Sie sind auf Seite 1von 59

HANDBOOK FOR REVIEW OF DESIGN AND

CONSTRUCTION OF TEMPORARY WORKS

FOR DEEP EXCAVATIONS


Forward

LTA carries out many of the deep excavations in urban areas where ground
movements have to be carefully controlled and monitored to avoid unacceptable
damage to adjacent structures. Temporary works are always the responsibility of the
contractors, however in most projects LTA retains the responsibility as the Supervising
Qualified Person (QP). It is therefore necessary for LTA staff to review the
submissions of temporary works design from the contractors. These reviews are to
ensure that the proposed temporary works satisfy the codes of practices and the LTA
contract requirements, and that the design and construction follow the best engineering
practice in the industry.

Information is provided in this handbook for the review of the temporary works design
submissions.

The handbook is prepared by a working committee comprised:

Wen Dazhi (Leader) Assistant Manager, Civil Design Department


Cai Min Design Engineer, Civil Design Department
Chan Chee Weng Senior Engineer, Development & Building Control Department
Chiam Sing Lih Design Manager (CCL3), Design Management Department
Joshua Ong Design Manager (CCL1 & 2), Design Management Department

In addition, the technical group led by Mr. Yong Ah Poh assisted the preparation of
the graphs and figures in the handbook.

The Working Committee is particularly grateful to Mr. Song Siak Keong, Manager,
Civil Design Department and Mr. Paul Fok, Manager, Design Management
Department for their guidance and comments in the preparation of the handbook.

2
HANDBOOK FOR REVIEW OF DESIGN AND
CONSTRUCTION OF TEMPORARY WORKS
FOR DEEP EXCAVATIONS

Contents

1 INTRODUCTION

1.1 Scope of the handbook


1.2 Temporary works
1.3 Cause of failure

2 DESIGN CONCEPT

2.1 Construction sequences


2.2 Design assumptions

3 DESIGN SOIL PARAMETERS

3.1 General
3.2 Geotechnical design parameters for Phase I/II MRT projects
3.3 Kallang Formation
3.4 Old Alluvium
3.5 Jurong Formation
3.6 Bukit Timah Granite Formation

4 STABILITY CHECKING

4.1 Base heave


4.2 Hydraulic uplift
4.3 Toe-in depth
4.4 Slope stability for open excavation

5 DESIGN ANALYSIS

5.1 Overview of analytical methods


5.2 One Dimensional Finite Element Method
5.3 Two Dimensional Finite element method
5.4 Drained Analysis vs Undrained Analysis
5.5 Total Stress Approach vs Effective Stress Approach

3
5.6 Note on Using Effective Stress Approach for Undrained Analysis
5.7 Sensitivity Analysis
5.8 Back Analysis
5.9 Use of Sacrificial JGP
5.10 Treatment of berm
5.11 Strut design

6 GROUND MOVEMENT

6.1 Prediction of ground movements


6.2 Factors affecting ground movements
6.3 Consolidation settlements

7 DAMAGE ASSESSMENT

7.1 Classification of damage


7.2 Methodology for building damage assessment
7.3 Steps for building damage assessment

8 INSTRUMENTATION AND MONITORING

8.1 General
8.2 Review Levels

9 PERFORMANCE OF TEMPORARY WORKS

REFERENCES

4
1 INTRODUCTION

1.1 Purpose of the Handbook

The purpose of this handbook is to serve as a guide for review of temporary works
design, prediction of ground movement and building damage assessment for deep
excavations. It also gives guidelines for construction supervision of the temporary
works to ensure that design assumptions are realised and design requirements are
complied with.

1.2 Temporary Works

For the purposes of this handbook, temporary works refer to retaining structures and
the associated supports for deep excavations. Slopes that are used to stabilise an
excavation during open cut are also considered as temporary works for deep
excavations.

The following retaining wall types are commonly used in Singapore to support
excavations:

• Sheet pile wall


• Soldier pile wall with timber lagging or shotcrete
• Diaphragm wall
• Secant pile wall
• Contiguous bored pile wall
• A combination of sheet pile wall and soldier pile wall with lagging

The above can be further divided into the following categories according to the forms
of support provided:

• Cantilever wall
• Strutted or braced wall
• Tie-back or anchored wall

Steel I-beams are commonly used in Singapore as struts and waling for strutted wall.

1.3 Cause of Failure

The failure of temporary retaining works during excavation can be caused by local
failure and/or global failures. Failure mechanisms which should be considered in the
design are shown in Figure 1-1.

5
Passive Anchorage Failure

Break
R
Break or
bend Tie-back
Outside
R Formation Anchor
Excessive Bending

Kickout Through Tip

(a) Rod failure, anchor failure, bending /pullout (b) Rotational instability
failure or toe kickout

Figure 1-1. Typpical Failure mechanisms that should be considered in design

6
2 DESIGN CONCEPT

2.1 Construction Sequences

Before any analysis can be carried out, a construction sequence needs to be developed.
The bending moments and shear forces in the retaining wall, the strut loading at every
tier of struts and the wall deflection are closely related to the construction sequence.
The reinforcement in diaphragm walls in the permanent stage is also affected by the
construction sequence as part of the earth pressures will produce forces that will be
locked in the walls during construction. Therefore the construction sequence adopted
in the design should be clearly shown in the drawing. Any changes to the construction
sequence should be brought to the attention of the designers, and such changes may
require a re-design of the retaining system.

2.2 Design Assumptions

Typically the design of temporary retaining walls for deep excavations is based on the
assumption of a plane strain condition. This assumption is generally conservative for
areas around corners or bends of retaining walls. However the reviewer should take
note of the following:

• Struts are typically spaced at 6 to 8m interval. The stiffness of the struts should
be calculated on per meter run basis. The total strut force on each strut will be the
reaction from the analysis multiplied by the strut spacing.
• If inclined struts or diagonal struts are used, the stiffness of the struts should be
resolved to the direction perpendicular to the retaining wall. The final strut force
should then be resolved to the axial direction of the struts.
• Where piles or barrettes are within the excavation, the stiffness of the piles or
barrettes, if used in the analysis, should be converted to stiffness per meter run.
The total forces (bending moment and shear force) should be calculated as the
reaction multiplied by the spacing of the piles or barrettes. Cautions should be
exercised when there is soft clay such as the marine clay or peaty clay, as the
clay can be squeezed through the gaps between the piles.

7
3 DESIGN SOIL PARAMETERS

3.1 General

The geotechnical parameters are covered in Table 5/3, Chapter 5 of the Design
Criteria. The Table gives minimum values for common design parameters for the soils
and weathered rocks of Singapore. The Table is re-produced in Table 3.1 below.

Table 3.1: Recommended Design Parameters from Design Criteria


MRTC Classifica- Bulk Coefficient Undrained Effective Effective
(1983) tion based Density of earth cohesion, cu cohesion, c’ angle of
classifica- on BS5930 (kN/m3) pressure at (kN/m2) (kN/m2) friction, φ’
tion (1999) rest (ko) (degrees)

Fill Fill 19 0.5 0 0 30


B B 19 0.5 0 0 30
E E 15 1.0 Figure 3.1 0 5
F1 F1 20 0.7 0 0 30
F2 F2 19 1.0 Figure 3.2 0 22
M M 16 1.0 Figure 3.3 0 22
O O Class A 20 1.0 Note 1 Note 2 Note 2
O O Class B 20 1.0 Note 1 Note 2 Note 2
O O Class C 20 1.0 Note 1 Note 2 Note 2
O O Class D 20 1.0 Note 1 Note 2 Note 2
O O Class E 20 1.0 Note 1 Note 2 Note 2
S1 SI 24 0.8 N/A Note 3 Note 3
S1 SII 24 0.8 N/A Note 3 Note 3
S2 SIII 22 0.8 N/A Note 3 Note 3
S2 SIV 22 0.8 N/A Note 3 Note 3
S3 F.C. 22 1.0 Note 1 10 28
S4 SV 21 0.8 N/A 0 30
S4 SVI 21 0.8 Note 1 0 30
G1 GI 24 0.8 N/A Note 3 Note 3
G1 GII 24 0.8 N/A Note 3 Note 3
G2 GIII 23 0.8 N/A Note 3 Note 3
G2 GIV 23 0.8 N/A 0 30
G4 GV 20 0.8 N/A 0 30
G4 GVI 20 0.8 Note 1 0 30
N/A: Not Applicable.
Note 1: Undrained conditions do not usually apply for deep excavations in these materials, but may be
applicable during tunnelling. The methods outlined in Clough and Schmidt (1981) may be used to
assess if undrained parameters are applicable. Typically, design should be carried out for both drained

8
and undrained parameters, and the more conservative of these designs should be adopted. For
undrained analysis, a value of 5 x N (SPT value in blows/300m) kPa, up to N = 50, may be adopted for
the undrained cohesion in these materials, where applicable.
Note 2: Effective Stress parameters for the Old Alluvium shall be established for each site based on
p’-q plots.
Note 3. Effective Stress parameters for these materials should be derived from site-specific data. The
Geological Strength Index method (Hoek and Brown, 1997) is considered appropriate for this.

Undrained Cohesion, Cu (kPa)


0 10 20 30 40 50 60
0
Depth below ground level (m)

10

15

20

25

Figure 3.1. Undrained cohesion for Estuarine clay (E)

U n d ra in e d C o h e s io n , C u ( k P a )
0 10 20 30 40 50 60
0

5
Depth below ground level (m)

10

15

20

25

Figure 3.2. Undrained cohesion for fluvial clay (F2)

9
Und ra in e d c o he sion , C u (k P a )
0 10 20 30 40 50 60 70
0
5
Depth below ground level (m)

10
15
20
25
30
35
40

Figure 3.3. Undrained cohesion for marine clay (M)

It is to be noted that the parameters given in the Table 3.1 are intended to represent the
lower bound design values of the soil properties as it exists in-situ, as required under
BS8002. Higher design values can be used if the designer can demonstrate as being
appropriate, based on the results of the soil investigation for the project.

The following sections give some examples of the design values that have been
published for design and research. These sections can serve as a guide when reviewing
proposed design parameters submitted from external parties. Where proposed design
parameters are outside the commonly adopted range, the reviewer should pay more
attention to examine the justification submitted from external parties in detail.

3.2 Geotechnical Design Parameters for Phase I/II MRT Projects

Dames and Moore (1983) summarised the geotechnical properties of Singapore’s soils
based on soil investigation data obtained from North-South and East-West MRT lines.
A range of some soil parameters can be found in this report. For ease of reference,
Table 3.2 summarises the recommended design parameters from the report.

Table 3.2. Design parameters recommended by Dames & Moore (1983).


Soil Type Parameters Value / Comment
B Bulk density 17 kN/m3
Relative density 35%
Modulus, Es = 1/mv 10 MPa
Effective cohesion, c’ 0
Effective angle of friction, φ’ 30 degrees
Coefficient of earth pressure at rest, ko 0.5
Permeability 5 x 10-7 to 5 x 10-3 m/s

10
Table 3.2. Continued
Soil Type Parameters Value / Comment
E Bulk density 14 kN/m3
Undrained shear strength, cu • Depth: 0 – 5m: 5 kPa
• Depth 5m +: increase linearly with depth
to 50 kPa at 25m depth.
Sensitivity 2
Undrained Modulus, Eu 200 - 400 cu
Effective cohesion, c’ 0
Effective angle of friction, φ’ 5 degrees
Compression index, cc cc = 0.0088 (wn – 14), where wn = natural
water content
Coefficient of 2nd compression, cα cα = 0.04 cc
Coefficient of consolidation, cv 10 m2/year
Coefficient of earth pressure at rest, ko 0.9 – 1.0
Permeability • Design value: 1 x 10-9 m/s
• Range: 8 x 10-11 to 2 x 10-7 m/s
F1 Bulk density 18 kN/m3
Relative density • Depth 0 to 25m: 35%
• Depth 25m +: 50%
Modulus, Es = 1/mv • Depth 0 to 25m: 10 MPa
• Depth 25m+: 20 MPa
Effective cohesion, c’ 0
Effective angle of friction, φ’ • Depth 0 to 25m: 30 degrees
• Depth 25m+: 34 degrees
Coefficient of earth pressure at rest, ko 0.43 to 0.67
Permeability 1.0 x 10-7 to 1.0 x 10-4 m/s
F2 Bulk density 19 kN/m3
Undrained shear strength, cu • Depth: 0 – 10m: 20 kPa
• Depth 10m +: increase linearly with depth
to 50 kPa at 25m depth.
Sensitivity 4
Undrained Modulus, Eu 14.5 MPa
Effective cohesion, c’ 0
Effective angle of friction, φ’ 22 degrees
Compression index, cc cc = 0.2
Coefficient of 2nd compression, cα cα = 0.05
Coefficient of consolidation, cv 10 m2/year
Pore pressure parameter, A 0.4
Coefficient of earth pressure at rest, ko 0.75 – 1.0
Permeability • Design value: 1 x 10-9 m/s
• Range: 4 x 10-10 to 2 x 10-7 m/s
M Bulk density 15 kN/m3
Undrained shear strength, cu • Depth: 0 – 6.7m: 10 kPa (except at
Marina area)
• Depth 6.7m+: increase linearly with depth
to 60 kPa at 40m depth.
Sensitivity 5
Undrained Modulus, Eu 200 cu
Effective cohesion, c’ 0
Effective angle of friction, φ’ 22 degrees

11
Table 3.2. Continued
Soil Type Parameters Value / Comment
M Compression index, cc cc = 0.54 (eo – 0.15), where eo is initial void
ratio
Average value: 0.85
Coefficient of 2nd compression, cα cα = 0.175 cc
Average value: 0.15
Coefficient of consolidation, cv 2 m2/year
Pore pressure parameter, A 0.8 – 1.0
Coefficient of earth pressure at rest, ko 0.6 – 0.7
Permeability • Design value: 1 x 10-9 m/s
• Range: 4 x 10-11 to 4 x 10-9 m/s
O Bulk density 20.5 kN/m3
Undrained shear strength, cu • Depth* 0 – 5m: 45 kPa
• Depth* 5 - 15m: 100 kPa
• Depth* 15m+: 150 kPa
Undrained Modulus, Eu • Depth* 0 – 15m: 50 MPa
• Depth* 15m+: 75 MPa
Effective cohesion, c’ 0
Effective angle of friction, φ’ 35 degrees
Coefficient of consolidation, cv 25 m2/year
Pore pressure parameter, A 0.1
Coefficient of earth pressure at rest, ko 0.75 – 1.0
Permeability • Design value: 1 x 10-9 m/s
• Range: 1 x 10-10 to 1 x 10-6 m/s
(* depths measured from top of Soil Type O.)
S2 Bulk density 22 kN/m3
Undrained shear strength, cu 200 kPa
Undrained Modulus, Eu 200 MPa
Permeability • Design value: 5 x 10-9 m/s
• Range: 1 x 10-10 to 1 x 10-7 m/s
S3 Bulk density 22 kN/m3
Undrained shear strength, cu 100 kPa
Undrained Modulus, Eu 100 MPa
Effective cohesion, c’ 10
Effective angle of friction, φ’ 28 degrees
Compression index (cc or cr) 0.03
Coefficient of consolidation, cv 30 m2/year
Pore pressure parameter, A 0.1
Coefficient of earth pressure at rest, ko 0.75 – 1.0
Permeability • Design value: 1 x 10-9 m/s
• Range: 1 x 10-10 to 1 x 10-5 m/s
S4 Bulk density 20.5 kN/m3
Undrained shear strength, cu • Depth* 0 – 3m: 45 kPa
• Depth* 3 - 12m: 85 kPa
• Depth* 12m+: 150 kPa
Remoulded shear strength 45 kPa
Undrained Modulus, Eu • Depth* 0 – 3m: 7 MPa
• Depth* 3 - 12m: 40 MPa
• Depth* 12m+: 150 MPa
Effective cohesion, c’ 0
Effective angle of friction, φ’ 30 degrees

12
Table 3.2. Continued
Soil Type Parameters Value / Comment
S4 Coefficient of consolidation, cv 40 m2/year
Pore pressure parameter, A 0.07
Coefficient of earth pressure at rest, ko 0.75 – 1.0
Permeability • Design value: 1 x 10-9 m/s
• Range: 5 x 10-8 to 2 x 10-10 m/s
(* depths measured from top of Soil Type
S4.)
G4 Bulk density 18.5 kN/m3
Undrained shear strength, cu
- Silty CLAY • Depth* 0 – 5m: 35 kPa
• Depth* 5 - 20m: 55 kPa
• Depth* 20m+: 100 kPa
- Clayey SILT • Depth* 0 - 20m: 55 kPa
• Depth* 20m+: 100 kPa
Remoulded shear strength 0.5 cu
Undrained Modulus, Eu • Depth* 0 – 5m: 5 MPa
• Depth* 5 - 20m: 8 MPa
• Depth* 20m+: 15 MPa
Effective cohesion, c’ 0
Effective angle of friction, φ’ 30 degrees
Coefficient of consolidation, cv 40 m2/year
Pore pressure parameter, A 0.17
Coefficient of earth pressure at rest, ko 0.8 – 1.0
Permeability • Design value: 1 x 10-8 m/s
• Range: 4 x 10-10 to 2 x 10-7 m/s
(* depths measured from top of Soil Type
G4.)

3.3 Kallang Formation

The Kallang Formation covers a wide range of soil types, which include soils of the
Marine, Alluvial, Littoral, Transitional (Estuarine) and Reef members and are all of
post Pleistocene age. The deposits vary in composition and occur as an infill to
ancient buried valleys formed in strata of the Bukit Timah Granite, Jurong Formation
and the Old Alluvium during sea level changes in the last 5000 years. The maximum
depth of the Kallang Formation has been proven to a thickness of some 60m.

Table 3.3, Table 3.4 and Table 3.5 list the published design parameters for the
Singapore Marine Clay (M), the fluvial sand (F1) and fluvial clay (F2), respectively.

13
Table 3.3 Published design parameters for Marine Clay

Undrained Effective Effective Undrained


Compression
Reference Cohesion, Cohesion Angle of Modulus
Index cc
cu (kPa) c’ (kPa) Friction Eu (MPa)
φ’ (deg.)

Nicholson (1987) cu/pc = 0.23 - - - 450 cu


Poh & Wong (1998) - - 20-24 0.83 -
Pitts (1985) - 0 22 - 100-200 cu
Wong, et al (1997) - - - - 200 cu
where pc = pre-consolidation pressure

Table 3.4 Published design parameters for fluvial sand (F1)

Effective Effective Modulus E


Reference Cohesion Angle of (MPa)
c’ (kPa) Friction
φ’ (deg.)

Chu et al (2000) 0 32-36 12.3 – 43.5


Flanagan et al (1999) 0 30 15
Pitts (1985) - - 10 - 20
Tan et al (1980) - 30 - 35 -

Table 3.5 Published design parameters for fluvial clay (F2)

Undrained Effective Effective Undrained


Compression
Reference Cohesion, Cohesion Angle of Modulus
Index cc
cu (kPa) c’ (kPa) Friction Eu (MPa)
φ’ (deg.)

Flanagan et al (1999) - 0 30 - 15
Pitts (1985) 20 - 120 0 22 - 15

3.4 Old Alluvium (OA)

The Old Alluvium, which is about 2 to 7 million years old, extends from southern
Johor to the east of Singapore. The thickness of this formation is generally very great
and has been proved to a depth up to 195 metres.

The formation consists of coarse angular quartz feldspar sand and gravel which have
probably been derived from granite and laid down in a fluvial environment. Joints
have been observed in this formation. However there is no evidence of any faulting.
Weathering and decomposition of the feldspar in the upper layers has produced
cohesive soils of clayey silty sand or sandy clay with some fine gravel. It should be
noted that the clay content of the Old Alluvium tends to increase with increasing
degree of the weathering and proximity to the ground surface. There is generally some
cementation; however, the degree of cementation varies. Where the SPT N-value is

14
low, the OA has been weathered after decomposition, and much of the cementation
has been lost.

Wong et al (2001) conducted a study on the data obtained from NEL projects for OA
soils. The study characterizes the OA into 3 main zones according to the degree of
weathering and SPT blow counts as follows:

OAI: N = 5 to 25 (residual soil zone)


OAII: N = 26 to 99 (weathered zone)
OAIII: N = 100 or more (cemented zone)

The shear strength of the three groups as recommended by Wong et al (2001) are
given in Table 3.6.

Table 3.6 Recommended Design parameters for OA, Wong et al (2001)


OAI OAII OAIII
Property
(N = 5 to 25) (N = 26 to 99) (N ≥ 100)
Undrained shear strength (kPa) Cu ~ 5.4 N
Effective apparent cohesion (kPa) 0 5 25
Effective friction angle (degrees) 35 35 35

3.5 Jurong Formation

The Jurong Formation, which is about 190 million years old (Mid Triassic to Lower
Jurassic) comprises alternating layers of mudstone, siltstone and sandstone which are
typically tightly folded and fissured. The thickness of the layers varies and the fissures,
which are often randomly oriented, are not governed by the bedding planes. The fine
grained mudstone is intensely fissured. The rock has been heavily weathered in its
upper parts.

Within the Jurong Formation a ‘boulder bed’ has been described by Pitts (1984). This
is believed (Shirlaw et al, 1990 and Han et al, 1993) to be a landslide deposit consisted
of often massive boulders of strong sandstone in a stiff fissured clay matrix. This
boulder bed is now known as the Fort Canning Boulder Bed, Shirlaw et al (2003).

Table 3.7 gives some examples of published data on the properties of the Jurong
Formation.

15
Table 3.7 Published design parameters for Jurong Formation

Undrained Effective Effective Undrained


Weathering
Reference cohesion, cohesionc cngle of Modulus
Description
cu (kPa) ’ (kPa) friction φ’ Eu (MPa)
(deg.)

Orihara et al (1999) S2 200 40 30-35 300-600


Leong et al (2002) S4 5* SPT N - - Eu/cu = 40 -
400
Lim (1995) S4 - 19-50 24-40 -
Lo et al (1988) S4 - 6 32 -
Orihara et al (1999) S4 65-200 0-20 33 30-100
Rahardjo (2000) S4 5-9 29-32
Yong, et al (1985) S4 - 12 13-40
Zhu (2000) S4 - 10-30 24-40 -

As required by the Design Criteria, the effective stress parameters for weathered rocks
should be derived from site-specific data. Typical assessed parameters for the
weathered rocks of the Jurong Formation based on the Geological Strength Index
method, Hoek and Brown, 1997 is listed in Table 3.8.

Table 3.8. Typical assessed parameters for rocks of Jurong Formation


Rock RQD
c’ φ’
(kPa) (degrees)
Mudstone <25% 40 30
Siltstone <25% 40 30
Sandstone <25% 40 45
Mudstone >25% 40 35
Siltstone >25% 40 35
Sandstone >25% 120 45

3.6 Bukit Timah Granite Formation

The Bukit Timah Granite is of lower to Mid Triassic age (220 million years) and
includes an entire suite of plutonic acid igneous rocks, particularly granite, adamellite
and granodiorite. The granite has been cut by numerous basic and acidic dykes which
post-date the granite and are believed to be of mid-Triassic age.

Weathering has produced a combination of residual soil and completely decomposed


rock near the surface of the granite that is underlain by the highly and moderately
weathered granite which, in turn is underlain by slightly weathered and fresh granite.
The residual soil and the completely decomposed rock can be described generally as
clayey silt, silty sand and silty clay that tend to be characterised by a reddish brown
colour.

Table 3.9 lists typical parameters for the residual soils of the Bukit Timah Granite
Formation.

16
Table 3.9 Published design parameters for the residual soils of the Bukit Timah Granite Formation
References Undrained Effective Effective Undrained
cohesion, cu cohesion c’ angle of Modulus
(kPa) (kPa) friction φ’ Eu (MPa)
(Degrees)
Leong et al (2002) 5* SPT N - - Eu/cu = 30 -
300
Poh et al. (1985) - 0-42 20-35
Yang & Tang (1997) - 5-10 35-40
Tan et al. (1987) - 0-40 30-35
Rahardjo (2000) - 12-50 29-33
0-14 27-31
KarWinn et al. (2001) - - 20-40
Zhou (2001) - 7 32

17
4 STABILITY CHECKING

4.1 Base Heave

Temporary works design for deep excavation shall include adequate precautions
against base heave.

In deep excavations in soft clay, such as the Singapore marine clay, there is a potential
for base failure accompanied by a consequent sinking of the surrounding ground and
failure of the support system. For an excavation that is adequately supported by struts
to prevent horizontal displacement of the walls, the soft clay at the base is under the
loading due to the weight of the surrounding soil masses. This loading tends to push
the clay towards and up into the excavation, see Figure 4-1.
Surcharge, q

Figure 4-1. Base heave checking

Bjerrum & Eide (1956) presented their study of the problem of base stability of deep
excavations in soft clays. They proposed that the factor of safety against base failure in
soft clays should be checked as below:

FoS = Nccu / (γD +q)

Where D is the depth of excavation, q is the surcharge, cu is the undrained shear


strength of the clay at the base, γ is the unit weight of the clay and Nc is the bearing
capacity factor.

Other formulae for various excavation dimensions and geological strata can be found
in the LTA in house training notes by Prof. Wong Kai Sin et al (2003).

Studies by Wong (2003) have shown that low factors of safety against basal failure
will produce high ground movements and wall deflection, see Figure 4-2.

18
Figure 4-2. Wall deflection vs. factor of safety against basal heave

CP18, Cl. 8.3.3.3 requires a minimum factor of safety of 1.5 against overstress.
Checking should be carried out not only for the final stage of excavation, but also for
intermediate stages where layers of soft clay exist between ground level and the final
formation level.

4.2 Hydraulic Uplift

Hydraulic uplift problems arise when water under pressure is trapped in permeable
strata beneath clays below the excavation, and if the water pressure exceeds the weight
of the overburden, the base of the excavation will be blown up, causing the base to
lose its strength, see Figure 4-3.

Standpipe

k1
γD D
k2
U
Figure 4-3. Checking for hydraulic uplift.

CP18, Cl. 8.3.3.3 requires the design of deep excavations in water bearing soils to
consider the possibility of hydraulic uplift. Where necessary and appropriate, i.e. there
will be no major impacts on adjacent properties, services and utilities, ground water
lowering methods should be used to lower the external head of ground water.

Davies (1984) reported a base failure in a deep excavation due to hydraulic uplift. The
excavation was in the residual soils of the Bukit Timah Granite, and the total
excavation depth was 8.0m. The rock level was at 13m below the ground surface. He
reported that the base of the excavation was reduced to a slurry when the excavation
was at about 6.5m deep. Construction traffic sank into the base of the excavation and it

19
was only possible to walk across the site on planks. To allow work to proceed, deep
wells were sunk to the permeable zone just above the rock-head, and the excavation
base became quickly firm and dry.

Excavations in the residual soils of the Bukit Timah Granite are susceptible to this type
of failure. Weathering tends to reduce the upper levels of the soil to a clayey material.
However the clay content decreases and the permeability of the soil increases with
depth. It is difficult to install retaining walls deep enough in the rock to reduce the
hydraulic gradient by lengthening the seepage path. To prevent failure due to hydraulic
uplift, the water pressure in the underlying permeable layer should be relieved by
using deep wells or drainage pipes prior to reaching the critical depth.

4.3 Toe-in Depth

Toe-in depth is to check the adequacy of the wall embedment below the formation
level. The method of checking is given in NAVFAC DM7.2 and is re-produced in
Figure 4-4.

DEFLECTED POSITION
SHEETING OR SOLDIER
PILES WITH LAGGING

Struts
H Ms
A
PA
l2
l1
PA1
Pp
D

PA = TOTAL RESULTANT ACTIVE PRESSURE

PA1 = RESULTANT ACTIVE PRESSURE BELOW POINT A

Figure 4-4. Checking for adequacy of toe-in depth.

PP * l 2
Factor of safety =
PAl * l1 - M S

where MS is the allowable moment in the retaining wall, PP is the force due to the
passive soil pressure and PAl is the force due to the active soil pressure below the
lowest level of struts.

CIRIA Report 104 (Padfield & Mair, 1984) recommends the following factors of
safety:

20
• For φ’ < or = 20o, FOS = 1.2
• For φ’ = 20o – 30o, FOS = 1.2 – 1.5
• For φ’ > or = 30o, FOS = 1.5

Where the mobilisation factors are used in accordance with BS8002, the minimum
required factor of safety can be 1.0.

4.4 Slope Stability for Open Excavation

The factor of safety for slope stability for open excavation should be checked by
carrying out slope stability analysis. Various software, such as SLOPE/W can be used.
The two major critical parameters are the cohesion of soil and the phreatic surface.

The cohesion should be justified by soil investigation data. Typically the ground water
table will be at 1 to 2 m below the ground level. Under clement weather the water table
can rise to the ground level, thus should be checked.

For permanent slope the factor of safety shall be a minimum of 1.5. For temporary
slopes, the factor of safety can be reduced, based on consequence of failures. A guide
is shown in Table 4-1, based on Geotechnical Manual for Slopes, Geotechnical
Control Office, Hong Kong.

Table 4-1. Factors of safety for slopes.


Recommended factor of safety against loss
Risk of life for a ten-year return period rainfall
Negligible Low High
Recommended factor of Negligible >1.0 1.2 1.4
safety against economic
loss for a ten-year Low 1.2 1.2 1.4
return period rainfall High 1.4 1.4 1.4

21
5 DESIGN ANALYSIS

5.1 Overview of Analytical Methods

Methods for analysing temporary retaining works can be broadly classified into the
following types:

• beam on elastic foundation


• finite element method

5.2 One-Dimensional Finite Element Method

This method assumes a retaining wall as an elastic beam founded on an elastic


foundation to derive the solution of the governing differential equations. The elastic
beam is assumed to generate a reactive earth pressure proportional to its deflection, i.e.
Winkler’s hypothesis, as a bed of springs. The commonly used commercial software,
which adopts this approach, is as follows:

• WALLAP
• FREW
• REDO

An advantage of the ‘beam on elastic foundation’ approach is its ability to account for
structure flexibility and soil stiffness. Thus the effects of stress redistribution in soil as
a result of differential structural deflections are accommodated. Although the
magnitude of the shear forces and bending moments in the wall and the strut or anchor
loads are not very sensitive to the values of spring stiffness used in the analysis, the
predicted deformations of the supporting wall are.

Typically the software will require the user to input the active and passive coefficients
of earth pressure, ka and kp. While it is reasonable to assume some friction between the
wall and the soil, it should however be noted that the friction between the wall and the
soil for kp should not be too high, as large wall movement is required to achieve the
corresponding kp value. CIRIA Report 104 (Padfield & Mair, 1984) recommends the
following maximum effective wall friction, δ:

• For active zone, δ = 2/3 φ’


• For passive zone, δ = 1/3 φ’

The reviewer should also check to ensure that the construction sequences adopted in
the analysis are the same as those in the construction drawings.

The following limitations of the ‘beam on elastic foundation’ approach should be


noted:

22
• There is an inherent difficulty in determining the appropriate spring stiffness
(constants) of the soil for analysis as these are not fundamental soil properties. The
spring constants are dependent on the Young’s modulus of the soil and also on the
dimension of the excavation problems being analysed. Users should not use
correlation of spring constants with undrained strength of soil as basis of deriving
the spring constants for analysis. Spring constants, if used, should be based on
Young’s modulus of the soil, which is a soil property and independent of
dimension of the excavation.

• The method can only predict the movement of the wall. Ground movements can
not be predicted by using such software.

5.3 Two Dimensional Finite Element Method

The finite element method offers the designer an analytical tool that can simulate the
complex facets of the strutted or tied-back wall except unquantifiable variables such as
workmanship or geological uncertainties. The commonly used commercial software is
as follows:

• PLAXIS
• SAGE CRISP
• EXCAV97

The assumption of plane strain condition usually adopted in finite element analysis of
strutted or anchored walls cannot be arbitrarily extrapolated to many practical
problems, particularly those involving soldier pile walls. Diaphragm walls, however,
were shown to approximately satisfy the conditions assumed in a plane strain analysis,
except at the corners of the retaining system. To simulate discontinuous wall elements
such as soldier piles, struts or tie-backs, the stiffness has to be represented on a ‘unit’
length basis. Analyses using this representation yield results that are characteristic of
the average values between those at the supports and those at the mid-point between
the supports.

The two single most important variables required in a finite element analysis are the
stiffness of the ground, which mainly affects the displacements, and the magnitude of
the initial horizontal stresses, which affect both the displacements and the bending
moments.

The in-situ horizontal stress in the ground is related to the in-situ vertical stress by the
coefficient of earth pressure at rest, ko. From conventional soil mechanics, it is well
known that ko, being dependent on stress history, is not a fundamental soil property.
For a soil in the normally consolidated (i.e. low ko) state, the stress change required to
reach the active state is small compared with that needed to mobilise passive
resistance. The contrary is true for a soil in the over-consolidated (i.e. high ko) state.
Therefore, the horizontal displacements of the wall and surface movements behind the
wall are dependent on the magnitude of the in-situ stress. The existence of a high in-
situ horizontal stress would likely result in large wall displacements and surface

23
settlements. For excavated walls in soils with a high initial ko value (≥2), the prop
forces and wall bending moments can greatly exceed those predicted by limit
equilibrium methods, Potts & Fourie (1984).

The advantage of the finite element method in the analysis of earth retaining structures
lies in its ability to predict both earth pressures and deformations with a minimum of
simplifying assumptions. Both structure and soil are considered interactively, so that
the effects of structural flexibility are taken into account.

Limitations in using the method primarily derive from a user’s inability to prescribe
appropriate constitutive models for the soil, and to determine the most appropriate
parameters needed for the constitutive models. Another limitation is that, in a 2-D
model, any element simulating a pile is in effect a wall, as the model is typically based
on plane strain condition. This simplification may underestimate the ground
movement.

The effects of corners can not be reflected in a 2-D analysis. This may lead to over-
conservative (excessive) prediction of ground movement and wall deflection at corners
of deep excavations. The significance or otherwise of the corner effects depends on
three factors: the length to depth ratio of the excavation, the depth to a relatively stiff
stratum and the stiffness of the strutting system, Lee et al (1998).

5.4 Drained Analysis vs. Undrained Analysis

A saturated soil comprises two phrases: the soil particles and the pore water. The pore
water pressure is constant at a value governed by a constant position of the water table.
This initial value is called the static pore water pressure. When a soil element is
loaded, pore water pressure is increased above the static value immediately. This
increase in pore water pressure causes a pressure gradient, resulting in a transient flow
of pore water towards a free draining boundary of the soil layer. This flow or drainage
will continue until pore water pressure again becomes equal to the value governed by
the position of the water table. The component of pore water pressure above the static
value is known as the excess pore water pressure. When the excessive pore water
pressure is dissipated the soil is said to be in the drained condition. Prior to dissipation
of the excess pore water pressure the soil is said to be in undrained condition.

The normal total stress applied to a soil element can be separated by means of the
principle of effective stress. The effective stress is the component of normal stress
taken by the soil skeleton. It is the effective stress that controls the volume and
strength of the soil. For saturated soils, the effective stress can be calculated as the
total normal stress minus the pore water pressure.

The conventional approaches to analysis of short term behaviour of excavations in clay


assume undrained response for the soil and no excess pore water pressure dissipation.
The question that is often asked is under what conditions the undrained approach does
prevail.

24
Osaimi & Clough (1979) reported their analysis of consolidation around excavations.
The results are shown in Figure 5-1.
k = 1.0E-06 m/s k = 1.5E-07 m/s
k = 1.0E-08 m/s k = 1.0E-10 m/s

120
% Consolidation at end of excavation

100

80

60

40

20

0
0 0.5 1 1.5 2 2.5 3 3.5
Rate of excavation (m/day)
Figure 5-1. Degree of consolidation at the end of one-dimensional excavation into homogeneous
clay as a function of excavation rate and permeability, k (Osaimi & Clough, 1979)

For a soil with a permeability of 1 x 10-10 m/s, essentially zero consolidation occurs for
even the slowest soil removal rate of 0.03 m/day. Thus for this example, undrained
behaviour is appropriate for analysis of the end of construction state. For higher soil
permeability, the degree of consolidation is strongly a function of the rate of
excavation, with significant drainage indicated for several cases. In the field the
presence of drainage lenses of sand and silt will probably speed up consolidation over
that shown in Figure 5-1. Practically it means that both undrained and drained
conditions may need to be considered in design analysis for excavations, with the
exception of excavations in marine clays or fluvial clays.

5.5 Total Stress Approach vs. Effective Stress Approach

There are alternative methods of doing drained or undrained analysis. These are:

i) Total stress approach


(a) Undrained – specify Eu and νu (νu ≈ 0.49)

ii) Effective stress approach


(a) Undrained – specify E’, ν’and kw, where kw is the bulk modulus of water,
kw ≈ 100K’ and K’ = E’/{3(1-2ν’)}
(b) Drained – specify E’ and ν’

iii) Consolidation analysis

25
(a) Always specify E’, ν’. After a short time interval, undrained response is
obtained, and after a long time interval a drained response is obtained.

The advantage of effective stress approach is that pore water pressure changes due to
undrained loading are calculated and printed explicitly.

Typically Mohr-Coulomb (M-C) failure criteria are adopted for analysis, as the
parameters, cu or c’ and φ’ for the criteria are well defined for soils commonly
encountered. For undrained analysis, cu should be specified and for drained analysis, c’
and φ’ should be specified. Based on the research work by Prof Wong Kai Sin of NTU
(2003), the following notes should be taken into account when carrying out a FE
analysis or reviewing an submission using FE analysis:

• Undrained analysis by total stress approach (Eu and νu ≈ 0.49) using total stress
M-C parameter (cu) can produces reliable results provided an appropriate Eu/cu
ratio is adopted for soft clays and stiff clays. The analysis can provide
reasonable matches at all stages for deep excavation in stiff clays, but it cannot
provide good matches at all stages of excavation for soft clays.
• Drained analysis by effective stress approach (E’ and ν’) using effective stress
M-C parameters (c’ and φ’) can produce reasonable results for certain stress
paths and questionable or unreliable results in some other stress paths.

5.6 Note on Using Effective Stress Approach for Undrained Analysis

Both SAGE CRISP and PLAXIS have the facilities to carry out undrained analysis by
using effective stress parameters. This is possible by using the kw, the bulk modulus of
water. With kw, it is possible to calculate the rate of excess pore water pressure, thus
allowing the determination of the total stress in a soil element. As the analysis is in
effect an undrained analysis, total stress parameter for M-C criteria, cu should be used.

In SAGE CRISP, the bulk modulus of water kw has to be specified by the user as an
input for undrained analysis using effective stress approach. The recommended value
is kw ≈ 100K’, where K’ = E’/{3(1-2ν’)} to ensure the simulation of pore water as
incompressible material compared with the soil skeleton.

In PLAXIS, the calculation of kw is automatically done. The user will only get a
warning when the input value of ν’ is larger than 0.35. It should be noted that in
PLAXIS, whenever the Material Type is set to Undrained, effective stress parameters,
E’ and ν’ should be specified in the input, and ν’ should be less than 0.35. If M-C
failure criteria is adopted, cu should be used as input for strength parameter, instead of
c’ and φ’.

In summary, there are a few methods available to carry out undrained analyses in the
program PLAXIS:

26
• Method A - Undrained setting, Mohr Coulomb soil model, effective stress strength
parameters, c' and φ', E' and µ' (< or = 0.35). Method A shall not be used for
analyses of deep excavations in soft clays, such as the marine clay.
• Method B - Undrained setting, Mohr Coulomb Model, undrained strength
parameter, cu, E' and µ' (< or = 0.35).
• Method C - Drained setting, Mohr Coulomb model, undrained strength parameter,
cu , Eu and µu = 0.495.
• Method D - Undrained setting, soft clay model, such as Cam Clay model, c', φ', λ,
κ.

Commonly used approaches for undrained analyses in SAGE CRISP are:

• Effective Stress Approach, Mohr Coulomb model, specify cu, E', µ' and kw (kw
>>soil bulk modulus, k' to avoid numerical error). (Equivalent to Method B)
• Total Stress Approach, Mohr Coulomb model, specify cu , Eu, µu = 0.495 and kw
= 0.0, where kw is the bulk modulus of water. (Equivalent to Method C)
• Effective Stress Approach, soft clay model, such as Cam Clay model, λ, κ, M, ecs,
µ' and kw (kw >>soil bulk modulus, k' to avoid numerical error). (Equivalent to
Method D)

5.7 Sensitivity Analysis

There are great dangers in relying on ‘one-off’ analyses, in which a single set of
geotechnical parameters are used, and the results of the analysis then taken as ‘the
prediction’ of deformations, loads and stresses that will actually occur. It is essential to
vary the input parameters, within a reasonable range corresponding to those actually
determined from the ground investigation, and to examine critically the effects of such
variations on the computed deformations, loads and stresses.

It is important to distinguish design from analysis – the former requires a full


consideration of all the factors that might affect the behaviour of the underground
works, whereas the latter must only be regarded as a tool to aid the design process,
albeit a very important tool. The temptation to rely on a single finite element analysis
as the basis for design should be resisted. A suite of analyses, used to explore the
influence of parameter variations, is necessary before the results can be used with
confidence in the design.

Sensitivity analyses shall be performed as part of the design to demonstrate that the
design and the model are not unduly sensitive to variations in any of the following
input parameters:

• Shear strength
• Soil stiffness
• Reduced wall stiffness due to cracking
• Pre-loading forces

27
• Over-excavation
• One-strut failure
• Consolidation parameters
• JGP strength and stiffness (where applicable)

It shall also be demonstrated that the model is not unduly sensitive to any other
variable for which assumptions are made within the FE or FD model.

The design shall be based on the envelope of upper and lower bounds of the sensitivity
analyses, with appropriate factor of safety for structural design of the walls and strut
system.

5.8 Back Analysis

Back analyses should be carried out comprehensively with careful reviews of ALL
design assumptions and the actual monitored behaviour of the structures.

While curve fitting of the deflection is an important part of the back analysis, other
monitored parameters should also be checked in the back analysis. There should be a
clear evidence of improvement between computed and measured behaviour to date, i.e.
from the beginning up to the stage where back analysis is carried out.

The review of design assumptions during back analyses should include


appropriateness of the models used in the original design, soil parameters, soil profiles
encountered, construction sequences, drainage conditions on site and the wall stiffness.
Any changes or revisions in the back analysis should not be arbitrary. There should be
a clear rationale and substantiation, for example field or laboratory tests or additional
evidence from construction records, for all changes in the model parameters.

5.9 Use of Sacrificial JGP Layers

Traditionally geotechnical design parameters are selected conservatively, i.e. design


values (strength and stiffness) are selected towards the lower bound values. For the
design of a sacrificial jet grouted slab, it is necessary to carry out sensitivity analyses
for the upper bound strength and stiffness values of the jet grouted slab. This is to cater
for the worst loading condition for the struts above the sacrificial jet grouted slab.
When the slab is removed, the load taken by it will be transferred to the struts above. If
lower bound strength and stiffness values are used, this will underestimate the strut
load for the struts above the sacrificial jet grouted slab. Close monitoring of the struts
and the walls should be carried out during the removal of the sacrificial JGP layers.

5.10 Treatment of Berm

Slope stability analysis should be carried out to ensure that the berm is stable during
the whole period of the construction.

28
More often than not, the berm size is limited, and it is unlikely full passive resistance
can be developed within the berm. The method recommended by Fleming (1981) can
be adopted to model the berm in the design and analysis of the retaining wall. The
empirical method is to treat the berm as causing an increase in the effective ground
level on the passive side of the wall. In the method the height of the berm is treated as
not more than 1/3 of the berm width as shown in Figure 5-2 and the effective ground
level is then taken as half the height of the berm at the point where it contacts the wall.

Actual Berm Size

Effective ground
level in analysis

3 1 H/2
H
H/2

Figure 5-2. Empirical method of berm treatment in analysis

Another alternative method is to convert the berm to an effective surcharge acting on


the potential passive failure zone. In this method the effective weight of the berm is
calculated and is then distributed over the approximate width of the passive failure
mechanism at the general final excavation level, see Figure 5-3, thus increasing the
passive pressure available accordingly.

Effective surcharge

Passive zone

Figure 5-3. Alternative method of treatment of berm in analysis

5.11 Strut Design

The figures in this section give the design charts for struts commonly used in
supporting excavations. The designs in accordance with both BS449 and BS5950 are

29
produced. In addition to the axial force used in producing the charts, the following
loads are also considered:

• Self-weight of strut
• Vertical superimposed load of 1kN/m
• Horizontal superimposed load of 1kN/m

In structural design of struts, walers and their connections, mixed use of different
codes, such as struts/walers are designed to BS449 and their connection details to
BS5950 should not be adopted. The design for all elements (struts, walers and their
connections) should be carried out consistently in accordance with the same code, i.e.
either BS449 or BS5950.
1200
Strut Capacity (T/Strut), SLS BS 449

1100
H200x200x65.7
1000

900
H200x200x57.8
800
H200x200x49.9
700
H200x200x41.4
600

500

400

300

200

100

0
0 2 4 6 8 10 12

Effective Length L (m )
Design Chart for Strut Capacity

Figure 5-4. Design chart for strut size H200x200 in accordance with BS449.

1800
Strut Capacity (T/Strut), SLS BS 449

H250x250x98.1
1600

1400 H250x250x81.6

H250x250x72.4
1200 H250x250x71.8

H250x250x66.5
1000

800

600

400

200

0
0 2 4 6 8 10 12 14

Effective Length L (m )
Design Chart for Strut Capacity

30
Figure 5-5. Design chart for strut size H250x250 in accordance with BS449.

2800
Strut Capacity (T/Strut), SLS BS 449

2600 H300x300x147
2400
H300x300x130
2200 H300x300x125
2000
H300x300x106
1800 H300x300x105
1600 H300x300x94
H300x300x93
1400 H300x300x87

1200
1000
800
600
400

200
0
0 2 4 6 8 10 12 14

Effective Length L (m )
Design Chart for Strut Capacity

Figure 5-6. Design chart for strut size H300x300 in accordance with BS449.

12000
Strut Capacity (T/Strut), SLS BS 449

11000 H400x400x605

10000

9000

8000
H400x400x415
7000

6000

5000 H400x400x283
H400x400x235
4000 H400x400x232
H400x400x200
H400x400x197
3000 H400x400x172
H400x400x147
2000 H400x400x140

1000

0
0 2 4 6 8 10 12 14

Effective Length L (m )
Design Chart for Strut Capacity

Figure 5-7. Design chart for strut size H400x400 in accordance with BS449.

31
2000
Strut Capacity (T/Strut), ULS BS 5950

1800
H200x200x65.7
1600
H200x200x57.8
1400

H200x200x49.9
1200

1000 H200x200x41.4

800

600

400

200

0
0 2 4 6 8 10 12

Effective Length L (m )
Design Chart for Strut Capacity

Figure 5-8. Design chart for strut size H200x200 in accordance with BS5950.

3200
Strut Capacity (T/Strut), ULS BS 5950

2800 H250x250x98.1

2400 H250x250x81.6
H250x250x72.4

2000 H250x250x71.8

H250x250x66.5
1600

1200

800

400

0
0 2 4 6 8 10 12 14

Effective Length L (m )
Design Chart for Strut Capacity

Figure 5-9. Design chart for strut size H250x250 in accordance with BS5950.

32
5000
Strut Capacity (T/Strut), ULS BS 5950

4500
H300x300x147
4000
H300x300x130
H300x300x125
3500
H300x300x106
3000 H300x300x105
H300x300x94
H300x300x93
2500 H300x300x87

2000

1500

1000

500

0
0 2 4 6 8 10 12 14

Effective Length L (m )
Design Chart for Strut Capacity

Figure 5-10. Design chart for strut size H300x300 in accordance with BS5950.

18000
H400x400x605
Strut Capacity (T/Strut), ULS BS 5950

16000

14000
H400x400x415
12000

10000
H400x400x283
8000 H400x400x235
H400x400x232
6000 H400x400x200
H400x400x197
H400x400x172
H400x400x147
4000 H400x400x140

2000

0
0 2 4 6 8 10 12 14

Effective Length L (m )
Design Chart for Strut Capacity

Figure 5-11. Design chart for strut size H400x400 in accordance with BS5950.

33
6 GROUND MOVEMENT

6.1 Prediction of ground movements

The prediction of ground settlements due to deep excavations should take into account
the effects of wall installation and extraction where applicable, wall deflection and
consolidation settlement. Other causes, such as ground treatment, piling should also be
considered, where appropriate.

Ground settlements due to diaphragm wall installation have been studied by Hulme et
al (1989) and Wen et al (2001). It is found that typically the settlement due to
diaphragm wall installation is in the order of 5 to 25mm.

The surface settlements due to wall deflection have been addressed by Clough &
O’Rourke (1990). They proposed that ground surface settlements due to deep
excavations take the form of the wall deflection as shown in Figure 6-1.

Figure 6-1. Surface settlement due to deep excavation.

This relationship has been confirmed by settlement measurements during the


construction of the cut & cover tunnels at Race Course Road for the North East MRT
Line, Wen et al (2001). Others have also reported similar correlation for deep
excavations in the Singapore marine clay, Hulme et al (1991), Nicholson (1987) and
Tan et al (1985). The ground surface settlements due to deep excavations in the Old
Alluvium and in the residual soils of the Bukit Timah Granite Formation have been
studied by Wong, et al (2001) and Lee et al (2001).

The prediction of movements during excavation shall be properly related to the results
of the excavation analyses. Finite element or finite difference methods can give
misleading results for the development of ground movements outside the excavation
unless the change of ground stiffness with strain is properly taken into account. All

34
analytical predictions should be checked against empirical methods based on
excavations in similar ground in Singapore.

6.2 Factors affecting ground movements

The following factors will affect the ground movement:

• Stiffness of retaining walls


• Factor of safety against base heave
• Factor of safety against failure through openings in retaining wall
• Loss of fines due to seepage through gaps in retaining walls or through timber
lagging
• Workmanship

Figure 6-2 shows the relationship of wall stiffness, factor of safety and settlement
behind the wall.

Sheetpile Wall Diaphragm Wall

Figure 6-2. Wall movement vs. wall stiffness / factor of safety against basal heave

6.3 Consolidation settlements

Typically consolidation settlement is caused by under-drainage of the soft clays such


as the marine clay and / or peaty clay. There are two sources that will cause a
drawdown of piezometric pressure in the soil layers below the soft clays:

• Seepage
• Stress reduction from ko to ka outside the excavation (active side)

35
• Stress increase from ko toward kp inside the excavation (passive side)

BS8002:1994 gives a simplified method to establish the pore water pressure at steady
state seepage condition, see Figure 6-3.

j
h

Note: Assume head difference


(h+i-j) is dissipated uniformly

Path Length
i along flow path length (2d+h-i-j)
d
2 (h + d - j) (d - i) γ w
PWT =
(2d + h - i - j)
PWT PWT
Figure 6-3. Pore water pressure at steady state seepage condition, BS 8002: 1994.

The mechanism of piezometric head draw-down has been studied by Shirlaw & Wen
(1999) and Wen & Lin (2001).

It has been shown by Hulme et al (1991) and Wen et al (2000) that consolidation
settlement can be up to 100% to 150% of the total measured settlements.

36
7 DAMAGE ASSESSMENT

7.1 Classification of Damage

The classification system for building damage adopted in the Design Criteria is based
on the work by Burland et al (1977) as given in Table 7-1. The classification in the
table relates only to visible or aesthetic damage, and the emphasis is laid on ease of
repair. Appropriate crack widths are listed and are intended merely as an additional
indicator rather than a direct measure of the degree of damage. The widths are based
on the views of engineers who have had the experience in observation of building
performance and the reaction of occupants.

Table 7-1 Classification of visible damage to walls with particular reference to ease of repair
Category Normal
of Degree of Description of typical damage
Damage Severity
0 Negligible Hairline cracks less than about 0.1mm.
1 Very Slight Fine cracks that are easily treated during normal decoration. Damage
generally restricted to internal wall finishes. Close inspection may
reveal some cracks in external brickwork or masonry. Typical crack
widths up to 1mm.
2 Slight Cracks easily filled. Re-decoration probably required. Recurrent
cracks can be masked by suitable linings. Cracks may be visible
externally and some re-pointing may be required to ensure weather-
tightness. Doors and windows may stick slightly. Typical crack
widths up to 5mm.
3 Moderate The cracks require some opening up and can be patched by a mason.
Re-pointing of external brickwork and possibly a small amount of
brickwork to be replaced. Doors and windows sticking. Service
pipes may fracture. Weather-tightness often impaired. Typical crack
widths are 5 to 15mm or several greater than 3mm.
4 Severe Extensive repair work involving breaking-out and replacing sections
of walls, especially over doors and windows. Windows and door
frames distorted, floor sloping noticeably1. Walls leaning1 or building
noticeably, some loss of bearing in beams. Service pipes disrupted.
Typical crack widths are 15 to 25mm, but also depends on the
number of cracks.
5 Very Severe This requires a major repair job involving partial or complete
rebuilding. Beams lose bearing, walls lean badly and require shoring.
Windows broken with distortion. Danger of instability. Typical crack
widths are greater than 25mm but depends on the number of cracks.

Burland and Wroth (1974) developed the concept of ‘critical tensile strain’ as a
parameter determining the onset of cracking; this was replaced by the concept of
‘limiting tensile strain’, εlim, by Burland et al (1977). Boscardin and Cording (1989)
analysed case histories of excavation induced subsidence, and showed that the damage
categories in Table 7-1 were related to the magnitude of tensile strain induced in the

37
building, and the range of strains were identified. The range of strains is adopted by
the Design Criteria, and is re-produced in Table 7-2.

Table 7-2. Relationship between category of damage and limiting tensile strain
Category Normal Degree Limiting tensile strain (∈lim)
of Damage of Severity (%)
0 Negligible 0 - 0.05
1 Very Slight 0.05 - 0.075
2 Slight 0.075 - 0.15
3 Moderate 0.15 - 0.3
4 to 5 Severe to Very Severe > 0.3

4.2 Methodology for Building Damage Assessment

The Design Criteria require all buildings and structures within the assessment zones
shall be assessed for damage by means of assessing the limiting tensile strains within
the buildings and the foundations. The assessment zones are defined in Chapter 20:
Assessment of Damage to Buildings and Utilities of the Design Criteria. The
methodology outlined by Mair et al (1996) can be adopted. The method of calculating
the limiting tensile strain is based on Burland & Wroth (1974). The method treats a
building as an idealised beam with span L and height H deforming under a maximum
deflection ∆. Expressions can be derived relating the ratio of ∆/L for the beam to
maximum bending strain (εb) and diagonal strain (εd) as follows:

∆  L 3I E 
= +  εb
L 12 t 2 t L H G 

and

∆  H L 2G 
= 1 +  εd
L  18 I E 

where:
H is the height of the building
L is the length of the building
E and G are respectively the Young’s modulus and shear modulus of the building
(assumed to be acting as a beam), E/G = 2.6 for masonary structures and E/G = 12.6
for framed buildings.
I is the second moment of area of the equivalent beam (i.e. H3/3)
t is the furthest distance from the neutral axis to the edge of the beam (=H)

The horizontal ground strains due to horizontal ground movement also contribute to
potential building damage. The average horizontal strain across a section of a building
is more appropriate in the context of potential damage than local horizontal ground
strains. The average horizontal strain, εh, that is transferred to the building can be

38
calculated as the difference of horizontal movement at both ends of a building divided
by the span length then gives an average horizontal strain.

The average horizontal strain is added to either the bending or diagonal strain to obtain
the maximum combined tensile strain, which is then compared with the limiting values
in Table 7-2 to assess the potential building damage.

The horizontal strain can be added directly to the bending strain giving

εbt = εh + εb

where εbt is the total bending strain.

Diagonal strains and horizontal strains can be combined by making use of a Mohr’s
circle of strain. Assuming a value of Poisson’s ratio of 0.3, the total tensile strain due
to diagonal distortion,εdt, is given by

εdt = 0.35εh + [(0.65εh)2 +εd2 ]

The Design Criteria require that protective works be provided to all buildings where
the predicted degree of severity of damage is moderate or above, with the aim to
restrict damage to the “slight” category or below.

4.3 Steps for Building Damage Assessment

The building damage assessment methodology is described by Mair, Taylor & Burland
(1996). A three staged building damage assessment can be adopted:

• Stage 1: If the predicted settlement is less than 10mm, no further assessment is


required.
• Stage 2: The building is assessed based on “green field” assumption, i.e. the
building is assumed to be flexible and move with the ground.
• Stage 3: Detailed evaluation is carried out, taking into account the building
stiffness and foundation.

In the detailed evaluation in Stage 3 or in a structural appraisal of a building, a holistic


approach should be adopted. Where applicable, the negative skin friction on the piles
should be evaluated in the assessment. The framing of the superstructures, layout of
the foundations and tie beams, reinforcement details in the structural members and
connection details should be considered as a whole.

Particular attentions should be paid to buildings on mixed foundations. These


buildings can be particularly sensitive to settlements. A detailed assessment needs to
be carried out for all buildings on mixed foundations.

A detailed assessment should also be carried out where a building is founded on piles,
and where the tip of the piles falls within a zone as defined in Figure 7-1.

39
Ground Level

Zone A
o o
60 60
Tunnel axis

Tunnel
Figure 7-1. Damage assessment for buildings on piles above the crown of a tunnel.

For utilities, allowable limits for settlements and deformations should be established
with the relevant utility agencies. Particular attention should be given to junctions of
pipe lines or joints of cables. The methodology outlined by Bracegirdle et al (1996)
can be followed for assessment.

40
8 INSTRUMENTATION AND MONITORING

8.1 General

The minimum monitoring required is covered in Chapter 19 of the Design Criteria.


Below are some guidelines to avoid common pitfalls in monitoring strut loads:

• Strain gauges should be mounted at the position not influenced by end effects
and also away from any joints or welds in the prop.
• The connections between strut and waling generally will result in non-uniform
stresses generated in the struts close to the ends. The strain gauges should be
installed at the middle of the strut.
• For case where bending stresses about the horizontal axis are significant they
should be measured by additional strain gauge at the top and bottom of the
strut (i.e. 4 gauges in total). The axial load can be found by averaging all four
gauges.
• The strain gauge temperature should be read and taken together with the strain
gauge reading. This information is important to understand the effects of
temperature changes on the strut load.
• The strain gauge readings should be taken immediately after the strut is
installed, when it carries no load and preferably with the strut weight being
supported along the length (i.e. no self-weight bending). This reading will
become the base reading and is essential to repeat the reading for 2 or 3 times.
• Readings should be taken after the strut is unloaded on removal as this can
check on the initial base reading and hence establish the actual measured strut
load.
• Strain gauges are to be protected by suitable housing as they are very
susceptible to damage.
• Excessive strut loads have been monitored in East-West Line Outram Park
station, Hwang et al (1987) and North East Line Dhoby Ghaut station
construction. Both stations were constructed in the Jurong Formation. The
struts load measured in the weathered rocks, which were sufficiently stable to
allow excavation using soldier piles without loss of ground, can be
significantly higher than those measured in the soft squeezing marine clay,
Shirlaw, et al (2000).

8.2 Review Levels

The Design Criteria, Chapter 19 specifies three review levels: Trigger, Design and
Allowable. This will be revised to Alert (=Trigger) and Work Suspension (= Design).

The monitoring results should be reviewed and interpreted. The reviewers should not
just look at the absolute readings of the instrument readings. Attentions should be paid
to the rate of movement and the stages of construction. An example of this is
demonstrated in Figure 8-1 and Figure 8-2.

41
30 60 100 δH

0.7 H
H

Final
H

Design
Level

Figure 8-1. Interpretation of wall movement, Scenario 1 (after Wong, 2003).

70 100 δH

0.3H 0.3 H
H
Final
H

Design
Level
Figure 8-2. Interpretation of wall movement, Scenario 2 (after Wong, 2003).

In scenario 1 (Figure 8-1), even though the excavation at 0.7H has not reached the
design level, yet, but based on the rate of the increase of deflection, the wall movement
would exceed the design level when the excavation reached the formation.

In Scenario 2 (Figure 8-2), even though at 30% of the excavation the movement has
reached 70% of the design level, the rate indicates that at the end of the excavation, the
wall deflection would still be within the design level.

The above example shows that it is important to examine the rate of movements when
interpreting the instrumentation results. This would ensure that any adverse trends and
events are identified in time. For critical structures a daily summary of the monitoring
results is particularly useful. An example of a daily summary for the monitoring of the
Foo Chow Church used during NEL C706 construction is shown in Figure 8-3.

Actions should be taken when pre-set review levels are reached. These actions can be
a review of the original design assumptions and re-design or additional ground
treatment measures. In the review of the original design assumptions and re-design of
the temporary support system, back analysis should be carried out, based on the
monitoring results, to verify the design parameters and design assumptions. This
would improve the accuracy of the predictions for the subsequent excavations, and
then revised review levels can be set.

42
Figure 8-3. Typical daily summary of monitoring results for critical structures.

43
9 PERFORMANCE OF TEMPORARY WORKS

9.1 General

The construction stage involves implementation of design and construction proposals


that are manifested in the form of shop drawings, design calculations and method
statement of works.

In the design of deep excavations the designer starts by analysing the problem with
appropriate methods and soil parameters to obtain a best estimate of the forces in the
retaining wall and strut loading. In the design certain assumptions will be made about
the ground condition, retaining system, strut arrangement and construction sequences.
During construction it needs to be ensured that these assumptions made in the design
are verified. The designers should be informed if the ground conditions and
construction deviate substantially from the design.

9.2 Ground Conditions

Ground is not a man-made material with well-defined characteristics and strength. The
distribution of different stratification, its thickness and the type of deposits depend on
the geological history. In addition, the ground conditions could be modified by
activities such as filling, cutting or excavation activities. Ground conditions can
therefore be very complex. Inevitably some assumptions and generalisation of the soil
profiles will have to be made during design stage. This generalisation of the soil
profiles for design purposes needs to be verified during construction.

Substantial soil information will be obtained after piling or after diaphragm wall
constructions. If the depth or thickness of the soft clay layer is found to be dipping
across the section or drastically different from the design assumption, the designer of
the retaining system should be informed. In such situations the designer should check
whether a re-design needs to be carried out with the additional soil information, in
particular whether the designed toe-in level is adequate if the dipping soft layer may
have resulted in a reduction in toe-in depth, see Figure 9-1. If the design is based on a
embedment depth of D2, then D1 should be checked again to ensure adequate
embedment of the retaining walls. This is important as inadequate toe-in will result in
over-stressing of the struts, excessive bending moment in the retaining wall and kick-
out of the retaining wall, causing excessive ground movement or the retaining system
to collapse.

9.3 Ground movements due to jet grouting

It is very common to use jet grouting to treat soft ground before excavations. While jet
grouting can increase the strength and stiffness of soft soils, it can also cause large
ground displacements, including deflection of diaphragm walls, see Figure 9-2. At
Race Course Road, localized heave was reported to be in excess of 300mm, Maguire
& Wen (1999). During the early phases of the construction of MRT lines, the

44
maximum measured heave was 1600mm, Shirlaw, et al (2000). As such, it is
important to sequence piling works after the jet grouting. Where possible, grouting
should be carried out in a direction away from areas of sensitive structures.

Fill Fill

Soft
Soft Retaining Clay
Clay Wall

Dipping of soft
layer

D2
OA
D1
OA

D1, D2 = Embedment depth in hard soil.


Figure 9-1. Adequacy of embedment of walls in dipping soft layers in the ground.

Deflection (mm)
-40.00 -30.00 -20.00 -10.00 0.00 10.00 20.00
102.00
Elevation

97.00

92.00

87.00
Grouting zone
82.00

77.00

72.00

67.00

62.00

Figure 9-2. Lateral diaphragm wall movement due to jet grouting works at Race Course Road.

During the jet-grouting process large volumes of grout, air and water is injected into
the ground. Surplus material produced by the injection of these materials is expelled to
surface via the annulus between the drilling rods and the surrounding marine clay. If
this passage is restricted or blocked the pressure inside the cavity created by the high
pressure jetting will be built up and when the pressure exceeds the cavity expansion
pressure, both lateral and vertical ground movement will be initiated.

45
The hydrodynamic pressure created by the high pressure jetting exists only within a
zone of influence, which is about 300 times that nozzle diameter, Covil & Skinner
(1994). For a nozzle diameter of 2mm, the zone of influence is only 0.6m. Therefore
the jet pressure does not control the pressure within the cavity. It is believed that the
movement of the ground (both lateral and vertical) is not a direct result of the high
pressure jetting adopted in the jet grouting works. The ground movement during jet
grouting works is a result of the pressure built up to expel the sludge to the ground
surface, see Figure 9-3. When the annulus between the grouting tube and the ground is
blocked, pressure can be locked into the ground causing the ground to move both
horizontally and vertically.

Grouting tube

Sludge
flow
h

Water or
air jet P = γs*h + Pf
Grout γs – Density of sludge
jet Pf – Pressure required to
P expel the sludge to the
surface

Figure 9-3. Pressure required to expel the sludge to surface.

Based on the understanding of the cause of the ground movement during jet grouting,
measures that can be taken to reduce the pressures to expel the sludge to the surface,
thus to reduce the ground movement can be:

• Reducing the density of the sludge;


• Increasing the size of the annulus by drilling a bigger hole;
• Using casings to prevent blockage of the sludge flow.

Jet grouting can be done using single, double and triple tube methods. It should be
noted that the triple tube method generally causes least impact on the surrounding
ground as compared with other two methods. This is consistent with experiences with
the use of triple tube method in Singapore. Shirlaw et al (2003) showed that the
diameter of the casing should be 200mm or larger for it to be effective. Other measures
can be pre-grouting to condition the soft clays and the use of pressure relief holes,
which aim to ensure free flow of the sludge and to reduce the pressure locked into the
ground.

46
9.4 Ground movements due to vibration

The driving of piles, installation of casings will often generate vibration. Vibration can
cause settlements in loose sands. Buildings particularly old masonry structures are
susceptible to vibrations. Vibrations can be reduced or minimised by the use of silent
pilers, casing oscillators which does not cause significant vibration.

Chiselling should be used with extreme caution in the vicinity of sensitive structures as
they can cause collapse of ground, vibration and cracking in adjacent structures.

9.5 Surcharge

Typically temporary retaining walls are designed for a nominal surcharge as a


uniformly distributed load behind the retaining walls. This nominal load is only
adequate for normal construction equipment. For heavier machinery, such as cranes,
and silos, etc, the nominal value is not adequate and should be considered separately.
On drawings designated area should be indicated where design checks have been
carried out to allow for heavy equipment.

9.6 Soldier piles and lagging

Soldier pile with timber lagging can be used where the ground conditions are
favourable and could remain stable without support for a limit period of time while the
lagging is put in place. However, where water table draw down is a concern, this
method should not be used as the wall is not continuous.

For areas where pre-bored holes are used to install the soldier piles, attention should be
paid to how the boreholes are to be filled after the soldier piles have been placed into
the bore. Sandy infills can be washed out easily during clement weather.

How the lagging is installed behind the king posts plays an important part in
minimising ground movement. There could be additional ground movement due to the
voids between the lagging and excavated face. The size of the voids depends on the
workmanship. It can be minimised by packing with concrete or cement mortar.

9.7 Gaps in retaining system due to underground services

Where utilities are not diverted, gaps may exist in the retaining system. Extra care
should be exercised in excavation. Often jet grouting columns are proposed to close
the gaps. In this method the soil within the gap is designed to arch over the gap. It
should be noted that the stiffness of the wall at the gap will be reduced. This will result
in large ground movement behind the wall. Loads at the gap will be transferred to the
adjacent retaining walls which should be checked and strengthened if necessary. The
struts should be checked as well.

It should be highlighted that the JGP columns may not be effective because of the
untreated zone underneath the utilities, see Figure 9-4. The size of this untreated zone

47
depends on the depth and width of the utilities. Where the untreated zone can be
avoided, a single row of JGP would not be adequate as there often exist gaps in the
interlocking of the jet grout columns. A minimum of two rows should be used to
minimise the risk of gaps in the JGP interlocking. Other chemical grouting, such as
cement-silicate may be used to pre-treat the fluvial sands, see Wen et al (2003). It
should be highlighted that the selection of grout should consider the permeability of
the ground and the participle size distribution of the soil to be grouted. Grouting
should be carried out before excavation commences.

Ground level

Underground services
Retaining wall

Untreated zones below the


service

Jet grouted piles to close


the gap in diaphragm wall

Final excavation level

Figure 9-4. Untreated zones due to obstruction of services.

9.8 De-clutching and welding of sheetpile wall

Sheet pile retaining wall should be checked during installation and during excavation
for signs of de-clutching. The sheetpile section modulus can only be achieved when
the individual sheetpiles are interlocked. Any de-clutching will reduce the section
modulus substantially. There should be proper guides and control of verticality so that
the wall does not deviate significantly out of position or de-clutched. If so, the gap
should be sealed by installing a secondary backing row of sheet piles or by stabilising
the gap by grouting.

Often sheetpiles have to be welded the piles are to be installed deeper than 12m. It is a
good engineering practice to stagger the welding connections between adjacent piles to
avoid forming a weak section in the retaining system.

9.9 Misalignment of secant pile walls

Another similar loss of interlocking between the individual elements is the


misalignment in a secant pile retaining wall. Due to excessive vertical deviation of a
pile, a large gap may exist in a secant pile wall at depth. Such gap may result in loss of
ground when exposed during excavation. Sigl et al (2003) reported such a case. When

48
the excavation reached 24m below ground, an unexpected ingress of ground water and
soil into the excavation took place as a results of gaps between the piles due to poor
verticality. The excavation had to be backfilled for 5m to implement a grouted curtain
outside the retaining wall. The cause of the failure was the misalignment of the secant
piles, leaving large gap for soil and water to flow into the excavation. The mechanism
is due to the hydraulic uplift as described in Section 4.2.

9.10 Pre-loading of struts

Pre-loading is a very effective way of reducing ground movement associated with


excavation. LTA Materials and Workmanship Specification requires a minimum pre-
loading of 50% of the design load on the strut at the particular stage of excavation.
Even a nominal amount of pre-load will go a long way in improving the system as the
slack in the load transfer system is taken up. Pre-load are usually applied onto
longitudinal and sometimes on diagonal struts. In applying the pre-load, it is important
to check that the forces in the support system are well balanced. Therefore preloading
should be done systematically and locked off. After excavation, if the pre-load is
found to drop, the loads can be re-applied. However caution should be exercised, when
the load trend indicates increasing loads for some and reduction for others. This may
signify an unbalanced strutting system that is not very desirable and could be
potentially unstable.

9.11 Excavation

Where the excavation is large and there is sensitive structures behind the retaining
wall, excavation work could be done in compartments or area by area to minimize the
exposure time and to install the support as soon as possible to restrict and minimize the
ground movement.

Over-excavation can occur in a number of ways. The commonest way is excavating


exceeding the designed depth prior to the installation of support or strutting. Others
include local excavation of pits for lift pit, sump, pile cap and etc. The design should
cater for over-excavation and the depth should be the maximum depth including the
excavation of sumps and pile caps.

If soil berms are used, it is to be noted that the berm size is very important in limiting
movement of the ground and deflection of the wall, Powrie & Daly, 2002. During
construction, the berm size should be monitored to avoid over-excavation and for
stability of the berm. To ensure full passive resistance, the berm must extend
sufficiently to develop the shear resistance and the slope must not fail. To prevent
deterioration of slope, the slope can be shotcreted.

The excavation particularly the last stage to formation level should be planned and
sequenced to minimise the ground settlement and movements. An effective way to
limit ground deformation in soft ground is to thicken the lean concrete slab at the base
of the excavation to act as temporary support at formation level.

49
Excavation in segments with strut installed within the excavated segment prior to the
next excavation is also useful in controlling the ground movement. Opening-up a long
stretch without installation of struts should be avoided. The length of each segment
will depend on the situations and practicability. 3 to 6m long are commonly adopted.

9.12 Strut removal

Backfill and compaction shall be carried out stage by stage in line with the removal of
struts.

When the construction involve a temporary wall, and if the permanent wall is not cast
against the temporary wall, a gap will exist between the permanent wall and the
temporary wall. In the limited space available it is difficult to achieve good
compaction during strut removal. Even when reasonable compaction has been
achieved there have been several instances on NEL projects where there has been
significant ground movement after the top strut is removed. To limit this movement,
one possible way is to entirely backfill the gap between the temporary wall and
permanent wall with lean mix concrete. Another way is to backfill the 1m below the
soffit of the struts with lean mix concrete. To limit movement, it is also a good practice
to backfill the gap between the temporary wall and the permanent wall over the full
height of the base slab, and also the 1m below the top of the roof slab. A typical
arrangement is shown in Figure 9-5.

9.13 Detailing of Connections

The connection details are important to ensure effective transmission of strut forces,
shear and moments.

Struts are basically designed as compressive members. To ensure no buckling, they


must be restrained laterally and vertically. At the interface area where there is high
point load, the walers must also be stiffened to prevent web buckling. Typical details
of bracing struts to the king-posts and runner beams are shown in Figure 9-6.
Stiffening of walers is shown in Figure 9-7. The dimensions indicated these figures are
indicative only. The use of sand bags or concrete as stifferers to walers should not be
accepted.

As far as possible, horizontal struts should be used instead of raked struts. It is difficult
to pre-load the raked struts and they will also subject the wall to vertical forces. Where
raked struts are used, the connection should be detailed properly for the transmission
of forces at the wall and where the load will be resisted. It is important that the strut at
the support is properly anchored as rotation of the strut may cause interfacing areas to
separate. There are different ways of providing the support, one of which is shown in
Figure 9-8.

50
Strut Level 1

Strut Level 2

Roof Slab
Strut Level 3
Lean concrete
Strut Level 4
Backfill soil
compacted to
Strut Level 5 specification

Base Slab

Temporary wall

Figure 9-5. Backfilling to avoid ground movement – typical arrangement

- ∅20 THREADED
MILD STEEL BAR
WITH M20 NUT &
SPRING WASHER
STRUT
600
450

C
L OF
INTERMEDIATE
PILE

- 450

600

PLAN

Figure 9-6 (a). Connection between struts and runner beams – Plan

51
L-100x100x10
STRUT

∅20 THREADED
MILD STEEL BAR
WITH M20 NUT & STRUT
SPRING WASHER

L-100x100x10

75 450 75

600

SECTION A-A
Figure 9-6 (b). Connection between struts and runner beams – Section

600
450

L-100x100x10

∅20 THREADED
MILD STEEL BAR
WITH M20 NUT &
SPRING WASHER STRUT

STRUT
L-100x100x10
SECTION B-B

Figure 9-6 (c). Connection between struts and runner beams – Section

52
Stiffener
Waler
Strut
D/Wall Waler Stiffener

G30 Concrete
Packing

G8-8 Bolts

6 70

Brackets
650
Anchor Strut
Bolt (b) Plan.
(a) Elevation.

Figure 9-7. Stiffeners for waler at strut / waler connection

Inclined “Kicker” or
“Spur brace”

Plate if necessary 45
Weld C
Prop Waling

Weld A

Stiffeners top Soldier


and bottom at beam
prop

Horizontal wood
Weld A = Weld B & C
sheeting (lagging)
= vertical component of
prop load

Figure 9-8. Typical details for raked struts.

53
It is considered a good engineering practice to use packing between retaining walls
and the walers to ensure an even distribution of loading from the retaining wall to the
walers, see Figure 9-9.

For diagonal struts at corners it is essential to ensure that the walers are supported at
the end, for example by end blocks to prevent slip of the walers against the retaining
walls, see Figure 9-10. Every strut must have a clear load path and point of resistance.
In the example shown in Figure 9-8, not only must the waler be supported by the end
block, but the end block itself must be connected to the wall and the wall must find
sufficient resistance from the soil. Each component in the load path must have
sufficient strength to transfer the load.

Figure 9-11 shows a typical end block details anchored to a diaphragm wall. In the
detail, the strut is installed directly against the end block.

Retaining wall
Soil G30 concrete fill

200 Typ. Shear


connection

Waler

Strut Strut

Figure 9-9. Concrete packing between waler and retaining wall.

Retaining
walls

Walers
Diagonal
struts

End block to prevent


slip of waler against
the retaining wall

Figure 9-10. Diagonal struts and walers at corners.

54
T20

630
45o
T13

Diagonal 20
strut 100
20
20
6
100
End plate

PLAN

Figure 9-11a. Typical details of end block anchored to diaphragm walls - Plan

141

15 278.5 278.5 278.5 278.5 15


15
600

30
15

SECTION A - A

Figure 9-11b. Typical details of end block anchored to diaphragm walls - Section

55
100

6 200 200 200


T20 6

6
15

48
60
30
15

6
T13

200 500

SECTION B - B
Figure 9-11c. Typical details of end block anchored to diaphragm walls - Section

56
REFERENCES

Bjerrum, L. & Eide, O. (1956). Stability of strutted excavations in clay, Geotechnique, Vol. 6, No. 1,
March 1956.
Boscardin, M.D. & Cording, E.J. (1989). Building response to excavation induced settlement, Journal
of Geotechnical Engineering, ASCE, Vol. 115, 1989.
British Standard BS8002. (1994). Code of practice for earth retaining structures. British Standard
Institution.
Burland, J.B., Broms, B.B, & De Mello, V.F.B. (1977). Behaviour of foundations and structures,
Proceedings of the 9th International Conference of Soil Mechanics and Foundation Engineering,
Tokyo, 1977.
Burland, J.B. and Wroth, C.P. (1974). Alloeable and differential settlement of structures, including
damage and soil-structure interaction, Proceedings of Conferenceon Settlement of Structures,
Cambridge University,Canbridge, UK, 1974.
Chu, J., Wen, D., Kay, R.E. & Tay, T.H. (2000). Engineering properties of fluvial sand at Race Course
Road, Proceedings of International Conference on Tunnels and Underground Structures,
Singapore, 26-29 November 2000.
Clough, G.W. & Schmidt, B. (1981). Design and performance of excavations and tunnels in soft clay.
Soft Clay Engineering, ed E.W. Brand and R.P.Brenner, Elsevier, Amsterdam.
Covil, C.S. & Skinner, A.E. (1994). Jet grouting, - a review of some of the operating parameters that
form the basis of the jet grouting process. Proceedings Grouting in the ground, A.L. Bell ed.,
Thomas Telford.
Dames and Moore (1983). Mass Rapid Transit System, Singapore: detailed geotechnical study –
interpretative report, Provisional Mass Rapid Transit Authority, Singapore.
Davies, R. V. (1984). Some geotechnical problems with foundations and basements in Singapore,
Proceedings of International Conference on Tall Buildings, Singapore, 22-26 October 1984.
Flanagan, R., Schmidt, B. & Tan, B.T. (1999). Major sewer tunnel construction beneath metro tunnels
in Singapore. Proceedings of the Rapid Excavation & Tunnelling Conference, Florida.
Fleming (1981). Piling engineering, Published by John Wiley & Sons, 1981.
Geotechnical Manual for Slopes, Geotechnical Control Office, Engineering Development Department,
Hong Kong, 1984.
Han, K.K., Wong, K.S. Broms, B.B. & Yap, L.P. (1993a). The origin and properties of bouldery clay
in Singapore, Geotechnical Engineering, 24.
Han, K.K., Wong, K.S. Broms, B.B. (1993b). Singapore bouldery clay: the origin properties and load
tests of bored and cast-in-place piles. Proceedings of 11th Southeast Asian Geotechnical
Conference, May, Singapore.
Hoek E. and Brown, E.T. (1997). Practical estimates of rock mass strength. International Journal of
Rock Mechanics and Mining Science, Vol. 34(8).
Hulme, T.W., Potter, L.A. & Shirlaw, J.N. (1989). Singapore Mass Rapid Transit System:
construction, Proceedings of Institution of Civil Engineers, Transportation Engineering Group,
Part 1, 1989.
Hulme, T. W., Chapman, P.H.J., Foo, P.H., Copsey, J.P., Kraft, B., Sripathy, P., Potter, L.A.C. &
Shirlaw, J.N. (1991). Discussion, Proceedings of the Institution of Civil Engineers, Transportation
Engineering Group, Vol. 90.
Hwang, R., Quah, H. P. & Buttling, S. (1987). Measurements of strut forces in braced excavations.
Proceedings of the Singapore Mass Rapid Transit Conference, 6-9 April 1987, Singapore.
KarWinn, Rahardjo, H. & Seh, C.P. (2001). Characterisation of Residual Soils in Singapore.
Geotechnical Engineering, 23(1).
Lee, F.H., Yong, K.Y., Quan, C.N. & Chee, K.T. (1998). Effect of corners in strutted excavations:
field Monitoring and case histories. Journal of Geotechnical and Geoenvironmental Engineering
ASCE, April 1998, Vol. 124, No. 4.
Lee, F.H., Tan, T.S. & Wang, X.N. (2001). Ground water effects in soldier-piled excavations in
residual soils. Proceedings of Underground Singapore 2001, Singapore, 29-30 November 2001.

57
Lim, T.T. (1995). Shear strength characteristics and rainfall-induced matric suction changes in a
residual soil slope, MEng Thesis, School of Civil & Structural Engineering, Nanyang Technological
University, Singapore.
Lo, K.W., Leung, C.F., Hayata, K. & Lee, S.L. (1988). Stability of excavated slopes in weathered
Jurong Formation of Singapore. Proceedings of 2nd Int. Conf. On Geomechanics in Tropical Soils,
Singapore, 1: 277-284. Balkema: Rotterdam.
Maguire, C. & Wen, D. (1999). Practical aspects of jet grouting in Singapore marine clay.
Proceedings of International Conference on Rail Transit, Singapore.
Mair, R., Taylor, R.N. & Burland, J.B. (1996). Prediction of ground movements and assessment of
risk of building damage due to bored tunnelling, Proceedings of International Symposium on
Geotechnical Aspects of Underground Construction in Soft Ground, London, 1996.
Orihara, K. and Khoo, K.S. (1998). Engineering properties of Old Alluvium in Singapore and its
parameters for bored pile and excavation design, Proceedings of 13th Asian Geotechnical
Conference, Taiwan.
Orihara, K., Makino, M. & Tse, T.K.D. (1999). Excavation for construction of new Dhoby Ghaut
MRT station. Proceedings of International Conference on Rail Transit, 11-13 March 1999,
Singapore.
NAVFAC DM7.2 (1982) Foundations and Earth Structures, Design Manual 7.2, Department of the
Navy, Naval Facilities Engineering Command, Alexandria, VA. 22332, USA.
Nicholson, D.P. (1987). The design and performance of the retaining walls at Newton station.
Proceedings of the Singapore Mass Rapid Transit Conference, Singapore, 6-9 April 1987.
Oaimi, A.E. & Clough, G.W. (1979). Pore pressure distribution during excavation, Journal of
Geotechnical Engineering Division, ASCE, Vol. 105.
Padfield, C.J. & Mair, R.J. (1984). Design of retaining walls embedded in stiff clay. CIRIA Report
104, Construction Industry Research and Information Association, London.
Pitts, J. (1984). A review of geology and engineering geology in Singapore. Quarterly Journal
Engineering Geology, Vol. 17. London.
Pitts, J. (1984). A survey of engineering geology in Singapore. Geotechnical Engineering, Vo. 15.
Pitts, J. (1985). Discussion on “ A review of geology and engineering geology in Singapore”.
Quarterly Journal Engineering Geology,Vol. 18,London.
Potts, D.M. & Fourie, A.B. (1984). The effect of wall stiffness on the behaviour of a propped retaining
wall. Geotechnique, Vol. 35.
Poh, K.B., Chua, H.L. & Tan, S.B. (1985). Residual granite soil of Singapore. Proceedings of 8th
Southeast Asian Geotechnical Conference, Kuala Lumpur.
Poh, T.Y & Wong, I.H. (1998). Effects of construction of diaphragm wall panels on adjacent ground:
field trial. Journal of Geotechnical and Geoenvironmental Engineering.
Powrie W. & Daly, M.P. (2002). Centrifuge model test on embedded retaining walls supported by
earth berms. Geotechnique, 52, No. 2.
Rahardjo, H. (2000). Rainfall-Induced Slope Failures. NSTB 17/6/16 Main Report, Nanyang
Technological University, Singapore.
Shirlaw, J.N., Poh, KB. Hwang, R.N. (1990). Properties & origins of Singapore boulder bed.
Proceedings of 10th Southeast Asian Geotechnical Conference, April, Taipei.
Shirlaw, J.N. & Wen, D. (1999). Measurement of pore pressure changes and settlements due to a deep
excavation in Singapore marine clay. Proceedings of the 5th International Symposium on Field
Measurements in Geomechanics, 1-3 December 1999, Singapore.
Shirlaw, J.N., Hencher, S.R. & Zhao, J. (2000). Design and construction issues for excavation and
tunnelling in some tropically weathered rocks and soils. Proceedings of GeoEng2000, 19-24
November, 2000. Melbourne, Australia.
Shirlaw, J.N., Wen, D., Nadarajah, P., Yoon, S. I. & Sugawara, S. (2000a). Construction issues related
to jet grouted slabs at the base of excavations. Proceedings of Tunnels and Underground Structures,
an International Conference, 26-29 November 2000, Singapore.
Shirlaw, N.J., Broome, P.B., Chandrasegaran, S., Daley, J., Orihara, K., Raju, G.V.R., Tang, S.K.,
Wong, I.H., Wong, K.S. & Yu, K. (2003). The Fort Canning Boulder Bed. Proceedings of
Underground Singapore 2003 and Workshop “Updating the Engineering Geology of Singapore”,
27-29 November 2003, Singapore.
Shirlaw, J.N., Wen, D., Kheng, H.Y. & Osborne, N. (2003). Controlling heave during jet grouting on
Marine Clay. Proceedings of the RTS Conference, Singapore.

58
Sigl, O., Jedlitschka, G. & Weheli, J.M. (2003). DTSS Shaft R2: shaft and tunnel excavation in
weathered Bukit Timah Granite. Proceedings of Underground Singapore 2003, 27-29 November
2003, Singapore.
Singapore Standard CP18 : 1992. Code of Practice for earthworks, Singapore Institute of Standards
and Industrial Research, 1992.
Tan, S.B., Loy, W.C. & Lee, K.W. (1980). Engineering geology of the Old Alluvium in Singapore. 6th
Southeast Asian Conference on Soil Engineering, Taipei.
Tan, S.B., Tan, S.L., Lim, T.L. & Yang, K.S. (1987). Landslide problems and their control in
Singapore. Proceedings of 9th Southeast Asian Geotechnical Conference, Bangkok, Thailand.
Wen, D. & Lin, K.Q. (2002). The effect of deep excavation on pore water pressure changes in the Old
Alluvium and under-drainage of marine clay in Singapore. Proceedings of the 3rd International
Symposium on Geotechnical Aspects of Underground Construction in Soft Ground, 23-25 October
2002, Toulouse, France.
Wen, D., Ow, C.N. & Yoon S.I. (2001). The monitoring of cut and cover tunnel construction at Race
Course Road next to Foochow Methodist Church, Proceedings of Underground Singapore 2001,
Singapore, 29-30 November 2001.
Wen, D., Heng, C.H., Choo, H.Y. & Yoon, S.I. (2003) Gaps in diaphragm walls and remedial
measures for the deep excavations at Race Course Road. Proceedings of Underground Singapore,
27-29 November 2003, Singapore.
Wong, I.H., Poh, T.Y. & Chuah, H.L. (1997). Performance of excavations for depressed expressways
in Singapore. Journal of Geotechnical Engineering and Geoenvironment Engineering, ASCE, Jult
1997.
Wong, K.S., Li, W., Shirlaw, J.N., Ong, J.C.W., Wen, D. & Hsu, J.C.W. (2001). Old Alluvium:
engineering properties and braced excavation performance. Proceedings of Underground
Singapore 2001, Singapore.
Wong, K.S. (2003). Observational approach to avoid failures in temporary works, Proceedings of BCA
Seminar: Avoiding Failures in Excavation Works, 11 July 2003.
Wong, K.S., Goh, A.T.C. & Ganeshan, V. (2003). Design of temporary works, an in-house short
course for Land Transport Authority, 11 & 15 December 2003, Land Transport Authority,
Singapore.
Yang, K.S. & Tang, S.K. (1997). Stabilising the slope of Bukit Gombak. Proceedings of 3YGEC,
Singapore.
Yong, R.N., Chen, C.K., Sellappah, J. & Chong, T.S. (1985). The characterization of residual soils in
Singapore. Proceedings of 8th Southeast Asian Geotechnical Conference, Kuala Lumpur,
Malaysia.
Zhou, Y. (2001). Engineering geology and rock mass properties of the Bukit Timah Granite.
Proceedings of Underground Singapore 2001, Singapore.
Zhu, H. (2000). Evaluation of load transfer behaviour of bored piles in residual soils incorporating
construction effects. Ph.D. Thesis, School of Civil & Structural Engineering, Nanyang
Technological University, Singapore.

59

Das könnte Ihnen auch gefallen