTORS FIGURE 55 I


Ll

I
Ground level " -,, _
V §

B/2 Ground surface


H D (not greater
than B I Q
I B
<———-————--———->
I , .
Equivalent horizontal base
_____ ______ _

Loading arrangements for Brinch Hansen’s general equation

Eccentricity with .
ex / respect to centroid er Inclmed Ioud
Eccentricity applied V +”*I OI IOIHICIOIIOH . I
vertical load /T ' I Q~

____5'=a-26,, I I I
-<————-——-> <
I I A 8|I I
IIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIlllllllllm
Pressure distribution
I I I I I on equivalent base
l V
Point of application = _|"_'T
of resultant vertical <_?!5_+ 2e, B XL
force l/
e,I'y L__
_L=L- 4
I
.

Centroid of /14:1-__
foundomn ,9‘ K‘ \~ Equivalent base of area A‘
<—-—i——I—> I I
i :8-28/Y =BXL

I - I IL

Transformation of eccentrically-loaded foundation to equivalent


rectangular area carrying uniformly distributed vertical pressure.

LOADING ARRANGEMENTS FOR BEARING A


CAPACITY FORMULA E I FIGURE 56
Shape factor for vertical loading
2O

¢=45°
I.8

1.6

1.4 E =4o°
fShape
asc
ctor, =3o°
=35° S sc-1
1.2 ._ I C-——-*---i

Nq
,=0-25°
1.0 =3o°
~s +“ E0-255'
O8W fl

40°

=45°
fShape
act.9),
or, 0.6
I It ‘Iii I

O 50 Q2 0.4 Q6 0.8 I.O


Breadth / length, B/L

Approximate values for the shape factor for centrally-applied vertical loading

Shape of base sc sq s,

Continuous strip 1.0 1.0 1.0


Rectangle 1+0.2B/L 1+0.2B/L 1—O.4B/L
Square 1.3 1.2 0.8
Circle (B = diameter) 1.3 1.2 (1.6

Shape factors for inclined loading, then become


s,,B = 1 +0.2icBB’/L’, '
sc,_ = 1 +O.2ic,_L'/B’,
sqe = 1+$In <I’IqeB'/U, NOTE : See Figure 56
for Loading
3111- = 1 'I'SIn <I’IqLL'/B,’ Arrangments
SYB = 1 _O.4i.YBBI!L,,

S,,_ = 1 —O.4i,LL'/B’.
The shape factors s,B and s,,, must not be less than 0.6.

BRINCH HANSEN BEARING CAPACITY FORMULA


— SHAPE FACTORS I FIGURE 57
4.0
5.67

3.5
¢=45

3.0
3.59
n =4o°
4-.
I\) U)

Q-I
hfac
dc
p
or
35° I2.5e
for
D/ues5’=oo
De
2.0 _. o Va
39 2»-2.04
“ =2s°
1.5 .. TE =0-10° oiovkioooig

I. 4
OO 5 IO I5 2O
Depth /breadth, 0/5

dq—dC—T.
dc 1

d,=1

BRINCH HANSEN BEARING CAPACITY FORMULA I FIGURE 58


— DEPTH FACTORS
l.O ¢
IQ"

0.8 _ _ =lO°
=20‘)

/c
:-

or _
.0 = 30°
actor
q_.
Tl
.- =4o°
4>
.9
HCFIOO

0.2"

Q I I I I
Q 0.2 0.4 O5 0.8 1.0
/fiPLc+Vmn¢
I.O

0.8

T,/I4
cn
.0
¢_
F7? =1O
§b-

CIOf!CO
'¢-
Cl
0.4 -.-
:- =20“
0C"
0.2
=30"
O I I I ‘4°i
O 0.2 0.4 0.6 0.8 IO
/-//V + B'L'c cot ¢

1.0

0.8
._>~
.- an
.9

fat:
or
n
@-
O.4

|"|C
O
FICI
¢ -1o
O 2
-200

O I I =30“ I I=4o°
o o.z 0.4 (me o.e 1.0
H/V + BLC cot gb
I after Tomlinson, 1987)
BRINCH HANSEN BEARING CAPACITYI FORMULA HGURE 59
- LOAD INCLINATION FACTORS
TOE MOMENT EFFECT ON HEEL

The toe support moment produces a


loading on the heel. It it is
assumed that no moment is trans-
mitted into the stem, an equivaient
parabolic heel loading is as shown
below, with the maximum ordinate

MI I L I . given by _

p-I- = 2.4 IVI'|-/a2

L"—1 dl
where MT is the toe support moment.

WEIGHT OF BACKFILL
ABOVE HEEL
YI 2

Wxd Illllllll SELF WEIGHT OF HEEL

WI I III LOADING FROM


TOE MOMENT

"flW;‘>
ASSUMED FOUNDATION
BEARING PRESSURES I
l>
fl1OX
.?
r.__.. _._ .4

RESULTANT _LOADING ON
+VE HEEL (May be fully positivel

-VE

Note: Pressure diagrams not to scale.

DESIGN LOADING ON HEEL SLAB

IFIGURE 60
/ ANCHORAGE LENGTH

ASSUMED 45° /\\t4“I°


CRACK LINE
FROM e. / EQUIVALENT Ia FOR MOM.ENT
BI ° AT e.
ASSUMED 45., gI // EQUIVALENT jd FOR
CRACK LINE I=RoM/FD MOMENT AT II-
A. , 45°
A MAIN TENSILE STEEL
IN COUNTERFORT.

A = Point of maximum moment (maximum allowable stress in all main tensile


reinforcement). .

B = Section ot lesser moment than at A. It some of the reinforcing bars of


the main tensile steel were eliminated then there would be maximum
allowable stress in the remaining bars.

The Icut oft‘ position tor some ot the bars of the main tensile reintorcement
is to be the greater of:

(6) Anchorage length past the assumed 450 crackline from A.

(blflmmm past the assumed 450 cracked line from B.

'jdI can be taken as the perpendicular distance from the centroid at the
steel to the midpoint of the stem slab.

'CUT OFF' POSITIONS OF MAIN TENSILE


STEEL IN COUNTERFORT
IFIGURE at I
TRIPLE WALL

DOUBLE WALL I°°


SINGLE WALL
E

I0

Ha
H21/mg”
H1(max) /

32
(min)
L1_2{/ 1
“Q.
é%
I/‘I23
/mn}
3.15 "I H33(m,.n) H

ASSUMPTIONS : Soil properties :¢',c=0. ‘a'=19.5kN/m3


-__-_.- ¢ I = 40 0
woii properties 1 s =%-I0, Ww=15.5kN/ma
-————¢=ao Wall slope ;p=_11.'IIiinII
Water-table below base of wall.
Liveuload surcharge equal to 0.6m of
soil included.
Fs (sliding) = 1.5 min .
a:":* _L Fs IoverturningI=2.0 min.
3

HTS . _

Hit"
10 Ii R"r'~ _L §(_\ ‘ H32

-—~
t L L
E
3
I\L_ ~I\
_\ 33

WALL
LL
O I TI‘ ~t\
\\
3
/
GHT
HE
SI:J;1tJ*rI>t EN
I7 JCT?“ —I~ ___ \ $7 T
/
e/is/‘E/
wa \ /
/Q
I
~ __

/7 1‘
‘I
L ~
I
"' \ \\

/I“
I yI/
SLOPE \___J\_J$

o . 5 10 15 20 25
BACKFILL SLOPE w Id=9r==$l I

CRIBWALL DESIGN CURVES


' TRIPLE WALL
ooueuz WALL “O
SINGLE WALL

8
‘ £5
H1
(max
-P
H22
H32
1”,,,,
E
Io~——i——
. m
Q

'
Assumptions: Soil properties 9; l"| = =19-5kN/m3
Wall properties = 8: 19-0 €I>e =15-5kN/m3
9 Wall slope 2 p=-1li° I4 in ll .
Water table below base of wall.
Live load surcharge equal to
8 O-6m of soil included.

\ F lsliding)=1-5min.
7 ‘\\%’ F loverturningl=2-Omin.
\\ \\ Seismic Coefficient
H
Cl0I= 0'2
°\\ \\\ --— ¢'=1.o~
‘SK ---- —- ¢'=3o<=
\ \ ' ,_
'~
E 5 \ \ H3|

\ \
Q [:5-gs A/§\_\ \_\ 52
\\ \\\\\ \\ Z1
\II_Z1\ \\ \ Z \
\\ \ \ fi\
— 3
' \< as
OF
WALL
HE
GHT i
E
i
\
Q\ \\ \
SLOPE 2
\\\\\\ \\
/
/
/7
//
, I
//,as
///
\ I

5 _ 10 15 20 25
BACKFILL SLOPE 0.)“ (degrees)

CRIBWALL DESIGN CURVES-EARTHQUAKE LOADSI FIGURE 63


WMZO
EDWWME_Q|_<|_U@“_
¢®I
<_>n_>m,_ g

UQQ
m®J+Lo%
UCW_m
_LIQ_OLMQ>@
+l*0
&
CUW_
95+
@<&
+CM
CQ_3m®L £m3OL£+ W+O@
$0
(a
w£+
CO_+O®WL®+C_ I_
D
M
CmmE *0 $0
05+
UCQHi
W@C__
COw+U®_N&

<N\
Nm w
W®L0+
>@E
ODD_U®mU
mm¥ N£+On UCG
®UZ+__CO_+O®L_U

I/F4FmA, \
X
QOPZ
_n_
n_Z_<PmO OD>JOn_N’wUZOh_ UCWfil
UQLU
®F>n
U_3N
U_©U i
wmC_mCmw3
nEO_+3UoO3U®

Q
E1]ENwo_
AWKDWO /
_ (n_
B ‘H% fi% fi %Q_N€
__3+
ZDO¢
e”\
"
Q QUE
®_$2
_U_ OC°U
_EEom__
u_______E
8_ 8_ E
Q
V
my}
ki I,|‘\\\\&
'

Hv‘\“‘l | IMI I V
,Seam
o_€__m
%HHHHY%fi %fi$I,

_ '_I_h_h‘_ .‘ \Htx ‘HI2W“_HI“§i‘Ax“_ H

m
W *gs 2E90
BUEW
263
Emo nhe
_ o_uC__EEUm__h_E__2_g>
\\\

\\\\
~l_
>II
I o b
//svzsw / I. '1.
I
—-.'- /
$

H ——-P ¥ I
P
I
Iw>I_
I
U1
"4"-—*'PH _
\ Q
-I
‘JOm

FIIII
A
I"°
/ FAILURE d c A
I/ SURFACE
/" I-———-I
DEFLECTED POSITION PH =o,e5 KAYH '

EXCAVATION EXCAVATION
IN SAND IN CLAY

The above apparen+ pressure diagrams may be used for de+ermining The s+ruT
loads in braced excava+ions.

EXCAVATION IN SAND

Area abcd Is The pressure disfribufion. The resuI+anT, PH = 0.65 KA Y H2


ac+s'a+ 0.50 H above +he base.
See Figures 5 To 8 for KA.

EXCAVATION IN CLAY

Area abcd is +he pressure dis+ribuIIon. The shape of This diagram and
The magni+ude of fhe pressures depend on fhe value of +he s+abiIi+y
number (N3)

2<N5$5 s<~,,_<|o |o<~$<2o 20<Ns


\ PH .75»-:::“ -avnbfl (I-25--O38NsIHPH' -5Hp,,
I PH O.4yH "7:-I-4c 'YH—(8—-4Ns_)C TH
A '25H “ O O 0

5 -50H -75H (I-5--O75N5 II-I O


I ‘ I

R '50H -44-H 338 H ~33H

BRACED EXCAVATION Pnaessuzz DISTRIBUTIONS I FIGURE 65 I


q q N
F (base) =——-Pi-
F II ‘I S

1 [ c = AVERAGE UNDRAINED SHEAR


STRENGTH OF THE SOIL FROM BASE
[ H. LEVEL To A DEPTH OF o.2sH BELOW
z _ THE BASE
B—-_;—-%.:—..:_i] Nb: STABILITY FACTOR

L= EXCAVATION LENGTH
U /“Q? _

‘° I I I TI I I I I I I
Q i

/
.0 SQUARE OR
Z
3— CIRCLE %=1 "
0.5
FAC-TOR 7__ >__
LTY
I 0 , INFINITE STRIP

STAB

STABILITY FACTOR FOR VARIOUS


S [GEOMETRIES OF CUT I

0 I 2 3 L 5
H

B (After Janbu at CII. I956)

FACTOR or SAFETY wnn RESPECT Tu BASE HEAVE


QLSIIJIICQ from l:_><cav.
Max. Depth of Exctat-'.
0 0.1. 0.8 1.2 1.6 2.0
b\
Q
I I I I 4,> i .

Distance from Excav .


Max. Depth of Excav.
P
L. O 1.0 2.0 T 3.0 4.0

,..
(
(T
0.2 e” . VlII"""'
....
Id

t_,-)
De)EMF
ofth °-3 ZONE FOR MEDIUM TO 1.0
DENSE SAND WITH Excav
of III
‘T/IELIX
~ 0.1» INTERBEDDED STIFF ement ~_I
f——————CLAY, AVERAGE TO ¢_:
‘.4
2-0 Ial SETTLEMENT DATA
GOOD WORKMANSHIP Se Dep
T
' [After Peck,1969]
Distance from Edge Of Excav. 5
2 .
Depth Of Exeav. 3.0
P’)
O O-5 I-Q 1.5 2.0
I
»-I
I I I
\r

X Zone I
1-O -
>
‘-0 Sand and soft to hard clay average
,-
2'0_ I bl workmanship
L
O
... Zone II
... 3-0“-
SOI"
I I at \/rry soft to soft clay
'3 I I Limited depth of clay below
;. 4.0-
h<>ttoIr~. of excavation,
ta
2) Significant depth of clay
Arg 5-O below bottom of excavation
but I\Ib< 5.I1I.
I bl SETTLEMENT DATA L»)
A .
butt Ionic-nts affected by
c.onstru<:tion difficulties _
t offer 0’ Rourke et GI, 19751
N°"@ Zone I I I
May be used for approximate
Very soft to soft clay to a
guidance only for residual _$0iIs. significant depth below bottom of
excavation and with Nb) 5. I4 ,
K I-I
where NI) I-= T
I -
Simplified
settlement
r\J=*F*»,/ \ Struts
‘ I
profile \ \ I
|
I
I (CI GENERAL TRENDS OF GROUND MOVEMENTS
lateral _id.__.-'--"I
movemems I \ I Exaggerated scale]

base heave
\ flex
I i b lé side su'pp orts
..l I-

LARE EXIIAVATIUNS — SETTEEMENT GUIDE


OPUS
INTERNATIONAL
CONSULTANTS

You might also like