You are on page 1of 165

Citation: Ahmed S.A.

(2014) "State-of-the-Art Report: Deformations Associated with Deep Excavation and


Their Effects on Nearby Structures," Ain Shams University, Faculty of Engineering, Structural Engineering Dept.
DOI: 10.13140/RG.2.1.3966.9284.

Ain Shams University


Faculty of Engineering
Structural Engineering Department
Geotechnical Engineering Group

STATE-OF-THE-ART REPORT:
DEFORMATIONS ASSOCIATED WITH DEEP EXCAVATION AND
THEIR EFFECTS ON NEARBY STRUCTURES

Sayed Mohamed El-Sayed Ahmed


April, 2014

CONTENTS
1.

INTRODUCTION ................................................................................................................... 9

2.

FACTORS AFFECTING EXCAVATION DEFORMATIONS ........................................... 14


2.1.

Soil Type ........................................................................................................................ 15

2.2.

Wall Stiffness and Excavation Stability ......................................................................... 15

2.3.

Overconsolidation (OCR) and At-Rest Earth Pressure Coefficient (Ko) ....................... 22

2.4.

Groundwater Conditions and Control Measures ............................................................ 23

2.5.

Strut/Tie-back Prestressing ............................................................................................ 28

2.6.

Construction Sequence ................................................................................................... 30

2.7.

Wall Lateral Deformation Patterns ................................................................................ 32

2.7.1.

Settlement pattern associated with the wall cantilever deformation mode ............. 35

2.7.2.

Settlement pattern associated with the wall bulging deformation mode ................ 37

2.8.

Time-Dependent Effects ................................................................................................ 39

2.9.

Excavation Geometry and Three-Dimensional Effects .................................................. 39

2.9.1.

Excavation dimensions and depth to firm layers .................................................... 39

2.9.2.

Corner effect ........................................................................................................... 41

2.9.3.

Parallel distribution ................................................................................................. 42

2.10.

Wall Installation Effect ............................................................................................... 43

2.11.

Building Stiffness and Weight .................................................................................... 49

2.12.

Wall-Soil Interface ..................................................................................................... 51

2.13.

Workmanship.............................................................................................................. 52

3.

EMPIRICAL AND SEMI- EMPIRICAL METHODS ......................................................... 55

4.

NUMERICAL MODELING ................................................................................................. 73


4.1.

Beam-Column on Elastic Winkler Springs .................................................................... 73

4.2.

The Finite Elements ....................................................................................................... 81

4.2.1.

Constitutive soil modeling ...................................................................................... 83

4.2.2.

Simulation of excavation and construction sequence ............................................. 84

4.2.3.

Interface modeling. ................................................................................................. 86

4.2.4.

Two-dimensional versus three-dimensional analyses ............................................. 87

4.2.5.

Modeling of structures affected by excavations ..................................................... 87

5.

ANALYTICAL APPROACH ............................................................................................... 88

6.

ARTIFICIAL NEUTRAL NETWORKS (ANNS)................................................................ 90

7.

PHYSICAL MODELING USING CENTRIFUGE .............................................................. 93

8.

OBSERVATIONAL METHOD AND MONITORING ....................................................... 97


2

9.

INSTRUMENTATION AND MONITORING..................................................................... 99


9.1.

Deformation Instrumentations...................................................................................... 102

9.2.

Stress Measurements .................................................................................................... 105

9.2.1.

Piezometers ........................................................................................................... 105

9.2.2.

Strain Gauges ........................................................................................................ 107

9.3.

Real Time Monitoring .................................................................................................. 108

9.3.1.

Robotic total stations (RTS).................................................................................. 108

9.3.2.

Three-dimensional Laser scanning ....................................................................... 109

9.4.

Trigger Levels for Monitoring ..................................................................................... 110

9.5.

Monitoring Experience in Egyptian Deep Excavation Projects ................................... 111

10.

BUILDING DAMAGE CRITERIA ................................................................................ 114

10.1.

Superstructure Damage criteria ................................................................................ 114

10.1.1.

The maximum angular distortion () criterion .................................................. 117

10.1.2.

The maximum deflection ratio (/L) criterion .................................................. 118

10.1.3.

The limiting tensile strain criterion ................................................................... 118

10.1.4.

The crack width criterion .................................................................................. 126

10.1.5.

The maximum settlement (Smax) maximum rotation (max) criterion ............. 128

10.1.6.

The Damage Potential Index (DPI( criterion .................................................... 128

10.2.

Assessment of the Induced Building Damage .......................................................... 129

10.2.1.

Primary assessment ........................................................................................... 129

10.2.2.

Second stage assessment ................................................................................... 130

10.2.3.

Detailed assessment........................................................................................... 130

10.3.

Features Affecting Structural Damage ..................................................................... 131

10.3.1.

Ratio of the buildings Young modulus to its shear modulus (E/G) ................. 131

10.3.2.

Grade beams ...................................................................................................... 133

10.3.3.

Building-Soil Relative Stiffness ........................................................................ 133

10.4.

Damage Assessment for Deep Foundations ............................................................. 136

10.5.

Design Deep Excavations for Admissible Structural Deformations ........................ 140

11.

SURVEY OF DAMAGE DUE TO INDUCED GROUND MOVEMENT .................... 141

12.

RISK MANAGEMENT AND MITIGATIONS .............................................................. 143

13.

SUMMARY OF THE PRESENTED WORKS ............................................................... 148

14.

REFERENCES ................................................................................................................ 152

List of Figures
FIGURE 1. GROUND AND BUILDING DEFORMATIONS INDUCED BY A DEEP EXCAVATION (HSIAO, 2007) ................................................9
FIGURE 2. FAILURE OF A BUILDING IN CHINA IN 2009 THAT WAS INITIATED BY A NEARBY DEEP EXCAVATION .......................................10
FIGURE 3. FAILURE OF A DEEP EXCAVATION ADJACENT TO NICOLL HIGHWAY, SINGAPORE (LEE, 2008) .............................................10
FIGURE 4. A MASONRY WALL SUFFERED FROM SEVERE CRACKING DUE TO GROUND DEFORMATIONS (VATOVEC ET AL., 2010) ...............11
FIGURE 5. TYPICAL FORMATIONS IN THE GREATER CAIRO AREA (EL-SOHBY AND MAZEN, 1985) ......................................................12
FIGURE 6. EFFECT OF THE SOIL TYPE ON THE SETTLEMENTS INDUCED BY DEEP EXCAVATION (PECK, 1969A) ........................................15
FIGURE 7. EFFECT OF WALL STIFFNESS AND SOIL STABILITY NUMBER ON THE WALL DEFORMATIONS IN CLAYS (GOLDBERG ET AL., 1976) ..16
FIGURE 8. EFFECT OF THE BASAL HEAVE STABILITY ON THE WALL DEFORMATIONS INDUCED BY DEEP EXCAVATIONS IN CLAYS (MANA &
CLOUGH, 1981) .....................................................................................................................................................16
FIGURE 9. EFFECT OF THE BASAL HEAVE STABILITY AND THE SYSTEM STIFFNESS ON THE WALL DEFORMATIONS INDUCED BY DEEP
EXCAVATIONS IN CLAYS (CLOUGH AT AL., 1989) ........................................................................................................... 17
FIGURE 10. NORMALIZED FIELD MEASUREMENTS OF THE LATERAL DEFORMATIONS AGAINST CLOUGH & OROURKES (1990) SYSTEM
STIFFNESS AND BASAL HEAVE FACTOR OF SAFETY FOR CASES WITH LOW FACTOR OF SAFETY (FOS<3) (LONG, 2001) ................... 18
FIGURE 11. NORMALIZED FIELD MEASUREMENTS OF THE LATERAL DEFORMATIONS AGAINST CLOUGH & OROURKES (1990) SYSTEM
STIFFNESS AND BASAL HEAVE FACTOR OF SAFETY FOR CASES WITH HIGH FACTOR OF SAFETY (FOS>3) (LONG, 2001) .................. 19
FIGURE 12. NORMALIZED FIELD MEASUREMENTS OF THE LATERAL DEFORMATIONS AGAINST CLOUGH & OROURKES (1990) SYSTEM
STIFFNESS AND BASAL HEAVE FACTOR OF SAFETY FOR SOFT CLAY (MOORMANN, 2004)......................................................... 20
FIGURE 13. NORMALIZED FIELD MEASUREMENTS OF THE LATERAL DEFORMATIONS AGAINST CLOUGH & OROURKES (1990) SYSTEM
STIFFNESS AND BASAL HEAVE FACTOR OF SAFETY FOR STIFF CLAY (MOORMANN, 2004) ........................................................ 20
FIGURE 14. NORMALIZED LATERAL WALL MOVEMENTS VS. RELATIVE STIFFNESS RATIO, R, FOR DEEP EXCAVATIONS IN COHESIVE SOILS
(ZAPATA-MEDINA, 2007). .......................................................................................................................................21
FIGURE 15. EFFECT OF THE FACTOR OF SAFETY ON THE SETTLEMENT TROUGH WIDTH (MANA AND CLOUGH, 1981) ............................22
FIGURE 16. CONTOURS OF STRESS LEVEL AT AN EXCAVATION DEPTH OF 13.26M: (A) KO=2; (B) KO=0.5 (POTTS & FOURIE, 1984) .......22
FIGURE 17. GROUNDWATER FLOW PATTERNS ENCOUNTERED IN DEEP EXCAVATIONS (CLOUGH & OROURKE, 1990) ..........................23
FIGURE 18. INFLUENCE OF THE DEWATERING WORKS ON THE GROUND SETTLEMENT.....................................................24
FIGURE 19. GROUNDWATER CONTROL MEASURES FOR BRACED EXCAVATION (PULLER, 2003) .........................................................24
FIGURE 20. COLLAPSE OF CITY ARCHIVE BUILDING IN COLOGNE (GERMANY) DUE SOIL PIPING INDUCED BY DEWATERING (ROWSON, 2009)
...........................................................................................................................................................................25
FIGURE 21. THE COLLAPSED CITY ARCHIVE BUILDING IN COLOGNE (GERMANY) (ROWSON, 2009) ..................................................26
FIGURE 22. DAMAGE DUE TO SUBSIDENCE ALONG AN UNDERGROUND STATION OF THE NORTH-SOUTH TRAIN LINE IN AMSTERDAM (VAN
BAARS, 2011). ......................................................................................................................................................26
FIGURE 23. LEAKAGE AND DAMAGE AT THE BUILDING PIT IN MIDDELBURG, THE NETHERLAND (VAN BAARS, 2011) ............................27
FIGURE 24. FAILURE OF A DIAPHRAGM WALL IN THE INFINITY TOWER IN DUBAI IN 2007. THE CHRONOLOGICAL SEQUENCE OF EVENTS IS
(A) TO (D) ..............................................................................................................................................................27
FIGURE 25. SCHEMES FOR GROUNDWATER CONTROL IN A DEEP EXCAVATION (EL-NAHHAS, 2006). .................................................28
FIGURE 26. THE EFFECT OF PRESTRESSING ON THE WALL DEFORMATIONS (CLOUGH, 1975) ............................................................29
FIGURE 27. THE EFFECT OF THE STRUT STIFFNESS ON THE MAXIMUM LATERAL DEFORMATION OF THE WALL AND THE MAXIMUM
SETTLEMENT (MANNA & CLOUGH, 1981) ................................................................................................................... 29
FIGURE 28. CONSTRUCTION PROCEDURE STEPS FOR THE GREATER CAIRO METRO - LINE 1 (EL-NAHHAS ET AL., 1988) ........................30
FIGURE 29. ROD EL-FARAG STATION (AHMED AND ABD EL-SALAM, 1996) ................................................................................31
FIGURE 30. SETTLEMENT PATTERNS ASSOCIATED WITH DIFFERENT WALL DEFORMATION MODES (GOLDBERG ET AL. 1976). ..................32
FIGURE 31. MODES OF DEFORMATION OF THE WALL (CLOUGH AND OROURKE, 1990).................................................................33
FIGURE 32. LATERAL AND VERTICAL DISPLACEMENT PATTERNS: CONCAVE ON LEFT, SPANDREL ON RIGHT (BOONE 2003; BOONE &
WESTLAND, 2005). ................................................................................................................................................33
FIGURE 33. THE RATIO BETWEEN THE MAXIMUM HORIZONTAL TO VERTICAL DISPLACEMENT AS A FUNCTION OF THE COEFFICIENT OF
DEFORMATIONS (OROURKE, 1981) ..........................................................................................................................34
FIGURE 34. THE RATIO BETWEEN THE MAXIMUM HORIZONTAL TO VERTICAL DISPLACEMENT (MANA & CLOUGH 1981) .......................34
FIGURE 35. DEFORMATIONS PREDICTION FROM LATERAL WALL DEFLECTION VALUES PROPOSED BY AYE (2006): (A) SETTLEMENTS; (B)
LATERAL DEFORMATIONS .......................................................................................................................................... 35
FIGURE 36. SPANDREL-TYPE SETTLEMENT TROUGH (OU ET AL., 1993) .......................................................................................36

FIGURE 37. ASSUMED GAUSSIAN DISTRIBUTION FOR LATERAL AND VERTICAL GROUND DEFORMATIONS (LEE ET AL., 2007) ...................36
FIGURE 38. CONCAVE SETTLEMENT PROFILE (HSIEH & OU, 1998).............................................................................................37
FIGURE 39. RELATIONSHIP BETWEEN WALL MOVEMENT AND GROUND SETTLEMENTS FOR SOFT/LOOSE SOILS (KARLSRUD, 1997). ..........37
FIGURE 40. VERTICAL AND HORIZONTAL GROUND MOVEMENT PATTERNS AS A FUNCTION OF THE EXCAVATION DEPTH (HE) AND THE
DISTANCE FROM THE WALL (D) (SCHUSTER ET AL. 2009) ................................................................................................ 38
FIGURE 41. SUBSURFACE SETTLEMENT DISTRIBUTION FOR CONCAVE SETTLEMENT PROFILES (AYE ET AL. 2006)...................................39
FIGURE 42. EFFECT OF THE EXCAVATION WIDTH ON THE MAXIMUM GROUND SETTLEMENT AND THE WALL DEFLECTION (MANA & CLOUGH,
1981) ..................................................................................................................................................................40
FIGURE 43. EFFECT OF THE DEPTH TO FIRM LAYER ON THE MAXIMUM GROUND SETTLEMENT AND THE WALL DEFLECTION (MANA &
CLOUGH, 1981) .....................................................................................................................................................40
FIGURE 44. THE EFFECT OF THE HARD STRATUM ON THE COMPUTED WALL DEFLECTION (HSIAO, 2007) ..........................................41
FIGURE 45. PLANE STRAIN RATIO (PSR) AS A FUNCTION OF THE ASPECT RATIO (B/L) AND DISTANCE FROM THE CORNER (D)
(OU ET
AL., 1996) ............................................................................................................................................................ 41
FIGURE 46. THREE-DIMENSIONAL DISTRIBUTION OF SETTLEMENT AND LATERAL MOVEMENT AROUND FINITE DEEP EXCAVATION (FINNO &
ROBOSKI, 2005; ROBOSKI &FINNO, 2006) ................................................................................................................42
FIGURE 47. SETTLEMENT ASSOCIATED WITH TRENCHING IN HONG KONGS MTR (MORTON ET AL., 1980) .......................................43
FIGURE 48. MAXIMUM BUILDING SETTLEMENTS DUE TO SLURRY TRENCH EXCAVATION FOR DIAPHRAGM WALLS AS A FUNCTION OF
FOUNDATION DEPTH IN HONG KONGS MTR (COWLAND & THORLEY, 1984) .................................................................... 44
FIGURE 49. BUILDING SETTLEMENT DUE TO DIAPHRAGM WALL INSTALLATION IN HONG KONGS MTR (BUDGE-REID ET AL., 1984) ......44
FIGURE 50. SETTLEMENT DUE TO INSTALLATION OF A DIAPHRAGM WALL (CLOUGH AND OROURKE, 1990) .......................................45
FIGURE 51. LATERAL DEFORMATION ASSOCIATED WITH TRENCHING FOR SECANT PILES INSTALLED IN CHICAGO CLAY
(FINNO ET AL.,
2002) ..................................................................................................................................................................45
FIGURE 52. THE EFFECT OF SLURRY LEVEL VARIATION AND ITS HOLDING TIME ON THE LATERAL DEFORMATIONS ASSOCIATED WITH
TRENCHING (POH AND WONG, 1998) ........................................................................................................................ 46
FIGURE 53. VERTICAL DEFORMATIONS DUE TO DIAPHRAGM WALL INSTALLATION (GABA ET AL. 2003) ..............................................47
FIGURE 54. INFLUENCE OF PANEL LENGTH ON LATERAL DISPLACEMENTS (GOURVENEC & POWRIE, 1999)........................................47
FIGURE 55. THE SETTLEMENT ENVELOPES FOR SHALLOW AND DEEP FOUNDATION DUE TO TRENCHING TO A DEPTH OF 21M IN THE NILE
ALLUVIUMS IN THE GREATER CAIRO (ABDEL-RAHMAN & EL-SAYED, 2009). ......................................................................48
FIGURE 56. THE RELATIONSHIP BETWEEN THE LATERAL DEFORMATIONS AND THE MAXIMUM SETTLEMENT DUE TO TRENCHING IN THE NILE
ALLUVIUMS IN THE GREATER CAIRO (EL-SAYED & ABDEL-RAHMAN, 2002). ......................................................................49
FIGURE 57. BUILDINGS AND INTERFACES USED IN CENTRIFUGE TESTS (ELSHAFIE, 2008) .................................................................50
FIGURE 58. WALL DEFLECTIONS VARIATION WITH THE VARIATION OF THE WALL-SAND FRICTION (YU & GANG, 2008) .........................51
FIGURE 59. MAXIMUM SETTLEMENT VERSUS THE WALL-SAND INTERFACE FRICTION ANGLE (YU & GANG, 2008) ................................51
FIGURE 60. THE VARIATION OF THE RATIO BETWEEN THE MAXIMUM SETTLEMENT TO THE MAXIMUM WALL DEFLECTION VERSUS THE WALLSAND INTERFACE FRICTION ANGLE (YU & GANG, 2008) ................................................................................................. 52
FIGURE 61. DIFFERENT METHODS OF PLACING LAGGING (PECK, 1969A) .....................................................................................53
FIGURE 62. SUBGRADE REACTION MODEL FOR ANALYSIS OF WALLS SUPPORTING DEEP EXCAVATIONS (DELATTRE, 2001). .....................74
FIGURE 63. HORIZONTAL SUBGRADE MODULI, KH (AFTER PFISTER ET AL., 1982) ..........................................................................75
FIGURE 64. IDEALIZED ELASTOPLASTIC EARTH RESPONSE-DEFLECTION CURVE WITH TWO SUBGRADE REACTIONS (DAWKINS 1994B) ........76
FIGURE 65. ELASTOPLASTIC SAND SUBGRADE DIAGRAM USING REFERENCE DEFLECTION METHOD (WEATHERBY ET AL., 1998) .............77
FIGURE 66. ELASTOPLASTIC CLAY SUBGRADE DIAGRAM USING REFERENCE DEFLECTION METHOD (WEATHERBY ET AL., 1998) ..............77
FIGURE 67. GROUND ANCHOR T-Y CURVE (STROM AND EBELING, 2001) ....................................................................................78
FIGURE 68. SHIFTED R-Y METHOD TO MODEL CONSTRUCTION STAGES (WEATHERBY ET AL, 1998) ..................................................79
FIGURE 69. A SCHEMATIC SHOWING THE DISCRETIZATION OF DEEP EXCAVATION PROBLEM INTO A FINITE ELEMENT MESH......................81
FIGURE 70. TYPICAL EXCAVATION SEQUENCE IN DEEP EXCAVATION SUPPORTED BY STRUTS (HASHASH & WHITTLE, 1996) ....................82
FIGURE 71. INACCURATE EVALUATION OF THE DIFFERENTIAL SETTLEMENT AFFECTING BUILDINGS DUE TO THE UTILIZING OF
UNREPRESENTATIVE CONSTITUTIVE MODEL (KUNG, 2010) .............................................................................................. 83
FIGURE 72. VARIATION OF MODULUS WITH STRAIN LEVEL (OBRZUD, 2010)................................................................................84
FIGURE 73. STRESS REVERSAL APPROACH (GUTIERREZ ET AL., 2002)..........................................................................................85
FIGURE 74. INTERFACE STRESSES DURING EXCAVATION AND TENSIONING (STROM AND EBELING 2001) ...........................................86
FIGURE 75. DISPLACEMENT FIELDS: (A) INCREMENTAL DISPLACEMENT FIELD FOR WIDE EXCAVATION; (B) INCREMENTAL DISPLACEMENT
FIELD FOR NARROW EXCAVATION; (C) PLASTIC DEFORMATION MECHANISM FOR CANTILEVER RETAINING WALLS IN UNDRAINED
CONDITIONS. (OSMAN & BOLTON, 2007; LAM & BOLTON, 2011) .................................................................................. 88

FIGURE 76. ANN ELEMENTS (MAREN ET AL., 1990; FAUSETT, 1994; SHAHIN ET AL., 2001 & 2008) ............................................90
FIGURE 77.COMPARISON BETWEEN MEASURED GROUND DEFORMATIONS AND ANN PREDICTIONS (FAYED, 2002) .............................92
FIGURE 78. TEST SETUP UTILIZED BY LAEFER (2001) ...............................................................................................................93
FIGURE 79. SOIL SURFACE SETTLEMENT OBTAINED BY LAEFER (2001) PLOTTED WITH RESPECT TO PECKS (1969A) ZONES....................93
FIGURE 80. CENTRIFUGE MODEL PACKAGE FOR EXCAVATIONS CUT AND PROPPED IN FLIGHT (LAM ET AL, 2011) .................................95
FIGURE 81. POTENTIAL BENEFITS OF THE OM ACCORDING TO CIRIA 185 (NICHOLSON AT AL., 1999) ............................................97
FIGURE 82. MEASURING POINT FOR MONITORING SURFACE SETTLEMENT ..................................................................................102
FIGURE 83. INCLINOMETER MEASUREMENT OF DISPLACEMENT ................................................................................................103
FIGURE 84. ROD EXTENSOMETER.......................................................................................................................................103
FIGURE 85. MAGNETIC MULTIPLE POINT EXTENSOMETERS.......................................................................................................104
FIGURE 86. FEATURES OF DIFFERENT PIEZOMETERS (MURRAY, 1990) ......................................................................................106
FIGURE 87. DIFFERENT TYPES OF VIBRATING WIRE STRAIN GAUGES (EL-NAHHAS, 1980) ..............................................................107
FIGURE 88. A REFLECTORLESS ROBOTIC TOTAL STATION (RRTS) MEASURING RSPS AND PRISMS (TAMAGNAN & BETH, 2012) ..........108
FIGURE 89. OPERATING SCHEMATIC OF A TLS SCANNER (LATO, 2012) .....................................................................................109
FIGURE 90. SETTING TRIGGER LEVELS FOR A BUILDING SUBJECT TO SETTLEMENT FROM A DEEP EXCAVATION .....................................110
FIGURE 91. INSTRUMENTED SECTION AT ORABI STATION, CAIRO METRO LINE 1 (EL-NAHHAS, 2006) ...........................................111
FIGURE 92. MEASURED PORE WATER PRESSURE BELOW THE DEEP EXCAVATIONS OF CAIRO METRO LINE 1 ....................................112
FIGURE 93. LAYOUT THE MONITORING SETTLEMENT POINTS (ABDEL RAHMAN & EL-SAYED; 2002A, 2002B & 2003; EL-SAYED & ABDEL
RAHMAN, 2002; ABDEL RAHMAN & EL-SAYED, 2009) ...............................................................................................112
FIGURE 94. THE MONITORING SYSTEM FOR AL-TAHRIR GARAGE (ABDEL-RAHMAN, 2007). ..........................................................113
FIGURE 95. DEFINITION OF THE DEFORMATIONS AFFECTING THE BUILDING BASED ON BURLAND & WROTH (1974 & 1975) AND
BOSCARDIN & CORDING (1989)..............................................................................................................................115
FIGURE 96. DEFINITION OF SAGGING AND HOGGING DEFORMATION MODES (MODIFIED FROM FRANZIUS, 2003) ..............................116
FIGURE 97. DAMAGE CRITERIA BASED ON ANGULAR DISTORTION (BJERRUM, 1963)....................................................................118
FIGURE 98. DEEP BEAM MODEL (BURLAND AND WROTH, 1974 & 1975; BURLAND ET AL. 1977; BURLAND ET AL., 2001) ..............119
FIGURE 99. THRESHOLD OF DAMAGE FOR SAGGING OF LOAD BEARING WALLS, E/G = 2.6 (BURLAND AND WROTH, 1974 & 1975) ....123
FIGURE 100. THRESHOLD OF DAMAGE FOR HOGGING OF LOAD BEARING WALLS, E/G = 2.6 (BURLAND AND WROTH, 1974 & 1975) .123
FIGURE 101. RELATIONSHIP OF DAMAGE TO ANGULAR DISTORTION AND HORIZONTAL EXTENSION STRAIN
(BOSCARDIN &
CORDING, 1989) ..................................................................................................................................................124
FIGURE 102. TENSILE STRAIN COMPONENTS DUE TO HORIZONTAL STRAIN, ANGULAR DISTORTION AND TILTING FOR WALL WITH L/H=1 &
E/G =2.6 (SON & CORDING, 2005) ........................................................................................................................125
FIGURE 103. DAMAGE ZONES WITH DIFFERENT CRITICAL TENSILE STRAINS (SON & CORDING, 2005) ..............................................125
FIGURE 104. DAMAGE CRITERION ACCORDING TO BURLAND (1997). ......................................................................................126
FIGURE 105. BOONES (1996 & 2001) PROCEDURE TO ASSESS THE EXPECTED CRACK WIDTHS......................................................127
FIGURE 106. DAMAGE CRITERION ACCORDING TO CRACK WIDTHS (BOONE,1996). ....................................................................127
FIGURE 107. THREE-PHASED DAMAGE ASSESSMENT FLOW CHART (BURLAND, 1995; MAIR ET AL., 1996; AND SON & CORDING, 2005)
.........................................................................................................................................................................129
FIGURE 108. EFFECTS OF E/G AND NEUTRAL AXIS LOCATION FOR DEEP BEAM ANALYSIS (FINNO ET AL., 2005) ................................131
FIGURE 109. ESTIMATION AND OF THE EQUIVALENT WALL MODULII .........................................................................................132
FIGURE 110. EFFECT OF GRADE BEAMS FOR TWO-STORY AND THREE-BAY STRUCTURES.................................................................133
FIGURE 111. THE MODIFICATION FACTOR FOR THE DEFLECTION RATIO (GOH, 2010, GOH AND MAIR, 2011)..................................134
FIGURE 112. THE MODIFICATION FACTOR FOR THE HORIZONTAL STRAIN (GOH, 2010, GOH AND MAIR, 2011) ...............................135
FIGURE 113. BASIC PROBLEM FOR THE PILE RESPONSE (POULOS & CHEN, 1997).......................................................................137
FIGURE 114. BASIC BENDING VERSUS DISTANCE FROM EXCAVATION FACE (POULOS & CHEN, 1997). ............................................137
FIGURE 115. BASIC MOVEMENT VERSUS DISTANCE FROM EXCAVATION FACE (POULOS & CHEN, 1997). ........................................137
FIGURE 116. CORRECTION FACTORS FOR BENDING MOMENT (POULOS & CHEN, 1997). .............................................................138
FIGURE 117. CORRECTION FACTORS FOR LATERAL PILE MOVEMENT (POULOS & CHEN, 1997). ....................................................139
FIGURE 118. ITERATIVE DESIGNING OF EXCAVATION SUPPORT SYSTEMS FOR ADMISSIBLE DEFORMATIONS (ZAPATA-MEDINA, 2007) ...140
FIGURE 119. CRACK PATTERNS ASSOCIATED WITH DIFFERENT MODES OF GROUND SETTLEMENTS ...................................................141
FIGURE 120. CRACK PLOTTING ON BUILDING ELEVATION ........................................................................................................142
FIGURE 121. AL-TAHRIR GARAGE, ITS SURROUNDING STRUCTURES AND MONITORING SYSTEM (ABDEL-RAHMAN, 2007). ..................145
FIGURE 121. PREDICTED LATERAL DISPLACEMENT OF THE DIAPHRAGM WALL WHILE ADVANCING THE CONSTRUCTION STAGES (ABDELRAHMAN, 2007) ..................................................................................................................................................146

LIST OF TABLES
TABLE 1: SUMMARY OF SOME OF THE ACKNOWLEDGED INTERNATIONAL EMPIRICAL AND SEMI-EMPIRICAL STUDIES ..............................56
TABLE 2: SUMMARY OF SOME OF THE ACKNOWLEDGED NATIONAL EMPIRICAL AND SEMI-EMPIRICAL STUDIES .....................................70
TABLE 3. WEIGHTED VALUE OF DIFFERENT TYPES OF DEFORMATION MEASUREMENTS (NEGRO ET, 2009) ........................................101
TABLE 4. WEIGHTED VALUE OF DIFFERENT TYPES OF STRESS MEASUREMENTS (NEGRO ET, 2009) ...................................................101
TABLE 3. EXPRESSION FOR THE LIMITING DEFLECTION RATIO (BURLAND & WROTH, 1974 & 1975; BURLAND ET AL., 1977) .............119
TABLE 4. DAMAGE CATEGORIES ACCORDING TO BURLAND ET AL. (1977) & BRE DIGEST 251 (1995) ...........................................121
TABLE 5. RISK CATEGORIES ACCORDING TO RANKIN (1988) ....................................................................................................128
TABLE 6. LEVELS OF BUILDING DAMAGE VERSUS DPI THRESHOLDS (SCHUSTER ET AL. 2009) ........................................................128
TABLE 7. EXAMPLES OF UNCERTAINTY IN THE GEOTECHNICAL WORKS (PATEL ET AL., 2007) ..........................................................143
TABLE 8. CONTINGENCY PLANS FOR DEEP EXCAVATION (ABDEL-RAHMAN, 2007) .......................................................................147
TABLE 9. SUMMARY OF THE FINDINGS OF THE COMMON EMPIRICAL/SEMI-EMPIRICAL METHODS FOR AVERAGE WORKMANSHIP IN TERMS OF
THE EXCAVATION DEPTH (HE) OR THE TRENCH/PILE DEPTH (D).......................................................................................149

The most fruitful research grows out of practical problems.


Ralph Peck

1.

INTRODUCTION

There is a worldwide increasing demand to utilize the underground space in the developments of
the urban congested areas for different purposes such as transportation tunnels, underground
parking garages, basements and utilities. Such developments call for deep vertical excavations
and underground tunneling that are frequently close to existing structurally-sensitive buildings
and utilities. As deep excavations initiate lateral and vertical ground deformations due to the
stresses relaxation and bottom heave associated with the excavation process, the adjacent
buildings and buried utilities become kinematically loaded by the induced ground deformations
which depend in magnitude and direction on the building proximity to the excavations as
schematically demonstrated in Figure 1. It is well-acknowledged that the control of ground
movements and protection of adjacent or overlying structures is a major element in the design
and construction of deep excavations and tunneling in urban areas (Gill & Lukas, 1990; Son,
2003; Son & Cording, 2005 & 2007; Hsiao, 2007; Zapata-Medina, 2007; Lam, 2010 and others).

Figure 1. Ground and building deformations induced by a deep excavation (Hsiao, 2007)

To date, failures of structures or roadway adjacent to excavation occur despite the recent
advances made in assessing the stability of excavations and the effects of excavations on nearby
properties. Figure 2 shows a very recent example of a failure case history for a collapsed 13-floor
building by toppling in Minhang District of Shanghai, China. The failure, which happened in in
2009, was due to a nearby deep excavation which overloaded the piles of the collapsed building.
Chai et al. (2014) indicated that the failure was initiated by lateral overloading on the pile
foundation due to excavation near one side of the collapsed building and stockpiling the
excavation at another side of the building. The unbalanced excavation and fill on the sides of the
collapsed building induced lateral loads on piles were also accompanied by unforeseen soil
softening due to a rain event. This failure case history indicates the viral need for supporting
walls for deep excavations in urban areas even in soils that can sustain cuts to avoid affecting the
adjacent structures with the induced deformations. Commonly, a wall is required to support deep
excavations especially in urban areas to minimize the induced deformations. Therefore, the term
deep excavation herein is meant as a supported deep vertical excavation by means of a peripheral
wall. Deep excavations are also termed herein and in the literature as braced excavations (Puller,
2003).
9

Another very well-known recent failure is the failure of Nicoll Highway in Singapore, Figure 3,
which occurred due to insufficient site investigations, misinterpretation of the observations,
faults in design of the bracing system, and utilization of unsuitable method for wall strutting by
jet grouting (Whittle & Davies, 2006; Lee, 2008).

Figure 2. Failure of a building in China in 2009 that was initiated by a nearby deep excavation

Figure 3. Failure of a deep excavation adjacent to Nicoll Highway, Singapore (Lee, 2008)

10

Serviceability problems associated with the substantial foundation settlement and lateral
deformations induced by deep excavations are much more widespread than failures. Structure
may experience distresses such as cracking of structural or architectural elements, uneven floors,
or inoperable windows and doors due to the induced deformations. Figure 4 shows an example of
a cracked external wall due to a nearby excavation. The amount of the tolerable deformations
and the severity of excavation-related damage depend on the building type, configuration and
stiffness as well as the characteristics of excavation support, the ground geotechnical conditions
and the construction sequence. Both geotechnical and structural engineers are required to
collaborate in quantifying the amount of building settlement, assess the possible structural
damages and set up the counter measures and risk mitigations to avoid such damage (Boscardin
and Cording, 1989; Burland, 1995; Boone, 1996 & 2001; Boone et al., 1998 & 1999; Long,
2001; Finno and Bryson, 2002; Finno et al., 2002; Son, 2003; Son and Cording, 2005 & 2007;
and others).

Figure 4. A masonry wall suffered from severe cracking due to ground deformations (Vatovec et al., 2010)

It is acknowledged that the effect of deformations associated with deep excavation depends on
the geotechnical characteristics of the soils. The less strength and more compressible the soils
have, the more pronounced effects and deformations are anticipated. Awkwardly, most of the
deep excavations are in urban areas that have deltaic soils originating from rivers and oceans;
they comprise sediments such as silts, clay and sands under shallow groundwater table. Such
deltaic soils are often encountered in the most densely populated areas in the world. This fact
emphasizes the need to predict, control and mitigate the deformations resulting from deep
excavations (Peck, 1969a; El-Nahhas, 1992 & 2006; Boone, 1996 & 2001; Bolton, 2008; and
others).
11

Nationally, most of the developments that need deep excavations in Egypt are located in the
Greater Cairo area which is characterized by recent Nile alluviums with shallow groundwater
table. Geologically, the Nile developed its course in this area through the down faulting of the
limestone extending between the El-Muqattam cliff and the Pyramids plateau and deposited
recent alluviums of alternating layers of cemented silty sand, clayey sand and medium to coarse
sand underlain by very stiff plastic clay that rests on the limestone marine formations as
illustrated in Figure 5 (Said, 1981; El-Sohby & Mazen, 1985; El-Ramli, 1992; El-Nahhas, 2006;
and others).

Figure 5. Typical formations in the Greater Cairo area (El-Sohby and Mazen, 1985)

The geotechnical conditions of the Nile alluviums are considered problematic for deep
excavation particularly as the expected deformations impose risks on the adjacent structures and
utilities including possible loss of support to existing foundations and structurally distressing
buildings, pavements and utilities surrounding the excavation. Notwithstanding these
engineering challenges, there is an ever growing need for utilization of underground space in
Greater Cairo in the last decades due to the scarcity of the ground space and the high cost of
lands in this area (El-Nahhas et al., 1988 & 1990; El-Nahhas, 1992 & 2006; Abdel-Rahman,
1993; Abd El-Salam, 1995; Ahmed & Abd El-Salam, 1996; Ahmed et al., 2005; Abdel-Rahman,
2007; Abdel-Rahman & El-Sayed, 2009 and others).
Precise evaluation of the ground displacements induced by a deep excavation is not simple to be
achieved due to the uncertainties in soil properties, constitutive modeling, construction stages,
three-dimensional and time-dependent natures of the problem, and the need for incorporation of
human factors such as workmanship in the models. Notwithstanding that, reasonable assessments
can be reached if the diverse methods for analysis are carefully studied by an experienced
geotechnical engineer to reach a solid evaluation. Predicting the induced movements and
mitigating them by suitable means became more feasible with the development of observational
methods and the non-linear finite element analysis procedures and software since the early
1970s. (Peck, 1969b; Lambe, 1970; Clough & Duncan, 1969 & 1971; Duncan & Clough, 1971;
Goldberg et al. 1976 ; ORourke et al., 1976; Boscardin et al., 1979).
12

Generally the methods to obtain the deformation field can be categorized into empirical/semiempirical methods, numerical methods, analytical methods, physical/centrifuge modeling, and
Artificial Neural Networks (ANNs). It is also to be noted that the assessment of deformations
associated with deep excavation depends if there is a building in the vicinity of the excavation or
not. For the case of no buildings, the ground deformations are designated as free-field or
greenfield. The presence of the building modifies the induced deformation due to the building
weight and stiffness. Heavy flexible buildings may have more deformations that the expected
greenfield while light rigid building may have less deformations that the anticipated greenfield
(Chang, 1969; Chandrasekaran & King, 1974; Clough & Manna, 1976; Brown & Booker, 1986;
Powrie & Li, 1991; Ng, 1992; Abdel-Rahman, 1993; Morrison, 1995; Bentler, 1998; El-Nahhas
et al., 1989, 1994 & 1998; Seok et al., 2001; Fayed, 2002; El-Nahhas & Morsy, 2002; and
others)
The traditional single, fully developed design with no intention to vary the design during
construction does not exist in geotechnical engineering, particularly, in deep excavation and
tunneling projects. Peck (1969b) coined the observational approach to be adopted in geotechnical
projects. In this approach, the instrumentation and monitoring are required to be carried out to
provide confidence to the administratively controlling authorities and the affected third parties
such as the owners of the adjacent buildings. Monitoring results often integrate with the design
and enhance the reliability of the design assumptions by validating the design parameters as the
construction proceeds. Inverse analysis of the monitoring data provides the appropriate tool to
combine observational and analytical approaches to enhance risk mitigations and managements
(Clough, 1975; Powderham, 1994, 1998 & 2002; Powderham & Nicholson, 1996; Powrie &
Kantartzi, 1996; Nicholson & Penny, 1999; Hashash et al., 2004 & 2010; Abdel-Rahman, 2007;
Lee at al., 2007; and others).
In this state-of-the-art report, the following issues are highlighted:
1. Factors affecting the deformations associated with deep excavations;
2. Assessment of the ground deformations outside the deep excavation using different
approaches (viz., empirical/semi-empirical, numerical; analytical, physical modeling and
ANN approaches);
3. Influence of the presence of buildings on the displacements;
4. Structural damage criteria;
5. Monitoring programs for deep excavation projects; &
6. Risk management and mitigation for deep excavation projects
In addition to the International studies presented in this state-of-the-art covering the
abovementioned points, National experiences in assessments of the deformations induced by
deep excavations, monitoring programs, risk assessment and risk mitigation are also highlighted.

13

2.

FACTORS AFFECTING EXCAVATION DEFORMATIONS

Ground deformations associated with deep excavations are inevitable. The relaxation of the
horizontal stress by the excavation induces horizontal movements of the wall and the soil
towards the excavation accompanied by vertical deformations of the soil around the excavation.
The vertical deformations are mostly downward deformations (settlements); yet, sometimes
upward deformations (heaves) are measured adjacent to the wall or at far distances from the wall.
Settlement may be associated with the instability of the excavation base in clayey soils.
Deformations may also occur due to the increases in the effective stresses during lowering
groundwater table (Caspe, 1966; Goldberg et al., 1976; ORourke, 1981 & 1993; Clough and
ORourke, 1990; Ou et al., 1993 & 2000; Hseih & Ou, 1998; Poh et al., 2001; Kung, 2003; and
others).
Prior to the revolutionary state-of-the art paper of Peck (1969a) in which he demonstrated that
substantial deformations associated with deep excavations and tunneling may occur,
geotechnical engineers used to assume that that deformations are negligible if the excavation
have adequate factor of safety against potential failures. Later, ORourke et al., (1976) and
Boscardin et al. (1979) reported that sensitive structures were damaged due to the deformations
induced by an adjacent deep cut in Washington D.C., even though the bracing experienced no
structural distress. Since then, it has been acknowledged that excavations are commonly
accompanied by significant deformations that may considerably affect adjacent facilities even if
they have adequate factor of safeties against possible failures.
It is a common practice to support deep excavations by continuous walls in urban areas to limit
the induced movements. The excavation support systems for deep excavations consist of two
main components: a wall, and its lateral supporting elements. Many types of walls and supports
have been used in deep excavations. Walls supporting deep excavation may be classified into
following three major categories according to the form of supporting measures provided for
them:
1. Cantilevered wall (usually for shallow excavation);
2. Strutted/braced wall; &
3. Tied-back or anchored wall
Under each of the above support category, the following wall types may be utilized:
1. Sheet pile wall;
2. Soldier pile and lagging wall (Berlin wall);
3. Contiguous bored piles wall;
4. Secant piles wall;
5. Diaphragm wall; &
6. Soil-mixing walls
Puller (2003) described the aforementioned systems and other less widely used support systems
in considerable detail. The excavation-induced deformations may be affected by a large number
of factors such as: wall stiffness, ground conditions, groundwater condition and control
measures, excavation depth, construction sequences, and workmanship. The following sections
address some of the important factors that profoundly affect the induced deformation and hence
the associated building damage.
14

2.1.

Soil Type

Peck (1969a) showed that settlements next to deep excavations correlate to soil type. He
proposed three zones of settlement profiles based on the prevailing soil conditions as illustrated
in Figure 6. In general, larger wall deflection and ground deformations are induced due to
excavations in soils with lower strength and stiffness. The same conclusion was reached in many
other following studies (e.g., Goldberg et al., 1976; Clough & ORourke, 1990; Bentler, 1989;
and others). This aspect is further elaborated in Section 3 of this state-of-the-art addressing
empirical and semi-empirical method for assessment of the deformation induced by deep
excavation.

Figure 6. Effect of the soil type on the settlements induced by deep excavation (Peck, 1969a)

2.2.

Wall Stiffness and Excavation Stability

Stability and deformation are interrelated. Walls with large factors of safety against potential
collapses have small strains around the excavation; consequently, the ground deformations are
also likely to be also small. Conversely, if the factors of safety (or some of them) are small,
strains around the excavation and ground deformations may become large. Additionally, the wall
stiffness greatly affects the induced ground movements. Goldberg et al. (1976) showed using
finite element and measured data that the maximum lateral deformations for deep excavations in
clays can be estimated using the stability number of the excavation H/cu (where is the soil unit
weight, H is the depth of the excavation and cu is the undrained shear strength) and the stiffness
of the supporting system EwIw/h4 (where Ew is the youngs modulus of the wall, Iw is the moment
of inertia of the wall per linear meter, h is a representative unsupported length of the wall such as
the average distance between struts). Figure 7 illustrates the findings of Goldberg et al. (1976).
15

Figure 7. Effect of wall stiffness and soil stability number on the wall deformations in clays (Goldberg et al., 1976)

Mana & Clough (1981) utilized the finite element and the field measurements to relate the
maximum wall movements with the factor of safety against basal heave in clays as shown in
Figure 8. The quasi-constant non-dimensional movement are at high safety factor is an indication
of an elastic response. The rapid increase in movements at lower factor of safety is a result of the
induced plastic deformations.

Figure 8. Effect of the basal heave stability on the wall deformations induced by deep excavations in clays (Mana &
Clough, 1981)

16

Clough et al. (1989) and Clough & ORourke (1990) utilized the nonlinear finite elements and
field measurements to determine the effect of the wall stiffness on the maximum lateral wall
movement in clays that is induced by excavation. They introduced a system stiffness factor,
similar to Goldberg et al. (1976), for estimating wall stiffness of unit thickness (plane strain)
which depends on wall material, section properties and support spacing; this factor is giving by:
EI
k
(1)
4
w have
where k
= Dimensionless system stiffness
E
= Youngs modulus of wall system
I
= Moment of inertia of wall system
have
= average vertical distance between tiebacks/struts
w
= unit weight of water = 9.81 kN/m3
The results of their analyses are shown in Figure 9.

Figure 9. Effect of the basal heave stability and the system stiffness on the wall deformations induced by deep
excavations in clays (Clough at al., 1989)

According to the introduction of the system stiffness factor, the retaining system can be
categorized into two-categories:
1. Flexible systems (e.g., sheet-pile walls, soldier beam and lagging): Generally the
stiffness factors k for these systems are less than 40.
2. Stiff systems (e.g., secant pile wall, tangent pile wall, diaphragm walls): Generally the
stiffness factors k for these systems are greater than 300.
The selection of system stiffness and spacing generally relies on economic as well as practical
issues such as minimum spacing to accommodate construction activities.

17

Clough at al. (1989) and Clough & ORourke (1990) concluded that the wall stiffness is less
effective in reducing movements than in cases with high factor of safety against basal instability.
However, in case of having low factor of safety against basal heave, the wall stiffness affects the
deformation greatly. The aforementioned studies suggest that the maximum lateral wall
movement for stiff systems (i.e., thick diaphragm walls or secant piles walls) in stable soils (i.e.,
factor of safety against bottom heave is greater than 3) is limited to approximately 0.2% of the
excavation depth regardless of the system stiffness.
Clough et al. (1989) ignores the increase of stability due to wall embedment. Hashash et al.
(2008) showed that the wall embedment and stiffness may limit the soil movements to much
lower magnitudes than what is predicted in Figure 9.
Long (2001) analyzed 296 case histories and checked Clough et al. (1989) chart against the study
data. Substantial scatter was noted as shown in Figures 10 & 11. He concluded that the wall
stiffness does not affect the deformation if the excavation has a factor of safety against basal
heave more than 3.

Figure 10. Normalized field measurements of the lateral deformations against Clough & ORourkes (1990) system
stiffness and basal heave factor of safety for cases with low factor of safety (FOS<3) (Long, 2001)

18

Figure 11. Normalized field measurements of the lateral deformations against Clough & ORourkes (1990) system
stiffness and basal heave factor of safety for cases with high factor of safety (FOS>3) (Long, 2001)

Moormann & Moormann (2002) and Moormann (2004) reached the same conclusion of Long
(2001) that Clough et al. (1989) and Clough & ORourkes (1990) system stiffness factor needs
to be revisited after reviewing more than 500 case history in both soft and stiff clays. Figures 12
& 13 show the substantial scatter of the database points with respect to Clough et al.s (1989)
curves. Moormann (2004) attributed the lack of dependency of lateral movements on system
stiffness as predicted by to the factors like:
1. Soil conditions at the embedment portion of the wall;
2. Groundwater conditions;
3. Effect of the surrounding buildings or geometrically irregularities;
4. Workmanship;
5. Unforeseen events and excavation sequence;
6. Pre-stressing of struts and anchors; &
7. Time-dependent effects.

19

Figure 12. Normalized field measurements of the lateral deformations against Clough & ORourkes (1990) system
stiffness and basal heave factor of safety for soft clay (Moormann, 2004)

Figure 13. Normalized field measurements of the lateral deformations against Clough & ORourkes (1990) system
stiffness and basal heave factor of safety for stiff clay (Moormann, 2004)

20

Zapata-Medina (2007) proposed a revised system stiffness factor which gives more reliable
results with the data using the data of 30 case histories that comprise soft, medium and stiff
clays. A graph showing the favorable correlation between the maximum lateral deformation, the
factor of safety and revised system stiffness is shown in Figure 14. The recommendations of
Zapata-Medina (2007) are further elaborated in the Section 3 in this state-of-the-art report.

Figure 14. Normalized lateral wall movements vs. relative stiffness ratio, R, for deep excavations in cohesive soils
(Zapata-Medina, 2007).

It is to be noted that Zapata-Medinas (2007) utilized Ukritchon et al.s (2003) factor of safety
(FS) against basal heave as shown in Figure 14. The factor of safety could reach as low values as
0.65 in the analyses without having a failure of the excavation; instead, large ground
deformations were observed. This note was not clarified by Zapata-Medina (2007)
Juang et al. (2011) explained the lack of dependence of the induced deformations on the system
factor by stating that the ground movements are essentially functions of the following six
parameters in addition to the system stiffness:
1. Excavation depth;
2. Excavation width;
3. The depth from the bottom of excavation to the hard stratum;
4. The normalized clay layer thickness (Hclay /Hwall) where Hclay is the sum of the
thicknesses of the clay layers and Hwall is the wall depth;
5. The ratio of shear strength over vertical effective stress (su /v); &
6. The ratio of initial Youngs tangent modulus over vertical effective stress (Ei /v)
The factor of safety against basal failure also affects the shape of the settlement trough
associated with deep excavation. Mana and Clough (1981) found by numerical analyses that the
width of the settlement trough increases with the increase of the factor of safety as shown in
Figure 15.
21

Figure 15. Effect of the factor of safety on the settlement trough width (Mana and Clough, 1981)

2.3.

Overconsolidation (OCR) and At-Rest Earth Pressure Coefficient (Ko)

Overconsolidated soils generally have higher at-rest lateral pressure (Ko) than normally
consolidated soil (Mayne & Kulhawy, 1982). Potts & Fourie (1984) studied the behavior of a
single propped retaining wall has been using elasto-plastic finite element method. They
concluded that by increasing the at-rest coefficient (Ko), the deformations, forces and bending
moments in the wall substantially increase and even may exceed those calculated using the
simple limit equilibrium approach which is in common use. The behavior of excavated walls in a
high-Ko soil is dominated by the vertical unloading forces caused by the excavation as shown in
Figure 16. Additional horizontal restraint in the form of multi-propping, while reducing
horizontal movements of the wall and soil, has a much smaller effect on vertical movements.
Peck (1969a) noted that in highly overconsolidated clays, soil tends to heave near to the wall.

Figure 16. Contours of stress level at an excavation depth of 13.26m: (a) Ko=2; (b) Ko=0.5 (Potts & Fourie, 1984)

22

2.4.

Groundwater Conditions and Control Measures

Groundwater develops hydro-pressure against the walls of the deep excavation supporting
system causing them to deform which adds to the ground deformations associated with deep
excavations. Additionally, soils under water are generally weaker than being above them due to
the effect of water in reducing the effective stress. Moreover, the flow of groundwater towards
the excavations may endanger the excavations and the surrounding buildings, particularly if it
occurs through the wall itself in lieu of the dewatering system. The different groundwater flow
patterns associated with deep excavations are shown in Figure 17 (Clough & ORourke, 1990).

Figure 17. Groundwater flow patterns encountered in deep excavations (Clough & ORourke, 1990)

Settlements are generated by the groundwater table lowering as the soil is passing from a
submerged to a saturated unit weight which leads to an increase of the effective stress as shown
in Figure 18. The settlement value depends on the drawdown of the water table and the soil
stiffness. In sands, excessive pumping out the groundwater from a deep excavation results in a
significant drop of the groundwater table within the surrounding areas with possible excessive
settlement of the adjacent buildings and other structures and piping if the exist hydraulic gradient
at the bottom of excavation exceeded the safe value. Puller (2003) summarized the groundwater
control measures for deep excavations as shown in Figure 19.
23

Figure 18. Influence of the dewatering works on the ground settlement

Figure 19. Groundwater control measures for braced excavation (Puller, 2003)

24

Examples of groundwater-related failures and problems that occurred to deep excavations due to
improper groundwater considerations in design and construction:
1. The collapse of a deep excavation for an underground metro station in Cologne,
Germany in 2009, Figures 20 & 21, which in-turn caused the collapse of the historical
City Archive Building. This failure is anticipated to be a piping failure induced by the
groundwater high velocity that was not considered during the design of the dewatering
system (Rowson, 2009).
2. A diaphragm wall leaked during the construction of a deep exaction for a new
underground station of the North-South Train Line in Amsterdam, the Netherlands. This
leakage caused washing of sand below the foundations of surrounding buildings and a
subsequent subsidence of 23 cm as shown in Figure 22. The predicted costs have gone
up from 1.5 to 3 billion euros and the project completion was shifted from 2011 to 2017
(Van Tol, 2010; Van Baars, 2011).
3. In 2005, a diaphragm wall leaked and surrounding houses started to subside in a deep
excavation for a garage in Middelburg, The Netherland. To stop the subsidence, the pit
was filled with water until 2009, Figure 23, till new walls were placed in the pit and the
pit was filled with 13,350 m3 of concrete; a loss of almost half the volume of parking
space (Van Baars, 2011).
4. In 2007, a well-known failure of the diaphragm for The Infinity Tower in Dubai
occurred due to piping by seepage through a diaphragm wall joint as shown in Figure 24.

Figure 20. Collapse of City Archive Building in Cologne (Germany) due soil piping induced by dewatering
(Rowson, 2009)

25

Figure 21. The collapsed City Archive Building in Cologne (Germany) (Rowson, 2009)

Figure 22. Damage due to Subsidence along an underground station of the North-South Train Line in Amsterdam
(Van Baars, 2011).

26

Figure 23. Leakage and damage at the building pit in Middelburg, the Netherland (Van Baars, 2011)

(a)

(b)

(c)

(d)

Figure 24. Failure of a diaphragm wall in The Infinity Tower in Dubai in 2007. The chronological sequence of
events is (a) to (d)

27

To avoid problems associated with groundwater and to minimize the effect of groundwater
lowering on the adjacent buildings, the concrete diaphragm walls in the Greater Cairo Metro was
extended deeper without reinforcement and a low permeability grouted plug is provided at their
toes as shown on Figure 25-a to avoid the possible effects of the large groundwater drawdowns
as schematically shown in in Figure 25-b. The grouting materials were injected in two stages:
bentonite-cement slurry and soft-silica gel, in order to reduce the permeability of the sand to 10-6
m/s. Thickness of the grouted plug and its elevation are selected to satisfy a safe limit of the
average hydraulic gradient within the plug (less than 3) and an average buoyancy factor of safety
of 1.1 of the remaining soil mass below the final excavation level (El-Nahhas, 2003 & 2006; ElNahhas et al., 2006).

(b) without plug causing large drawdowns


(not utilized in the Greater Cairo Metro)
(a) With plug (utilized in Greater Cairo Metro)
Figure 25. Schemes for groundwater control in a deep excavation (El-Nahhas, 2006).

2.5.

Strut/Tie-back Prestressing

Support systems for deep excavations consist of two main components: The wall and the support
provided for the retaining wall as struts (braces), rakers, and tieback anchors. Clough (1975)
demonstrated the effects of pre-stressing of braces to control wall deformations. The wall
movement is plotted against the normalized prestressing force for sands and stiff clays as shown
in Figure 26. For both sands and stiff clays, the movements decrease with increasing prestressing
force of the tie-backs. Clough (1975) suggested that the optimum effect of prestressing in
reducing movements is achieved by using pressure levels slightly greater than those of Terzaghi
& Peck (1967).
ORourke (1981) observed that the effective stiffness of braces could be as low as two percent of
the ideal stiffness due to compression of the bracing and its connections without imposing an
initial prestressing. Mana & Clough (1981) showed that increasing the stiffness of the
strut/anchor/raker reduces the deformation 40% as shown in Figure 27.
28

Figure 26. The effect of prestressing on the wall deformations (Clough, 1975)

Figure 27. The effect of the strut stiffness on the maximum lateral deformation of the wall and the maximum
settlement (Manna & Clough, 1981)

29

2.6.

Construction Sequence

Generally, the following two different approaches are frequently utilized in deep excavations:
Conventional (down-top) construction: The excavation between the supporting walls
(i.e., pit excavation) starts after installing the supporting walls and operating the
groundwater control measures. The excavation proceeds sequentially with the
installation of the struts, tie-backs and/or rakers to the foundation level of the permanent
structure. After that, the construction of the permanent structure starts from bottom to
the top. The first element of the permanent structure to be cast is the raft foundations. An
example of this method of construction is the construction of the Greater Cairo Metro Line 1 as show in Figure 28. The supporting wall may or may not be integrated with the
permanent structure. If the walls are be integrated with the structure, then, they should be
diaphragm walls.

Figure 28. Construction procedure steps for the Greater Cairo Metro - Line 1 (El-Nahhas et al., 1988)

Top-down construction: In this method of construction, the basement slabs are formed
and poured on the existing subgrade. The top slab of the basement in the permanent
structure is cast after installing the supporting walls (walls should be diaphragm walls).
The top basement slab is considered the first support to the wall. After that, operation of
the dewatering system and pit excavation under the top slab proceed till reaching the
level of the second slab which is to be cast against the subgrade at this stage. Temporary
supports are also installed while excavating. Excavation shall continue to reach the
foundation level and then the raft foundation is to be cast. Excavation supporting walls
are always integrated within the permanent structure. An example of the top-down
construction is Rod El-Farag Station in the Greater Cairo Metro Line 2 as illustrated in
Figure 29.
30

(a) configuration of the station

(b) stages of construction


Figure 29. Rod El-Farag Station (Ahmed and Abd El-Salam, 1996)

31

It is anticipated the top-down construction has much less deformations than the conventional
down-top construction due to the early installation of the top slab which acts as a support for the
wall. However, since the tope slab cannot be prestressed, it appears that the conventional (downtop) construction with prestressed anchors/struts gives less cantilever deformations than the topdown construction in soft clays especially as the top slab may suffer from shrinkage after its
casting and this limits its efficiency (Long, 2001).
It is also to be noted that the top-down construction helps to eliminating the top supports such as
tie-backs which may interfere with the foundations of the adjacent buildings (especially if they
are shallow foundations) and the nearby utilities. On the other hand, top-down construction
reduces the rate of excavation since excavations works start under the cast slabs in a restricted
narrow space.

2.7.

Wall Lateral Deformation Patterns

Goldberg et al (1976) identified different settlement patterns following the wall lateral
deformations patterns as shown in Figure 30. They showed that the settlement behavior do not
only depend on soil type but also on the wall lateral deformations as well.

Figure 30. Settlement patterns associated with different wall deformation modes (Goldberg et al. 1976).

32

Clough and ORourke (1990) According to the method of construction the wall deform in two
modes: cantilever mode, and bulging mode. The settlement troughs associated with each mode
are different as shown in Figure 31. Boone (2003) and Boone & Westland (2005) concluded the
same effect of wall deformation on surficial settlement trough as shown in Figure 32.

Figure 31. Modes of deformation of the wall (Clough and ORourke, 1990)

Figure 32. Lateral and vertical displacement patterns: concave on left, spandrel on right (Boone 2003; Boone &
Westland, 2005).

ORourke (1981) envisaged a factor called the Coefficient of Deformation (CD) which is defined
as the ratio of the cantilever deformation component (Sw) to total deformations (Sw + Sw) where
Sw is the bulging component of the wall displacement. Figure 33 shows the relationship between
CD and ratio of the maximum wall lateral deformation to the maximum ground settlement
relationship for clays. Accordingly, the maximum surficial settlement associated with the
cantilever mode (CD=1) is about 0.63 times the maximum lateral cantilever deformation of the
wall; while, the maximum settlement associated with the wall bulging mode (CD=0) is about 2
times the maximum lateral wall bulging deformation. Mana & Clough (1981) suggested that the
maximum vertical cumulative deformation is ranged between 0.5 & 1 times the maximum lateral
deformation of the wall based on field measurements as shown in Figure 34.
33

Figure 33. The ratio between the maximum horizontal to vertical displacement as a function of the Coefficient of
Deformations (ORourke, 1981)

Figure 34. The ratio between the maximum horizontal to vertical displacement (Mana & Clough 1981)

In the following sections, the settlement patterns associated with cantilever and bulging modes of
deformations of the wall are elaborated. The empirical and semi-empirical patterns associated
with the cumulative pattern of the wall are demonstrated in Section 3 of this-state-of-the-art.
34

2.7.1.

Settlement pattern associated with the wall cantilever deformation mode

Caspe (1966), Bowles (1988), Aye et al. (2006) utilized analysis for the induced settlement that
is anticipated to be associated mainly with the wall cantilever mode as shown in Figure 35. The
lateral wall deflection are to be determined using the 1D beam-spring model (refer to Section 4.1
in this state-of-the-art report) and numerically integrated to obtain the volume of the wall
deflection (Vo) utilized in this method.

Figure 35. Deformations prediction from lateral wall deflection values proposed by Aye (2006):

(a) settlements; (b) lateral deformations

Ou et al. (1993) presented a tri-linear settlement profile called spandrel-type settlement based on
10 case histories of deep excavation in soft clays from Taipei, Taiwan. The maximum settlement
is located at the wall when the wall deforms as a cantilever. The settlement trough is shown in
Figure 36.
35

Figure 36. Spandrel-type settlement trough (Ou et al., 1993)

Lee et al. (2007) proposed that lateral Sh and vertical Sv deformations associated with the wall
cantilever mode can be presented in terms of the maximum wall deformations Sw and the trough
width W using Gaussian distribution, Figure 37, as follows:
(3)
(4)

Where is the ratio between the maximum wall deflection and the maximum surface settlement
and can be assumed to be 0.5 for diaphragm wall & 1 for sheet pile wall.

Figure 37. Assumed Gaussian distribution for lateral and vertical ground deformations (Lee et al., 2007)

36

2.7.2.

Settlement pattern associated with the wall bulging deformation mode

Hsieh & Ou (1998) presented a concave settlement profile for the bulging mode of wall based on
analysis of 9 case histories. The maximum settlement is assumed to occur at 0.5 He, where He is
the excavation depth from the wall. The settlement at the wall is approximated to 50% of the
maximum settlement as shown in Figure 38.

Figure 38. Concave settlement profile (Hsieh & Ou, 1998)

Karlsrud (1997a) proposed a relationship between the maximum wall deformation and the
surficial ground settlement concave pattern, Figure 39, based on data from sites with soft clays
and loose to medium dense sand and silts. The dashed lines close to the wall reflects impact of
the potential for movements of the tip of the wall. Thus for structures laying at distances from the
wall smaller than 0.2 times the depth to zero lateral displacement, the settlements may be quite
uncertain.

Figure 39. Relationship between wall movement and ground settlements for soft/loose soils (Karlsrud, 1997).

37

Schuster et al. (2009) proposed a concave settlement pattern along with its associated lateral
deformation patterns as shown in Figure 40. The settlement at the wall is about 20% of the
maximum settlement. The lateral deformation affecting nearby building changes from concave
shape at the ground surface to spandrel shape to depth of 5m depending on the foundation depth
of the building.

Figure 40. Vertical and horizontal ground movement patterns as a function of the excavation depth (He) and the
distance from the wall (d) (Schuster et al. 2009)

For the subsurface settlement associated with concave settlement trough, Aye et al. (2006)
proposed a vertical distribution similar to their recommendations for the spandrel-type settlement
as shown in Figure 41.
38

Figure 41. Subsurface settlement distribution for concave settlement profiles (Aye et al. 2006)

2.8.

Time-Dependent Effects

For an excavation in clays, longer durations for installing the strut or constructing the floor slab
may cause larger wall deflection due to the occurrence of consolidation or creep of clay. Studies
that addressed that aspect by assessing the soil consolidation, as one of the components of the
wall and ground deformations, were carried out based on finite element analysis since it is not
possible to separate the consolidation deformation component out of the total deformations from
the field data.
Osaimi & Clough (1979), Yong et al. (1989), Finno & Harahap (1991), and Ou & Lai (1994)
showed that significant consolidation can take place during the construction of a deep excavation
in clay and that the effects of consolidation are significant. Consolidation and swelling during
excavation result in changes in the shear strength of soils and time-dependent deformations. The
negative water pressure dissipates with time generated by the excavation at the base of the
excavation which causes loss of some passive resistance that occurs immediate after excavation.
This leads to time-dependent deformations in the wall and the soil behind the wall.

2.9.
2.9.1.

Excavation Geometry and Three-Dimensional Effects


Excavation dimensions and depth to firm layers

Manna and Clough (1981) utilized non-linear finite elements to study the effect of the excavation
dimensions and found that increasing the width of the excavation and the depth to firm layer
increase the maximum ground settlement and the maximum wall deflection as shown in Figures
42 & 43. Similarly, Hsiao (2007) demonstrated that the maximum wall deflection has to be
modified by deflection reduction factor (K) due to presence of hard stratum. The deflection
reduction factor (K) is related to the ratio of the depth to hard stratum, measured from the current
39

excavation level, over the excavation width (T/B). At smaller T/B ratios (T/B<0.4), the presence
of the hard stratum has a great influence on the magnitude of the calculated maximum wall
deflection, and at T/B>0.4, the influence of the hard stratum is negligible as shown in Figure 44.

Figure 42. Effect of the excavation width on the maximum ground settlement and the wall deflection (Mana &
Clough, 1981)

Figure 43. Effect of the depth to firm layer on the maximum ground settlement and the wall deflection (Mana &
Clough, 1981)

40

Figure 44. The Effect of the hard stratum on the computed wall deflection (Hsiao, 2007)

2.9.2.

Corner effect

Ou et al. (1996) performed parametric three-dimensional finite element analyses to investigate


the features of three-dimensional deep excavation behaviors. They found that close relationships
existed between the aspect ratio for excavation geometry (B/L) and wall deformation. B and L
are the excavation dimensions in horizontal plane in the direction of lateral wall measurements
and the perpendicular direction, respectively. Increasing the B/L decreases the wall deformation.
Additionally, the wall deformation of a deep excavation is directly related to the smallest
distance from the corner (d). The smaller is the value of d, the less is the wall deformation.
Ou et al. (1996) defined a ratio called the Plane Strain Ratio (PSR). PSR is defined as the ratio
of the maximum wall deformation of the cross section at a distance (d) from the excavation
corner to the maximum wall deformation in the plane strain conditions of the same geometry.
They established the relationship between (PSR), (B/L) & (d) based on the results of parametric
studies, as shown in Figure 45.

Figure 45. Plane strain ratio (PSR) as a function of the aspect ratio (B/L) and distance from the corner (d)
(Ou et al., 1996)

41

2.9.3.

Parallel distribution

Finno & Roboski (2005); and Roboski &Finno (2006) studied deep excavations in soft to
medium clays based on the settlements that were observed using optical survey around a 12.8 m
deep excavation in Chicago. The excavation was supported by a flexible sheet pile wall and three
levels of regroutable anchors. They suggest a parallel distribution for the deformation to account
for the corner effect. They found that the complementary error function (erfc) can be used to
define the three-dimensional settlement distributions of ground movement around excavation of
finite length.

(3)

where max can be either maximum settlement or maximum lateral movement, L is the length of
the excavation, and H is the height of the excavation as presented in Figure 46.

Figure 46. Three-dimensional distribution of settlement and lateral movement around finite deep excavation (Finno
& Roboski, 2005; Roboski &Finno, 2006)

42

2.10.

Wall Installation Effect

The wall installation process can cause significantly movements in the surrounding ground. The
assumption of negligible deformations associated with wall installation may lead to a substantial
underestimation of excavation-related lateral movements (Ng and Yan, 1999; Gourvenec and
Powrie, 1999; Abdel Rahman and El-Sayed, 2002a, 2002b & 2009; El-Sayed and AbdelRahman, 2002).
In a survey of the problematic deep excavations in The Netherlands carried out between years
2007-2012, Korff & Tol (2012) noted that many problematic deep excavation cases have been
reported as the designer of the wall supporting the deep excavation disregarded the installation
effects of the walls and foundations. Although a lot of efforts are often not saved into the design
of the wall stiffness and related assessment of possible damage to properties, the installation and
the associated deformations are often excluded which caused many problems later.
Morton et al (1980), Budge-Reid et al (1984), Cowland & Thorley (1984), and Thorley & Forth
(2002) reviewed the settlements induced by the construction of the diaphragm walls in Hong
Kong, particularly for the Mass Transit Railway project where soils are generally fill, marine
deposits and alluviums underlain by decomposed granite. Settlement values up to 150mm were
reported for shallow foundations while less settlement was reported for deep foundations as
shown in Figures 47, 48 & 49.

Figure 47. Settlement associated with trenching in Hong Kongs MTR (Morton et al., 1980)

43

Figure 48. Maximum building settlements due to slurry trench excavation for diaphragm walls as a function of
foundation depth in Hong Kongs MTR (Cowland & Thorley, 1984)

Figure 49. Building settlement due to diaphragm wall installation in Hong Kongs MTR (Budge-Reid et al., 1984)

44

Clough & ORourke (1990) showed that significant settlement may occur behind a diaphragm
wall after installation (up to 0.15% of the trench depth) as shown in Figure 50. Deep trenches in
Hong Kongs marine and alluvial deposits controlled the data presented by Clough and
ORourke (1990); therefore, it is anticipated that Figure 50 overestimates the ground movements
in most cases.

Figure 50. Settlement due to installation of a diaphragm wall (Clough and ORourke, 1990)

Finno et al. (2002) observed that 25% of the total lateral movement occurred after installation pf
secant piles wall in soft to medium Chicago clay, as can be shown in Figure 51. It was concluded
that lateral movements of this magnitude cannot be neglected and must be taken into account
when designing support systems, especially when sensitive structures are nearby.

Figure 51. Lateral deformation associated with trenching for secant piles installed in Chicago clay
(Finno et al., 2002)

45

Poh and Wong (1998) investigated the influence of specific construction methods utilized to
install the diaphragm wall on the magnitude of lateral displacements. They found that the lateral
displacements decrease only slightly (approximately 10 percent) as the slurry level increases;
while the lateral displacements by approximately 50 percent if the slurry level decreases. They
also noted that increasing the holding time (i.e. time after the completion of the trench, but
before concreting) slightly the lateral soil movements (approximately 20 percent after 24 hours)
as shown in Figure 52.

Figure 52. The effect of slurry level variation and its holding time on the lateral deformations associated with
trenching (Poh and Wong, 1998)

CIRIA report 580 (Gaba et al., 2003) summarizes horizontal and vertical wall movements due to
installation of diaphragm walls and bored pile walls in stiff clays as shown in Figure 53. While
Clough & ORourke (1990) predicted that the maximum settlement could reach 0.15% of the
trench depth, Gaba et al. (2003) found out that the maximum settlement is 0.04-0.05% of the
trench depth. The maximum lateral deformation is about 0.04 to 0.08% of the maximum trench
depth.

46

.
Figure 53. Vertical deformations due to diaphragm wall installation (Gaba et al. 2003)

Gourvenec and Powrie (1999) investigated the influence of panel length and construction
sequence on the lateral deformations of a diaphragm wall. Figure 54 shows the lateral
displacements, normalized with respect to the maximum lateral displacement corresponding to
the plane strain case, versus depth, normalized with respect to the wall depth, for different panel
lengths. It can be seen in the figure that the maximum lateral displacements for panel lengths of
2.5, 3.75, 5 and 7.5 m are approximately 90, 75, 65 and 40 percent of the displacements obtained
for plane strains conditions ( L = ), respectively.

Figure 54. Influence of Panel Length on Lateral Displacements (Gourvenec & Powrie, 1999).

47

Abdel-Rahman & El-Sayed (2002a & 2002b), El-Sayed & Abdel Rahman (2002) and AbdelRahman & El-Sayed (2009) studied a case history of diaphragm wall trenching and pit
excavation in Nile alluviums of the Greater Cairo. They augmented the field data with 2D and
3D finite elements. They have concluded the following:
Using 3D finite elements, the maximum settlement due to trenching was estimated to be
about 0.048% of the maximum height of the trench for deep foundations and 0.03% of
the maximum height of the trench for shallow foundations.
Using 2D finite element, the maximum trenching settlement is estimated as 0.045% of
the trench depth for both shallow and deep foundations as shown in Figure 55.
The maximum lateral deformation due to trenching is about 0.077% of the trench depth
for piles and 0.047 % of the trench depth for the case of shallow foundations.
The maximum settlement in both cases was estimated to be 61% of the lateral
displacement as shown in Figure 56.

Figure 55. The settlement envelopes for shallow and deep foundation due to trenching to a depth of 21m in the Nile
Alluviums in the Greater Cairo (Abdel-Rahman & El-Sayed, 2009).

48

Figure 56. The relationship between the lateral deformations and the maximum settlement due to trenching in the
Nile Alluviums in the Greater Cairo (El-Sayed & Abdel-Rahman, 2002).

2.11.

Building Stiffness and Weight

There is a mutual influence between a building located nearby deep excavations and the induced
deformations. Both stiffness and weight of the building affect the final shape of the
deformations. The building stiffness tends to flatten the deformation across the building; while
the building weight increases the deformation especially in location close to the deep excavation.
Potts & Addenbrooke (1997) found that deformation induced by tunneling building deformation
is an interactive problem that can be solved using two relative stiffness ratios: a ratio expressing
for the bending stiffness of the building and the other is for the axial stiffness of the building.
Goh (2010) and Goh & Mair (2011) modified the relative stiffness ratios that were initially
proposed by Potts & Addenbrooke (1997) for tunneling to be utilized for deep excavations. They
introduced design charts that allow considering the effect of building stiffness on the induced
deformations. Mair (2011) showed that field data confirmed the trend according to Goh (2010)
and Goh & Mair (2011). The approach of relative stiffness is elaborated later in Section 10.3.3 in
this state-of-the-art.
Burd et al. (2000) studied the deformation associated with tunneling and found the following
differences between the building influence in sagging and hogging ground deformation modes:
1. The stiffness of the building reduces the differential settlement in sagging deformation.
They suggest that the ground provides a certain amount of lateral restraint when the
building is subjected to sagging deformation similar to the conclusions of Burland &
Wroth (1974 & 1975).
2. In hogging mode, such a restraint is not provided and the structure behaves more flexibly
leading to higher degrees of damage than in sagging. Burd et al. (2000) related this
behavior to the imposition of building weight which alters the settlement behavior
compared to the greenfield deformations.
49

Elshafie (2008) performed centrifuge tests on model buildings subject to excavation-induced


ground displacements as shown in Figure 57. Buildings with two foundation types: raft and
isolated footings, were introduced near the deep excavation. Simulated buildings were made
from micro-concrete with variable stiffness, weights and interface roughness. He noted the
following:
1. Horizontal displacements are clearly influenced by a smooth interface, leaving the green
field soil displacements intact, even for higher axial stiffness. Rough interfaces restrained
the horizontal movements of the building.
2. The roughness of the buildings-soil significantly affects by the axial stiffness of the
blocks. Increasing the roughness increases the axial stiffness of the building. The effect
of the interface between the soil and the building is seen especially for buildings with low
bending stiffness. Stiff buildings tend to tilt regardless of the interface roughness.
3. The effect of building weight (up to 40 kPa) was small (maximum about 10% increase in
deflection ratio) as long as a high factor of stability (> 1.4) of the wall was maintained.
This conclusion is in line with the findings of Franzius et al. (2004) for the deformations
induced by tunneling.
4. Buildings with individual spread footings experience large differential deformations,
because footings outside the zone of influence do not follow the influenced part of the
building. This results in significant distortions and tensile strains concentrating at the
weak parts of the buildings.

Figure 57. Buildings and interfaces used in centrifuge tests (Elshafie, 2008)

Son & Cording (2005) showed that cracks initially develop around opening areas (e.g., doors and
windows) and propagate further with increasing ground movement. The formation of cracks
decreases building stiffness and increases the tendency for a building to conform more closely to
the greenfield movements.
50

2.12.

Wall-Soil Interface

Yu & Gang (2008) studied the effect of the friction along the wall-sand interface on the wall
deflections and ground surface settlements using parametric finite element analyses. Figures 58,
59 & 60 show that the variation of the maximum lateral deformation, maximum settlement and
their ratio with the variation of the interface friction angle. For smooth wall, the deformations
(viz., maximum wall deflection is 0.2% and maximum settlement is 0.27% of the excavation
depth) are larger than the rough wall (viz., maximum wall deflection is 0.1-0.125% and
maximum settlement is 0.04-0.1% of the excavation depth). The ratio between the maximum
settlement and the maximum wall deflection also decreases from 1.35 in smooth walls to 0.50.75 in rough walls.

Figure 58. Wall deflections variation with the variation of the wall-sand friction (Yu & Gang, 2008)

Figure 59. Maximum settlement versus the wall-sand interface friction angle (Yu & Gang, 2008)

51

Figure 60. The variation of the ratio between the maximum settlement to the maximum wall deflection versus the
wall-sand interface friction angle (Yu & Gang, 2008)

2.13.

Workmanship

Workmanship can be considered as the human and/or experience factors which plays an
important role in the success or failure of a certain project. It was initially introduced by Peck
(1969a) as one of the main controlling factor of the ground and wall movements in deep
excavation projects. This factor has never been defined thoroughly in literature; despite that, its
impact is important and well-acknowledged in the final outcomes of geotechnical projects. In
fact, Deep excavations are very special projects. They need the Designer and the Contractor to be
well-acquainted with the technical and constructional aspects of the site as well as the structural
nature of the adjacent buildings. The following important findings demonstrate the importance of
the workmanship factor:
1. About 10% of the problematic deep excavation cases in The Netherlands for the projects
carried out between 2007-2012, poor workmanship was a main factor in the encountered
problems; it was related to the installation of diaphragm wall, piled walls or anchors
(Korff & Tol, 2012).
2. Sowers (1993) demonstrated the importance of human factors in geotechnical
engineering. He studied of more than 500 well documented foundation failures showed
that the majority (88%) of failures were due to human shortcomings; only 12% of the
failures were due to lack of technology.
Methods to enhance the workmanship include documentation of the performance and
encountered problems in deep excavation projects and transfer the gained knowledge to other
contractors and other personnel. An example is Pecks (1969a) note that the methods of placing
lagging in soldier pile and lagging walls affects the maximum settlement. Four different
methods, shown in Figure 61, were considered. Peck noted that the settlements experienced next
to the excavation with method (b) were found to be three times larger than with method (a).
Methods (c) & (d) are better than methods (a) & (b) because the contact between the soil and the
lagging is better.
52

Figure 61. Different methods of placing lagging (Peck, 1969a)

53

Translating the findings of our research into simple concepts and procedures for
the guidance of the practicing engineer is, in my opinion, a duty and worthy
activity of our profession.
Ralph Peck

54

3.

EMPIRICAL AND SEMI- EMPIRICAL METHODS

The magnitude and distribution of the deformation induced by deep excavations are related to
many factors such as soil and groundwater condition, construction quality, excavation geometry,
excavation sequences, duration of excavation, surcharge condition, existence of adjacent
buildings, method of retaining wall construction, penetration depth, wall stiffness, type and
installation of lateral support, spacing and stiffness of struts.
Methods to predict the deformation derived purely from theoretical basis would be very complex
to be obtained to the complex interaction of the aforementioned factors. Therefore, many
predictive methods were envisaged based on the analysis of the databases of the previous wellinstrumented case histories. The resulting relationships may be completely based on
measurements (empirical relationships) or based on both measurements and numerical/analytical
analysis (semi-empirical relationships).
Peck (1969a) was the first to utilize this approach to predict the deformations and stresses
associated with deep excavation and tunneling using the results of monitoring program. He
presented simple graphs for assessment of the settlement induced-by excavation in terms of the
soil type, the depth of excavation, the basal heave stability and the workmanship. Since then,
many empirical and semi-empirical methods were envisaged to determine the ground
movements. Most of them do not consider the effect of buildings and hence they predict only the
greenfield deformations. Empirical and semi-empirical methods form the first important step in
predicting the effect of excavations on adjacent buildings.
Some of empirical and semi-empirical methods include all construction activities, whereas others
only describe a specific aspect, so that the different contributions have to be added to a total.
Clough and ORourke (1990) categorized the deformations induced by deep excavations into two
types:
1. Deformations caused by auxiliary construction activities (such as dewatering and wall
installation).
2. Deformations due to the excavation and support process,
As the induced deformations are mostly related to the shape of the wall stiffness/ type, the soil
type, the relative flexibility or stiffness of the system (including the wall and its supports)
compared to the soil stiffness, the safety against basal heave in clays and the duration of the
excavation, an ideal empirical relationships.
Table 1 presents some of the commonly acknowledged empirical and semi-empirical
International relationships for displacements due to the effects of deep excavations. Table 2
represents the finding of the National studies.

55

No.
1

Table 1: Summary of some of the acknowledged International empirical and semi-empirical studies
Reference(s)
Main findings
Peck (1969a)
The relationships for surficial settlements are classified into three zones, I, II and III, depending on the
workmanship and type of soil as summarized in the below figure :

Clough
(1975)

Recommended settlement envelopes Compromise all building activities, the stability of the excavation and
consolidation during construction.

Recommended settlement envelopes are conservative. The use of more contemporary technology such as
diaphragm wall, limited the maximum settlements to lower values than those predicted by Pecks method.

The maximum settlement behind the wall tends to be less 0.3% of the excavation depth except for limited cases
where the settlement more than that may be attributed to inferior workmanship, consolidation or due to lowering
of groundwater.

56

Remarks
This study was the first
paper
to
provide
a
comprehensive review of
the measured surficial
settlements associated with
deep excavations.

The considered excavation


supports are: Strutted sheet
pile walls; strutted soldier
piles with lagging (Berliner
walls)

Different soil types (sands


& clays) were considered.

Apparent
pressure
diagrams
were
also
provided for the design of
the supporting wall.

It was noted that the


ground next to deep
excavations
in
stiff
overconsolidated
clays
might rise instead of settle.

Mainly tied-back
were considered

Different soil types were


considered.

walls

No.
3

Reference(s)
ORourke et
al. (1976)

Main findings
The maximum surface settlements observed in deep excavations in Washington, which is characterized by stiff
clays Interbedded with dense sand, were equal to or less than 0.3% of the excavation depth near the edge of the
cut and 0.05% at a distance equal to 1.5 times the excavation depth.

For Chicago soft clay, the maximum settlements were in agreement with Pecks (1969a) recommendations.
Three zones of ground displacement can be distinguished and related to the characteristics of construction as
summarized below:

57

Remarks
The considered excavation
supports was strutted/raked
soldier piles with lagging
(Berliner walls).

The
considered
soil
formations are dense sand
Interbedded with stiff clay
of Washington and soft
clay in Chicago.

They recommended three


zones are similar to Peck
(1969a) with the exception
that the widths of the
settlement
zones
are
notably shorter.

No.
4

Reference(s)
Goldberg et
al. (1976)

Main findings
The maximum surficial settlement behind the wall was found to lie between 0.5-2.0 times the maximum
horizontal wall deflection.
The maximum settlements behind the wall are generally about 0.171% of the excavation depth in sands,
gravels and very stiff to hard clays and 1.22% for soft to stiff clays. The settlement envelopes are:

The maximum lateral displacement can be estimated from the below graph:

58

Remarks
Different supporting walls
were considered.

Different soil types (sands


& clays) were considered.

Recommended envelopes
Compromise all building
activities.

No.

Reference(s)

Clough et al.
(1989);
Clough and
ORourke
(1990)

Main findings

Remarks

Where Ew is the youngs modulus of the wall, Iw is the moment of inertia of the wall per linear meter, h is a
representative unsupported length of the wall such as the distance between struts
In stiff clays/residual soils/sands, the maximum horizontal wall deflection tends to average about 0.2% of the
excavation depth and the maximum soil settlement behind the wall is generally less than 0.30% of the excavation
depth. The average value of the maximum settlement can be considered as 0.15% of the excavation depth as
shown below

Different excavation supports


were considered.
Different soil types (sands &
clays) were considered.
Data are also derived from
case histories, aided by a
series of nonlinear finite
element analyses.
The recommended settlement
envelopes
pertain
to
settlements by excavation and
bracing
stages
of
construction.
Movements
associated
with
other
activities (such as dewatering
or wall installation) must be
estimated separately.

In soft clay, the wall stiffness and support spacing affect greatly the wall deflection as observed by Peck (1969).

59

No.

Reference(s)

Main findings
Clough and ORourke (1990) utilized Clough et al. (1989) method of predicting the maximum settlement in
clays:

Dimensionless settlement profiles were developed for sand, stiff to very hard clays, and soft to medium clays as
follows:

Dimensionless lateral deformation profiles were developed for clays as follows:

60

Remarks

No.

Reference(s)

Bentler
(1998)

Long (2001)

Moormann &
Moormann &
Moormann
(2002) ; and

Main findings

The maximum horizontal wall deflection for excavations in sand or hard clays is 0.19% H and for soft to stiff
clays 0.45% H, where H is the depth of excavation.

The average of the maximum settlement is 0.22% H in sands/hard clays and 0.55%H in soft-stiff clays.
In stiff clays, the average of the maximum lateral deformation is ranged between 0.16 - 0.19% of the depth of
the excavation and the average maximum settlement behind the wall is ranged between 0.12 - 0.2% of the
depth of the excavation.

For soft clay with low factor of safety against base stability, large lateral movements of the wall (up to 3.2%)
were observed. When there is a high factor of safety against basal heave, the average maximum wall deflection
is 0.39% of the depth of excavation with an average maximum vertical settlement of 0.50% of the depth of
excavation.

In stiff clay, the deformations are independent of the stiffness of the wall and the support as well as the kind of
support. The stiffness only affects the deformation significantly when dealing with deep excavation in soft
clays with a low factor of safety against base heave.
Displacements are generally smaller than Pecks (1969a) but with some large displacements in cases with soft
soils and low factor of safety against basal heave as shown below:

61

Remarks

This study is based on a


database comprising 41
deep excavations presented
between 1989 and 1998.
Different
excavation
supports were considered.

The database comprises


mostly data for case
histories in clays.

296 case histories were


analyzed.

Different
excavation
supports were considered

The

database

comprises

No.

Reference(s)
Moormann
(2004)

Main findings

Remarks
mostly data for
histories in clays.

Soil displacements tend to become zero at 2H, which is similar to results presented by Peck (1969).

The maximum wall deflection averages to 0.87% H for soft clays, with a rather large spread around it.
Maximum vertical soil displacements tend to be 0.5-2 times the horizontal deflection, with an average for soft
clays of 1.1%H and occur within a distance smaller than 0.5H.

62

case

530 case histories were


analyzed.

No.
9

Reference(s)
Boone
(2003);
Boone
&
Westland
(2005)

Main findings
The maximum horizontal wall movement *h,max is determined in according Clough et al. (1989) and modified
using s factors as follows:

63

Remarks
This method is based on
Boone
(2003)
which
augmented the works of
Manna & Clough (1981)
and Clough et al. (1989).

No.
10

Reference(s)
ZapataMedina
(2007)

Main findings
An new flexibility factor was presented as defined below:

Where
R = relative stiffness ratio,
Es = Secant Youngs modulus at the 50% stress level,
E = Youngs modulus of the wall,
I = moment of inertia per unit length of the wall,
SH = average horizontal support spacing,
SV = average vertical support spacing,
H = height of the wall,
He = excavation depth,
s = average unit weight of the soil,
su = undrained shear strength.

The maximum lateral deformation H(max) is given by the following graph as a function of the relative stiffness
R and the factor of safety against bottom heave (FS).

The distribution of the lateral displacement of the wall is given by the following graphs:

64

Remarks
A database comprised 30
case histories had been
selected to cover soft to
stiff clays.
The database was back
analyzed
using
threedimensional finite element
to for a base for parametric
analysis.

No.

Reference(s)

Main findings

The maximum settlement is given by the following equation:

The distribution of the maximum settlement is given by the following graphs:

65

Remarks

No.

Reference(s)

Main findings

66

Remarks

No.

Reference(s)

11

Hsiao (2007)

Main findings

Use the following functional form is adopted for transformation of the depth of excavation He, ln(EI/w h4avge),
half width of the excavation B/2, the ratio between the undrained strength and the effective vertical stress
su/v, the ratio between the initial modulus and the effective vertical stress E i/v:

Remarks

A database comprised 33
case histories in soft to
medium
clays
(from
Taipei, Singapore, Oslo,
Tokyo, and Chicago) had
been selected to form the
main database of the
analysis.

The database was back


analyzed using nonlinear
finite element to extend
the database by generated
data of 144 cases.

Where a1, a2 & a3 are given in the following table:

Determine the maximum lateral deformation of the wall without the effect of hard stratum (hm) using the

67

No.

Reference(s)

Main findings
following equation:

The coefficients for the above equation are as follows: b0 = 13.41973, b1 = 0.49351, b2 = 0.09872, b3 =
0.06025, b4 = 0.23766, b5 = 0.15406, b6 = 0.00093, b7 = 0.00285 & b8 = 0.00198.

Determine the effect of hard stratum (if any) using the following equations:

Where hm is the maximum wall deflection assuming very deep hard stratum, and hm,m is the modified
maximum wall deflection that accounts for the presence of the hard stratum.

Estimate the deformation ratio R which is defined as the ratio between the maximum settlement and the
maximum lateral deformation (i.e., vm /hm) using the below equation:

68

Remarks

No.

Reference(s)

Main findings
c0 = 4.55622, c1= 3.40151, c2 = 7.37697, c3 = 4.99407, c4 = 7.14106, c5 = 4.60055, c6 =8.74863,
0.38092 & c8 = 10.58958.

12

Konstantakos
(2008)

Remarks
c7 =

The surface settlement v at distance d from the wall is by:

A large portion of the strutted walls (either constructed bottom up or top down) deflected less than 0.2%H
where H is the depth of excavation.

The average deflection is 0.8%H for the braced walls, including some walls with large deflections due to low
factor of safety against basal heave.

Settlements behind the wall were given for half of the number of cases and averaged about 0.2-0.4%H for
braced and top-down constructions.

A relationship is found with the basal stability factor (BS), giving a sharp distinct between BS < 1.8 and BS >
1.8. At BS<1.8 the deflection of the wall was much higher than 0.2%H.

69

This study is based on a


database of 39 deep
excavation projects from
U.S.A.

Most
cases
involve
diaphragm
walls
and
excavation depths range
from 6 to 31 m in rather
soft soils with high ground
water tables.

No.
1

Table 2: Summary of some of the acknowledged National empirical and semi-empirical studies
Reference(s)
Main findings
El-Nahhas,
et The maximum settlement is about 0.15%H. Where H is the depth of excavation. The settlement
al., 1988, 1989,
trough is shown below:
1990,
1994;
Shalaby, 1989;
Abdel-Rahman,
1993; and ElNahhas 2006

Abel-Salam,
1995; Ahmed &
Abel-Salam,
1996;
Assaf,
1997;
ElNahhas et al.,
1998; Mostafa,
1998;
Fayed,
2002; Ahmed et
al., 2005 and ElNahhas, 2006

The width of the settlement trough is about 1.5 the excavation depth.

The maximum lateral wall deformation is about 0.2%-0.22% of the maximum excavation depth.

The ratio between maximum settlement and the maximum lateral wall deformation is about 0.75

The maximum settlement 0.10% of the maximum excavation depth.

The maxim lateral deflection is about 0.16% of the maximum excavation depth.

70

Remarks
The databases comprised the
measurements made during the
execution of the Greater Cairo
Metro - Line 1 at test sections that
were instrumented.

The soil formation is Greater Cairo


alluvial with shallow water depth.

The
entire
length
of
the
underground portion of line 1 of
Cairo Metro (about 4.5-km long)
was excavated using precast
concrete diaphragm walls.

A grout plug was utilized; therefore


the settlement does not include the
effect of substantial groundwater
lowering.

The databases comprised the


measurements made during the
execution of the stations for Greater
Cairo Metro - Line 2 at test sections
that were instrumented.

The soil formation is Greater Cairo


alluvial with shallow water depth.

A grout plug was utilized; therefore


the settlement does not include the
effect of substantial groundwater
lowering.

Abdel Rahman
&
El-Sayed
(2002a, 2002b
& 2003), ElSayed & Abdel
Rahman (2002),
and
Abdel
Rahman & El
Sayed (2009)

The settlement envelope (Strench) due to trenching at a distance x from the trench/wall is given by the
following formula
trenching 2d x
S trenching S max

2d

Settlement observations were made


for 5 buildings (3 on deep
foundations and 2 on shallow
foundations) located near deep
excavation supported by diaphragm
wall having struts and tie-backs.

The depth of the diaphragm wall


panels was deeper than the pile tips.

The settlement induced by the


excavation was observed for the
buildings on the shallow and deep
foundations during trenching for
installation of the diaphragm wall
and pit excavation.

Minor dewatering was carried out


as a grout plug was performed to
reduce the seepage to the
foundations.

The soil formation is typical Nile


alluvium with shallow groundwater
table.

The observations were amended by


2D and 3D back-analyses of the
excavation.

Where d is the depth of the excavation and Smaxtrenching is the maximum settlement associated with
trenching which equal to 0.045% for both shallow and deep foundations.
The settlement of due to pit excavation (Spit excav) is given by

x2
pit excav

S pit excav S max


exp
2
2( K H )
Where,

pit excav
S pit excav is the settlement at a distance x from the trench, S max
is the maximum

settlement at the wall location, K is a dimensionless factor and H is the final depth of pit
excavation.

Parameters of the equation describing pit excavation were determined using best fitting analysis on
the measured values. The results of curve fitting are summarized below:
Foundation Type

pit exc .
Smax

Trough width

Pile foundations
Shallow foundations

0.03% H
0.11% H

1.25
0.75

3.8 H
2.2 H

Total settlement S for the trenching and pit excavation is given by the following:
S = Strenching + Spit excav

The extent of the settlement troughs were found to reach up to a distance equivalent to 3.5 of the
depth of excavation in alluvial soils for both shallow and deep foundations.

Most of the settlements in buildings on pile foundations occurred with during the trenching stage.
This was also noted by Korff (2012). On the contrary, most of the settlement pile founded buildings
is attributed to pit excavation.

71

Theory and calculation are not substitute for judgment, but are the basis for
sounder judgment
Ralph Peck

72

4.

NUMERICAL MODELING

The influence of the individual factors affecting deformations induced by deep excavation cannot
be directly extracted from an empirical database due to the limited number of excavations in
similar soil and construction conditions. Abdel-Rahman (1993) and El-Nahhas et al. (1994)
noted that that empirical and semi-empirical method cannot recognize the effect of the
construction sequence on the deformations induced by deep excavation and the stresses acting on
the supporting system. Hence, observational and numerical approaches are frequently integrated
together to provide a reliable estimate of the ground deformations due to deep excavations. Many
of the semi-empirical methods that were previously presented (e.g., Mana & Clough, 1981;
Clough et al., 1989; Zapata-Medina, 2007) utilized parametric numerical parametric analyses to
augment the field data to reach reliable formulations of the deformations associated with deep
excavations.
Hashash & Whittle (1996) also noted that most of the empirical methods were envisaged based
on data obtained from excavations less than 15 m deep, with relatively flexible lateral earth
support systems (i.e., sheet pile and soldier pile walls); such limitations in the available
monitoring databases may impose significant uncertainties in extrapolating the empirical
observations to new situations of interests such as very deep excavations supported by stiff
structural systems such as thick diaphragm walls.
As such, numerical methods are currently gaining momentum to be increasingly utilized in
geotechnical problems such as deep excavations. Since their inceptions in the late 1960s and
early 1970s, numerical methods such as the finite elements, the finite difference and the
boundary elements have greatly improved. Nowadays, there are numerous realistic stress-strain
models, tools to ease the creation of the geometric model, and tools to provide powerful and
useful graphical output for quick interpretation and presentation of results. Numerical
simulations have become very common for the analysis of excavations in urban environments.
Most of the research in this aspect is directed towards the Finite Element Method (FEM). Many
advanced finite elements software packages become commercially available (e.g. ABACUS,
PLAXIS, etc.); hence, complex non-linear analysis using large meshes progressively become a
routine matter. Many nonlinear finite elements analyses show close agreement with field
measurements (Clough & Mana, 1976; El-Nahhas et al., 1990; Abdel-Rahman, 1993; Hashash
and Whittle, 1996; Assaf, 1997; Bentler, 1998; El-Nahhas and Morsy, 2002; Kung et al., 2007b
& 2009; and others).
In the following sections some key aspects related to the most common numerical methods (viz.,
beam-column on elastic Winkler springs; & finite elements) are highlighted.

4.1.

Beam-Column on Elastic Winkler Springs

Design and analysis of foundation systems through the use of beam on elastic foundation
analysis techniques has been available to engineers for many years. Therefore, it has also been a
common numerical model wall deformations supporting deep excavation. The wall is considered
as a beam-column and the soil around the wall as equivalent springs based on the classical
Winkler (1867) model as shown in Figure 62 (Delattre, 2001).
73

Figure 62. Subgrade reaction model for analysis of walls supporting deep excavations (Delattre, 2001).

The construction sequence for the wall is modeled by dividing a series of unloading and
reloading sequences into several stages on the basis of the excavation to the proper excavation
level. The struts, rakers and tie-backs are introduced into the analysis using spring with higher
stiffness than the soil. The prestressing in the struts or tie-backs is introduced as forces at their
lock-off stages (Hwang et al., 1996; Strom & Ebeling, 2001 and others).
In the Winkler analysis of walls supporting deep excavations, springs can be taken as either
linear or nonlinear with their response based on curves that relate soil resistance p (also called R
in some literature), to wall displacement y. A plot of the pressure p on the wall versus wall lateral
movement y is designated the p-y curve. In general p-y curves (aka, R-y curves) are nonlinear;
however, they can be approximated as ideal linear elastic- fully plastic systems. These curves can
be generated for various elevations along the wall height.
For the idealized Winkler spring the relationship between the horizontal reaction p and the
displacement y may be considered a linear relationship. The ratio p/y in the elastic zone is called
the spring stiffness and denoted by kh. The stiffness coefficient kh may increase linearly with
depth (approximating the behavior of cohesionless soils, normally consolidated silts, and
normally consolidated clays) or remain constant with depth (representing the behavior of overconsolidated cohesive soils). Terzaghi (1955), Haliburton (1968 & 1987), Pfister et al. (1982),
Dawkins (1994a) and others recommended values for the spring stiffness (kh). Figure 63 shows
the recommendation of Pfister et al. (1982) for calculations of the subgrade value (kh) as a
function of the soil friction angle and cohesion.
74

Figure 63. Horizontal subgrade moduli, kh (after Pfister et al., 1982)

75

Turabi and Balla (1968) and Haliburton (1968) introduced the nonlinear presentation of the
subgrade reaction spring so that the stress will vary from active pressure if the wall deflected
sufficiently away from the wall to passive earth pressure if the wall deflected adequately towards
the soil. Dawkins (1994b) suggested using two values for kh for the active and passive earth
pressures as shown in Figure 64.

Figure 64. Idealized elastoplastic earth response-deflection curve with two subgrade reactions (Dawkins 1994b)

Boudier et al. (1970), Fages & Bouyat (1971a & 1971b), Rossignol & Genin (1973), Balay et al.
(1982), Balay (1985), Dawkins (1994a & 1994b) and others developed software packages for the
design of retaining walls using the subgrade reaction method. Pearlman & Wolosick (1990)
modeled soldier piles and diaphragm wall using beam on elastic foundation analysis to study the
soil-structure interaction. An advantage of the beam and spring approach over limit analysis is its
ability to account for structure flexibility and soil stiffness. Thus, the effects of stress
redistribution in soil because of differential structural deflections are accommodated.
Weatherby et al. (1998) suggested that the displacements required to generate active or passive

conditions are constant with depth. This gives rise to the Reference Deflection Method for
developing elastoplastic soil spring. The Reference Deflections are those lateral deformations
necessarily needed to mobilize active and passive soil resistance depending on the soil type.
Figures 65 & 66 show the Reference Deflection curves for sands and clays.

76

Figure 65. Elastoplastic sand subgrade diagram using Reference Deflection Method (Weatherby et al., 1998)

Figure 66. Elastoplastic clay subgrade diagram using Reference Deflection Method (Weatherby et al., 1998)

77

Prestressed anchor spring resistance-displacement curve (T-y curve), Figure 67, is a spring
model similar to the Winkler soil model. The pre-yield response of the anchor spring is based on
the yield displacement, tendon yield strength, and tendon elastic stiffness. The yield
displacement is given by (Strom & Ebeling, 2001):
(1)
( f y 0.60 f y ) Lu
yy
cos
Es
The tendon yield strength (Ty) is

Ty AS f y cos
And the tendon elastic stiffness (kT) is
AE
kT S S cos
Lu
where fy
= yield stress of strand
Lu
= effective elastic length of the anchor (unbonded length + bonded length)

= anchor inclination with horizontal


ES
= Youngs modulus for anchor tendon
AS
= Area of anchor tendon

(2)
(3)

The load region beyond yield can assumed flat (plastic) or interpolated between the yield point
and ultimate capacity with fu replacing the first occurrence of fy in Equations 1 & 2 to generate yu
and Tu .

Figure 67. Ground anchor T-y curve (Strom and Ebeling, 2001)

78

Modeling construction stages/increments with beam-column-spring models is accomplished by


shifting the R-y curve with each increment, or modeling the soil as a static load if it reached the
active or passive limit state. The shifted R-y method adjusts the soil response curves in the new
increment to account for the effect of staging as shown in figure 68. The deflection y is
considered as the incremental deformation that should be summed to have the total lateral
deformation of the wall. The same procedure is applied on the load-deformation curve of the
anchor or strut (Weatherby et al, 1998).

Figure 68. Shifted R-y method to model construction stages (Weatherby et al, 1998)

79

Once the lateral deformation of the wall is determined using the beam-column on Winkler
springs model is determined, the settlement trough can also be determined by the methods
presented in Section 2.7 in this state-of-the-art after identify the amounts of wall deformations
that can be attributed to the cantilever and bulging lateral deformation modes.
Strom & Ebeling (2001) summarized and addressed the difficulties in obtaining reasonable
results from a Winkler spring analysis as the following aspects are not well-presented in such
analysis:
1. The load deformation characteristics of the soil are not linear and may not be suitably
represented by ideal elastoplastic behavior.
2. The soil stiffness varies with respect to confining pressure and zone of influence.
3. The soil stiffness changes with submergence.
4. The ultimate resistance of the soil is dependent on different failure mechanisms which
vary depending on whether the soil is near the surface or at some depth below the
surface.
5. The behavior of discrete wall systems (soldier beam systems) is different from
continuous wall systems because the earth pressure distribution behind the wall is
different (zone of influence is different) and because soil has a tendency to arch between
the structural elements of discrete wall systems.
The aforementioned aspects call for the use of finite element analysis of the wall in lieu of the
beam-column-spring models.

80

4.2.

The Finite Elements

Generally, the Finite Element Method (FEM) is a numerical technique for the solving differential
equations governing boundary value problems in science and engineering. The domain of the
problem is divided into smaller parts (i.e., finite elements) and the connected set of the finite
elements representing the whole domain is called the finite element mesh as shown in Figure 69.
The behavior of each element is described by its geometry, and constitutive relationships. The
literature related to the finite element method is vast (Zienkiewicz & Morice, 1971; Desai, 1972;
Naylor et al., 1981; Potts & Zdravkovi, 1999 & 2001; and others); however, the basic procedure
for the different finite element analyses in geotechnical engineering broadly remains the same for
different problems and can commonly be summarized as follows:
1. Select element configuration and discretize the domain.
2. Define strain-displacement and stress-strain relations (constitutive laws).
3. Derive element equilibrium equations (e.g., by applying the principle of virtual work).
4. Assemble global or system equilibrium equations and apply boundary conditions.
5. Solve for primary unknowns (displacements).
6. Solve for secondary unknowns (strains and stresses).
7. Display primary and secondary unknowns.

Figure 69. A schematic showing the discretization of deep excavation problem into a finite element mesh

The first systematic Soil-Structure-Interaction (SSI) finite element analyses of soil retained by
walls were presented by Chang (1969), Clough & Duncan (1969, 1971) and Duncan & Clough
(1971), Clough et al. (1972) and Christian & Wong (1973). The first SSI analyses for deep
excavation utilized the hyperbolic constitutive relationship developed by Duncan & Chang
(1970) to model the behavior of the soil and the joint element developed by Goodman et al
(1968) to model the interface.
Since then, the FEM has numerously been employed to model SSI problems, particularly, the
deformation analysis caused by deep excavation and tunneling to estimate the excavationinduced ground movements and the support loads. It gained widespread acceptance due to its
capability to model complex construction sequences, and to incorporate detailed site-specific
properties of the structural support system and surrounding soils (Finno & Harahap, 1991; Ng,
1992; Whittle et al., 1993; Abdel-Rahman, 1993; Hashash & Whittle, 1996 & 2002; Ng & Yan,
1999; El-Nahhas and Morsy, 2002; El-Sayed & Abdel-Rahman, 2002; Kung et al., 2007b &
2009; and others).
81

Finite element analysis of deep excavations and nearby buildings presents special challenges for
the following reasons (Woods, 2003):
1. The geometry changes incrementally as shown in Figure 70;
2. Short and long term solutions are required for some geomaterials such as clays;
3. Potentially discontinuous deformations at the interface of the wall and the soil; &
4. Stiffness contrast is usually high between the supporting wall and the building, and the
surrounding soils;

Figure 70. Typical excavation sequence in deep excavation supported by struts (Hashash & Whittle, 1996)

The following important aspects addressing the above key points to realistically simulate deep
excavations are highlighted in the following sections:
1. Constitutive soil modeling;
2. Simulation of excavation and sequential construction;
3. Interface modeling;
4. Two- and three-dimensional analysis; &
5. Modeling of the effect of structures located near the excavation.

82

4.2.1.

Constitutive soil modeling

For accurate numerical analysis of deep excavations, a representative constitutive formulation of


the soil must be utilized. Figure 71 schematically illustrates the effect of discrepancy in the
settlement distribution on the evaluation of damage potential of buildings between actual
observations and finite element predictions. Inaccurate finite element predictions resulting from
utilization of a constitutive model that ignores the effect of the strain level on the soil modulus
may underestimate the angular distortion of the building and consequently underestimate the
expected damage level of building caused by the excavation-induced settlement.
Notwithstanding that, simple constitutive models (such as Mohr-Coulomb elastoplastic model or
Duncan-Chang hyperbolic model) may be utilized to obtain realistic predictions of the wall
lateral deformation despite their poor prediction of the settlements (Jardine et al., 1986; Kung,
2010).

Figure 71. Inaccurate evaluation of the differential settlement affecting buildings due to the utilizing of
unrepresentative constitutive model (Kung, 2010)

Generally, the stress-strain behavior of soil is highly nonlinear which has a dominant influence
on the form and scale of the displacement distribution. Figure 72 displays the typical variation of
soil stiffness with the strain from less than 10-6 to 1. The degradation of soil stiffness especially
at the small stain level is significant. As such, it is believed that the capability of soil model in
describing the stress-strain-strength characteristics of soil at a wide range of strain (e.g., 10-5 to
10-2 which represent the expected range of strains associated with deep excavation) would
significantly affect the accuracy in the prediction of excavation-induced ground movements
(Burland, 1989; Benz, 2007; Kung et al., 2009; El-Nahhas & El-Mossallamy, 2009; Kung, 2010;
Obrzud, 2010).

83

Figure 72. Variation of modulus with strain level (Obrzud, 2010)

4.2.2.

Simulation of excavation and construction sequence

There are many procedures to model excavation. Gutierrez et al. (2002) surveyed the approaches
of the excavation simulation. The initial simulation was given by Brown and King ( 1966),
Dunlop et al. (1968), Clough & Duncan (1969, 1971), Desai & Abel (1972), Christian & Wong
(1973), , and others utilized the Stress Reversal Algorithm (SRA) to model excavation by
reversing the traction pressures between the excavated and the unexcavated elements. For
example, the simulation of excavation shown in Figure 73 as a shaded area may be carried out
using the following steps:
1. Find the tractions or boundary stresses transmitted to the remaining soil by the soil that
will be excavated.
2. Remove the stiffness of the excavated region from the stiffness of the whole region.
3. Apply to the remaining soil the tractions or boundary stresses with magnitudes equal to
those determined but opposite in sign (i.e., direction).
4. Add the incremental displacements, strains, and stresses the condition before excavation.

84

Figure 73. Stress reversal approach (Gutierrez et al., 2002)

Chandrasekaran and King (1974) presented a procedure that does not rely on interpolation of the
stresses along the excavation boundary. The procedure is based on unbalanced loads calculated
from the product of the stiffness matrix of the unexcavated elements and the displacements from
the previous stage. The procedure is shown to yield numerically identical results to SRA. Clough
and Mana (1976) presented a numerical solution based on calculating excavation forces from
element stresses at the Gaussian points and not the element edges.
Ghaboussi and Pecknold (1984) presented a generalized formulation for the simulation of
excavation forces in the context of the Newton-Raphson techniques which is widely used in
nonlinear analysis. Brown and Booker (1986) showed the uniqueness of the solution using the
Ghaboussi & Pecknolds (1984) procedure for linearly elastic which is a requisite for a
theoretically-sound algorithm. Borja et al. (1989) extended the Ghaboussi and Pecknolds (1984)
procedure to nonlinear materials and accounted for time-varying problem domain and boundaries
such as the changes due to sequential excavations.

85

4.2.3.

Interface modeling.

Interfaces are simulated using 1-D elements (i.e., elements without thickness) that are introduced
between the structure and the soil. Interface elements allow adjacent continuum soil and structure
elements to move independently of each other. The main purpose of interface elements is to be
able to permit slippage between the soil and the structure during construction and also during
subsequent loading of the structure. Goodman et al. (1968) presented the first formulation of
interface modeling as a 1D finite element. Interface behavior can have important influence on the
excavation forces due to the shear stresses between the soils and the wall. Many studies
emphasized the importance of the interface modeling (e.g., Ebeling el., 1990, 1993 & 1997;
Ebeling & Mosher, 1996; Ebeling and Wahl, 1977).
Figure 74-a shows interface stresses near the top of the wall before the start of excavation.
Immediately following excavation, the wall bends inward, reducing normal stress and generating
downward shear stress on the wall (Figure 74-b). After tensioning, normal stresses increase and
shear stress may reduce or reverse direction (Figure 74-c). Subsequent excavation starts the cycle
over again (Figure 74-d). As demonstrated, interface elements may be subjected to very complex
stress paths; hence, they should be capable of modeling such conditions (Strom and Ebeling
2001).

Figure 74. Interface stresses during excavation and tensioning (Strom and Ebeling 2001)

86

4.2.4.

Two-dimensional versus three-dimensional analyses

Generally, plane strain two-dimensional (2D) analysis is conducted to assess wall and ground
movements in the middle of each side of the excavation. This simplifying assumption sometimes
is not consistent with actual excavation but it is considered a conservative assumption (Ou et al.,
1996). The analyses may also be based on axisymmetric assumption when the excavation is
symmetrical.
Notwithstanding these simplifications in analysis, deep excavation is actually a threedimensional problem (Zapata-Medina, 2007). 3-D finite element analysis for braced excavation
may be required to study the corner effect and the effects on building s (Bono et al., 1992; Lee et
al., 1998). Yet, full three-dimensional (3D) analyses have rarely been applied in practice and
research due to high cost of computational and time constraints.
4.2.5.

Modeling of structures affected by excavations

Any structure within the predominance of the deformation trough induced by deep excavation is
affected. On the other hand, the presence of the structure modifies the trough due to its weight
and stiffness. Generally, heavy buildings and traffic roads adjacent to the excavation site may
cause extra wall deformations due to their high loads. Conversely, increasing the building
stiffness distributes the settlements and lateral deformations at a distance greater than the trough
width and hence the deformations effects on the structures may become milder than the
greenfield condition. The numerical models that include buildings located near deep excavations
may be three-dimensional or two-dimensional. The advantage of such a 3D model is that the
building can be considered in any geometrical configuration with respect to the deep excavation.
The most common model for buildings is the deep elastic beam model which was adopted by
Burland & Wroth (1974 & 1975) and others as elaborated in Sections 10.1 in this state-of-the-art
report. The structure is represented by an elastic beam with bending stiffness (EI) and axial
stiffness (EA) representing the overall stiffness of the structure. The advantages of this method
are the small amount of computational resources required and therefore the ability to perform
extensive parametric studies. Furthermore this approach is consistent to the damage assessment
outlined by Burland (1995), Mair et al. (1996) and Son & Cording (2005).
Attewell et al. (1986) and Yoshida et al. (1994) used a method of two-stage procedures to
investigate the effect the building existence on the induced settlement trough. Firstly, the freefield surface deformations are calculated using one of the analytical methods; then the final
settlement can be calculated using the following relation:

K K d K w
s

gd

(4)

gd

in which [Ks] presents the stiffness matrix of the structure, [Kgd] presents the stiffness matrix of
the ground interface with the building, {w} is the free-field displacement vector of the ground
interface and {ds} is structure response.

87

5.

ANALYTICAL APPROACH

Generally, the estimation of ground settlement is inferred from empirical database or estimated
by finite element analyses. Bolton and his co-workers (Osman & Bolton, 2004a, 2006 & 2007;
Bolton et al., 1989, 1990a, 1990b, 2008 & 2010; Bolton, 2011) presented a simple analytical
approach called Mobilizable Strength Method (MSD) that correlates laboratory shear tests or
shear stress-shear strain relationship to the behavior of deep excavation. The same approach was
utilized by Osman & Bolton (2004b & 2004c) to predict load-deformation relationship of
shallow foundations resting on clays.
In conventional plasticity solutions for the stability of braced excavations in soils (Terzaghi
1943; Bjerrum & Eide, 1956), displacements are only permitted between the sliding blocks and
the rigid wall. Bolton and his co-workers extended this approach by considering finite
deformations fields of soils and the wall (Figure 75). The equilibrium of the deep excavation is
studied using the Principle of Virtual Work using the established finite-strain mechanism.

Figure 75. Displacement fields: (a) incremental displacement field for wide excavation; (b) incremental
displacement field for narrow excavation; (c) Plastic deformation mechanism for cantilever retaining walls in
undrained conditions. (Osman & Bolton, 2007; Lam & Bolton, 2011)

88

Bolton and coworkers approach combines statically admissible stress fields and kinematically
admissible deformation mechanisms with distributed plastic strains to provide a solution for deep
excavation problems. Displacements in the MSD method are controlled by the average soil
stiffness in the zone of deformation. Stressstrain data from an undisturbed soil sample taken at
the mid height of the retaining wall prior to excavation can be used to deduce the average shear
strength that can be mobilized at the required shear strain in MSD calculations.
The main shortcomings of this approach are the difficulty to generally envisage kinematicallyadmissible deformation fields associated with sequential excavation using minimum independent
variables. Bolton (2008) explained the factors affecting the choice of the mechanism as quoted:
Fortunately, it seems that deformation mechanisms do not much depend on soil properties, or
even on gentle variations in stratigraphy. But mechanisms must depend on major geometrical
constraints of excavations such as the width to depth ratio, the wall depth to bedrock depth ratio,
and presence of a base-grouted plug, for example. The MSD method is sufficient to assess the
likelihood of damage to neighboring services and buildings. This approach is also limited to
clay undrained condition. These shortcomings limit the use of this approach due to the easy
application of the numerical approach especial due to the wide spread of the specialized software
especially designed for geotechnical engineering and in particular the deep excavation in all soil
types.

89

6.

ARTIFICIAL NEUTRAL NETWORKS (ANNS)

The Artificial Neural Network (ANN) is a learning-based approach based on Artificial


Intelligence (AI) concepts. It is an information processing model inspired by the biological
nervous system. An ANN system is configured for a specific application, such as pattern
recognition or data classification, through a learning process. It well suited to model complex
problems where the relationship between the model variables is unknown. Typical structure of
ANNs consists of a number of Processing Elements (PEs), or nodes, that are usually arranged in
layers comprising an input layer, an output layer and one or more hidden layers as shown in
Figure 76 (Maren et al., 1990; Fausett, 1994; Shahin et al., 2001 & 2008).

Figure 76. ANN elements (Maren et al., 1990; Fausett, 1994; Shahin et al., 2001 & 2008)

The network adjusts its weights on the presentation of a training data set and uses a learning rule
to find a set of weights that will produce the input/output mapping that has the smallest possible
error. This process is called learning or training. Once the training phase of the model has
been successfully accomplished, the performance of the trained model has to be validated using
an independent testing set
As described above, ANNs learn from data examples presented to them and use these data to
adjust their weights in an attempt to capture the relationship between the model input variables
and the corresponding outputs. Consequently, ANNs do not need any prior knowledge about the
nature of the relationship between the input/output variables, which is one of the benefits that
ANNs have compared with most empirical and statistical methods.
ANNs have been applied successfully in many geotechnical engineering areas including pile
capacity prediction, settlement of foundations, soil properties and behavior, liquefaction, site

90

characterization, earth retaining structures, slope stability and the design of tunnels and
underground openings (Shahin et al., 2001 & 2008).
Goh et al. (1995) developed a neural network model to provide initial estimates of maximum
wall deflections for braced excavations in soft clay. The neural network was used to synthesize
data derived from finite element studies on braced excavations in clay. The input parameters
used in the model were the excavation width, soil thickness/excavation width ratio, wall
stiffness, height of excavation, soil undrained shear strength, undrained soil modulus/shear
strength ratio and soil unit weight. The maximum wall deflection was the only output. Using
regression analysis, the scatter of the predicted neural network deflections relative to the
deflections obtained using the finite element method were assessed. The results produced high
coefficients of correlation for the training and testing data of 0.984 and 0.967, respectively. Some
additional testing data from actual case records were also used to confirm the performance of the
trained neural network model. The agreement of the neural network predicted and measured wall
deflections were encouraging. The study intended to use the neural network model as a timesaving and user-friendly alternative to the finite element method.
Hashash et al. (2003) utilized ANNs to rationally and systematically incorporating field
observations into numerical models (Inverse Analysis). Hashash et al. (2004) utilized ANNs to
formulate a material constitutive model in finite element analysis. The developed combined
ANN-FEM model was proven to be reliable in predicting the deformations associated with deep
excavation. It was found very comparable with the advanced constitutive model MIT-E3.
Hashash et al. (2010) refined this approach by incorporation of both ANN and Genetic
Algorithm (GA) to improve the results of the proposed AI combination with FEM in the Inverse
Analysis.
Fayed (2002) utilized ANNs to model the behavior and performance of deep braced excavation
of Greater Cairo Subway Stations located on Line 1-Phase 2 and also Line 2-Phase 2. Three
subway stations were analyzed using this methodology: Rod EI-Farag Station, EL-Bohoos
Station, and EL-Dokki Station. The ANN model was formed using 58 set of data generated from
nonlinear elasto-plastic finite element analysis using the modified hyperbolic model of Morrison
(1995) with dimensions and construction sequence utilized in Rod Farag Station. In the analyses,
the soil condition was simplified as one layer with shear strength parameters (cohesion C,
friction angle ); these two parameters represent the weighted average of the shear parameters of
the soil layers in the site. The training subset constitutes 70 % of the database whereas the testing
is 20 %, and the querying is 10 % of the total data. The groundwater level was considered at a
depth of 2.50 m below the ground surface as an approximate average groundwater depth for Line
2. The ANN was trained considering the following inputs:
1. The ratio between the excavation depth and total length of the diaphragm wall;
2. The diaphragm wall thickness;
3. The angle of internal friction of the soil; &
4. The soil cohesion.
Figure 77 shows the results of the ANN model compared with the monitoring results. It was
concluded that the ANN model compared favorably with the monitoring results.

91

(a) Rod El-Farag Station

(b) El-Dokki Station

(c) El-Behoos Station

Figure 77.Comparison between measured ground deformations and ANN predictions (Fayed, 2002)

92

7.

PHYSICAL MODELING USING CENTRIFUGE

Some studies used physical modeling to model deep excavation and its effect on structures (e.g.,
Anderson et al., 1982; Seok et al., 2001). Laefer (2001) conducted large-scale tests to evaluate
the response of excavation adjacent to reinforced concrete structures. A pair of 1/10th scale
concrete frames was set in dry sand perpendicular to a nearby excavation as shown in Figure 78.
Soil was incrementally removed from in front of a three-level tied-back, continuous sheet piling
wall in a series of lifts, and the tie-backs were loaded as the excavation proceeded. Resultant soil
maximum settlement was found in the range of 0.25-0.3% of the excavation in accordance as
shown in Figure 79.

Figure 78. Test setup utilized by Laefer (2001)

Figure 79. Soil surface settlement obtained by Laefer (2001) plotted with respect to Pecks (1969a) zones

93

As soil stiffness and strength are always stress-dependent; therefore, scale effect is always a
major issue in physical modeling. Scale effect is considered in centrifugal modeling in which the
characteristics of the soil are controlled by the increased acceleration level (aka, g-level).
Centrifugal modeling can be considered suitable for modeling drained excavations in sand and
undrained excavation in clay. In sands, the effective stress can develop almost instantaneously
with the increase in g-level, dissipation of excess pore pressure occurs almost immediately.
Clays have much lower permeability and hence the consolidation process requires much longer
period for the dissipation of excess pore pressure.
To model an excavation in a centrifuge, a method of simulating the soil removal ideally has to be
carried out in-flight. Currently, the following four methods are used to model an in-flight
excavation in centrifuge:
1. Pre-set excavation (Lyndon and Schofield, 1970): Soil in the excavation area is initially
removed in 1g condition before being subjected to increasing centrifuge acceleration until
failure.
2. Draining of a heavy fluid (Bolton & Powrie, 1987; Elshafie, 2008). Soils to be excavated
are presented by a fluid of density identical to the soil. In this method, the coefficient of
lateral stress is always one. This can be suitable to heavily over-consolidated soils. For
normally consolidated soil, it is not suitable (Kimura et al., 1993). Another limitation of
this method is that the passive earth pressure still remains as 1, which is not consistent
with what known soil mechanics rules.
3. Removal of a bag of material from the excavation area (Azevedo, 1983). Soil bags are
placed at the zone to be excavated and were removed during the excavation process. This
method has advantage over the first two methods, as the modeling of stress history of the
soil model is more realistic. Since the soil used in the bags is similar to the soil model, the
coefficient of lateral stress is consistent. Nonetheless, the interaction behavior between
the interfaces of soil bags with the retaining wall would be very difficult to quantify.
4. An in-flight excavator (Kimura et al., 1993; Loh et al., 1998). A small scale robotic
excavator is developed to remove the soil in-flight in the centrifuge.
The fourth concept above was developed by Cambridge Geotechnical Group. They introduced a
2D-servo actuator to work at up to 100g in a geotechnical centrifuge, Figure 80, to be used in
plane strain analysis of clays. A narrow blade is lowered (in the analogue vertical sense) a few
millimeters into a clay bed and it scrapes the clay into a preselected disposal area. Models are
viewed on a cross-section through a Perspex window. Two digital cameras record ground
movements to be analyzed stage by stage. Models were also instrumented with displacement
transducers, pore pressure transducers, bending moment gauges on the retaining wall, and load
cells mounted on the props. This centrifuge modeling was utilized for research on the ground
movements associated with braced excavations in clay to validate the Mobilizable Strength
Method (MSD) that envisaged by M. Bolton and his coworkers (Lam, 2010; Haigh et al, 2010;
Lam et al., 2011; Bolton, 2011).

94

Figure 80. Centrifuge model package for excavations cut and propped in flight (Lam et al, 2011)

95

No theory can be considered satisfactory until it has been adequately checked by


actual observations.
Ralph Peck

96

8.

OBSERVATIONAL METHOD AND MONITORING

Precise prediction of the deformations associated with deep excavations using advanced
numerical analysis is practically unfeasible due to the highly variable nature of geomaterials.
Additionally, the empirical and semi-empirical methods for prediction the deformations induced
by excavations are envisaged based on specific case histories and hence they can be claimed to
present all cases of excavations, shoring and soil conditions. Therefore, there are always
uncertainties about the assessed deformations associated with excavation. Consequently, the
risks of distressing adjacent buildings due to the deformations induced by deep excavation
cannot be waived by any pre-construction analyses alone.
To address uncertainties in geotechnical design and the associated risks, Peck (1969b) proposed
to utilize the observational approach as an effective tool in geotechnically related projects. The
following definition of the observational method is quoted from CIRIA 185 (Nicholson at al.,
1999):The Observational Method in ground engineering is a continuous, managed, integrated,
process of design, construction control, monitoring and review that enables previously defined
modifications to be incorporated during or after construction as appropriate. All these aspects
have to be demonstrably robust. The objective is to achieve greater overall economy without
compromising safety. The objective of the observational method is to achieve greater overall
economy without compromising safety. The benefits of the observational method are
schematically shown in Figure 81.

Figure 81. Potential benefits of the OM according to CIRIA 185 (Nicholson at al., 1999)

97

Peck (1969b) described the different stages to observational approach as follows:


1. Carry out exploration sufficient to establish the general nature, pattern and properties of
the deposits related to the project.
2. Assess the most probable conditions and the most unfavorable conceivable deviations
from these conditions
3. Design by adopting an acceptable level of risk to all parties.
4. Identify risks and mitigation plans based on the most conceivable unfavorable conditions.
5. Develop reliable monitoring plan
6. Predict monitoring outcomes and establish action plan (in advance) and establish risk
triggering thresholds
7. Progress construction in clearly defined phases/stages.
8. Monitor and report promptly
9. Modify design as construction progress based on the results of the monitoring.
10. Implement appropriate changes progressively and demonstrate acceptable performance
through observational feedback.
Peck (1969) suggested that design is initiated based on most probable conditions and utilize
monitoring as a tool to update the geotechnical related as aspects as the construction proceeds.
As such, Monitoring is considered the nucleus and the most important aspects in the
observational method. It assists in managing a safe work place and helps to mitigate the risks
associated with variability in geological conditions and the inappropriate interpretation of
geotechnical data.
Nowadays, monitoring ground and support system response, recording construction activities,
and learning from measured data to extract underlying soil behavior becomes an important
component in all deep excavations and tunneling projects. Instruments often are installed to
monitor and control the performance of excavations. If the observed performance of the
excavation shows intolerable deformations changes in the design and construction procedure of
excavation is made.
It is to be noted that The observational approach is not suitable for brittle behaviors in the
structure or rapid deteriorations that do not allow sufficient warning to implement any planned
modifications such as rapid deteriorations of soils caused by groundwater or non-ductile failures
of structural members (struts/waling connections) in multi-propped basements (Patel et al.,
2007).

98

9.

INSTRUMENTATION AND MONITORING

An instrumentation program is a comprehensive approach that assures that all aspects of


instrumentation from planning and design through maintenance and rehabilitation are
commensurate with the overall purpose. To be fruitful, such monitoring programs must be
carried out for well-defined purposes, be well planned, and be supported by competent staff
through completion and implementation of results from the monitoring program. The principal
objectives of a geotechnical instrumentation plan may be generally grouped into the following
four categories:
1. Analytical, assessment of the data obtained from geotechnical instrumentation may be
utilized to verify design parameters, design assumptions and evaluate construction
techniques.
2. Prediction of future performance, instrumentation data should be used in justifiable
predictions of future behavior of tunnels in similar environments.
3. Legal evaluation, valid instrumentation data can be valuable for potential litigation
relative to construction claims.
4. Development and verification of future designs, analysis of the performance of existing
tunnels and instrumentation data generated during operation, can be used to advance the
state-of-the-art of design and construction.
More than one type of the same instrument may be used to provide a backup system even when
its accuracy is significantly less than that of the primary system. Repeatability can also give a
clue to data correctness. It is often worthwhile to take many readings over a short time span to
determine whether a lack of normal repeatability indicates suspect data. Murray (1990) discussed
the layout of the instrumentation program. He concluded that the most practical approach is to
provide detailed monitoring at a few locations where the conditions are considered more critical
and to augment this information with settlement observations from elsewhere. The most critical
areas could be those where the structures are highest, where the soft soils are deepest, or where
the consequences of failure are most unacceptable. The wrong type of instruments placed in
inappropriate locations can provide information that may be confusing, or divert attention away
from other signs of potential distress. The factors needed to be considered in selecting a
particular design of instrument are summarized as following:
1. The instrument must be able to be transported and installed without damage or significant
change of calibration. It must be able to withstand the effects of construction and
sustained loading.
2. The use of simple direct operating mechanisms may avoid the problems of maintenance
and reliability. Conversely, if complex systems were to be used, possibly because of a
requirement for automation or accuracy, consideration would need to be given to how
such instruments should be maintained.
3. Damage by construction activity may be minimized or even prevented by placing the
monitoring system outside the area of the construction activity.
4. A particular design of instrument may more readily lend itself to a requirement for
automatic data logging.
5. The choice of instrument on the basis of cost should include the costs incurred in
installing, reading and evaluating the data.

99

6. Some types of instrument employ measurement principles that have better stability
characteristics than the other ones.
7. The selection of a particular instrument to attain a desired accuracy can have an influence
on most of the other factors referred to above.
Most instrumentation measurement methods consist of three components: a transducer, a data
acquisition system, and a linkage between these two components. A transducer is a module that
translates a physical change into analogous electrical signals whilst data acquisition systems are
the portable readout units. The different measuring techniques used in geotechnical
instrumentation are classified into one of the following categories:
(a) Pneumatic devices are used in pneumatic piezometers, earth pressure cells, and liquid
level settlement gages.
(b) Vibrating wire devices are used in piezometers, earth pressure cells, and liquid level
settlement gages and deformation gages.
(c) Electrical resistance devices are used in strain gauges, which in turn have been used in
many measurement devices.
(d) Other devices that use the electric or the magnetic field changes due to linear
deformations.
Generally, the extent of the utilization of instrumentation (e.g., number and spacing of different
types of measurements) depends on the variability of site conditions along and normal to the
different sides of the excavation. Karlsrud (1997b) proposed that the scope on monitoring may
be divided into the following categories of purposes:
Category A: Verification of the basis for- and soundness of the design (e.g., that the
overall performance of the design and the various elements: wall, struts, anchors, ground
improvement, etc.).
Category B: Detection of unacceptable performance that may lead to a failure or a serious
consequence to allow needed remedial measures to be undertaken in their due time.
Category C: Optimization of the design (e.g., use of the observational design approach).
Category D: Documentation of the influence and consequences on surrounding
structures.
Category E: Documentation and verification that the quality of the construction works are
as planned.
Category F: Provide data that may be utilized in improving numerical, analytical or
empirical design tools or procedures for works of similar type and in similar ground.
Category G: Provide data that may help to enhance the general understanding of the
ground response to the subject construction works and to calibrate soil models and
analytical tools.
Geotechnical instrumentation for deep excavation projects may be classified into two main types
namely: the deformation-measuring instruments, and the stress-measuring instruments. The
deformation instruments are used to assess the ground displacement fields and the lateral
deformations of the wall. The stress measuring instruments are used to measure the pore water
pressure, the soil pressure and stresses in wall. Negro et al. (2009) attempted to weigh the
usefulness of the main types of measurements in relation to deep excavation projects as shown in

100

Table 3 & 4. In the following sections, a brief description of each type of instrumentation is
given in the following sections.

Table 3. Weighted value of different types of deformation measurements (Negro et, 2009)
Category/Purpose

Hor. Disp.
at wall

A- Design verification
5
B- Warning against failure
5
C- Observational design
5
approach
D- influence of
4
surroundings
E-Verify quality of
4
construction
F- Improve design rules
5
G-Enhance knowledge
5
5 high (a must) - 1 low (not required)

Hor. Disp.
behind
wall

Ground
surface
Sett.

Vert. dist.
of ground
movement

Sett. of
surr.
structures

3
3

4
2

1
1

3
2

Tilt and
strain in
sur.
structures
1
1

4
5

4
5

2
5

4
4

3
3

Table 4. Weighted value of different types of stress measurements (Negro et, 2009)

Category/Purpose

Loads in
struts or
anchors

A- Design verification
3
B- Warning against failure
5
C- Observational design
5
approach
D- influence of
1
surroundings
E-Verify quality of
4
construction
F- Improve design rules
5
G-Enhance knowledge
5
5 high (a must) - 1 low (not required)

Temp. in
struts

Strain in
wall

Pore press.
Within
excav.

3
4

3
3

1-5
1-5

3-5
1-3

Earth and
pore
pressure
against
wall
1
1

1-5

1-3

1-2

3-5

1-4

1-4

3
3

3
5

1-4
2-4

1-4
3-5

4
5

101

Pore
pressure
outside
excav.

9.1.

Deformation Instrumentations

The ground displacement around tunnels could be classified as superficial and subsurface
displacements occurring in both vertical and horizontal directions. The instruments for
measuring ground deformation are grouped into the following categories:
1. Surveying is used to monitor the horizontal and vertical deformations of the ground
surface monuments near the deep excavations. Surveying methods include optical
leveling, measuring offsets from a baseline, triangulation, electronic distance
measurement, trigonometric leveling, photogrammetric methods, and the satellite-based
global positioning system. All surveying methods must be referenced to a stable
reference datum: a benchmark for vertical deformation measurements and a horizontal
control station for horizontal deformation measurements. Surface measuring points,
Figure 82, must be stable and robust and must survive throughout the whole project life.

Figure 82. Measuring point for monitoring surface settlement

2. Inclinometers are devices for monitoring deformation parallel and normal to the axis of a
flexible pipe by means of a probe passing along the pipe as shown in Figure 83.The
inclinometer probe can measure the tilt of a casing pre-installed in a borehole and grouted
in place. The internal face of the casing bears four longitudinal grooves (at quarter circle
locations) in which the probe to slide on two-pairs of spring-loaded wheels along the
casing, thus measuring the tilt along this orientation from bottom to top. After that, the
probe can be re-inserted in the perpendicular orientation of the casing and measure the tilt
along this line as well, thus permitting to measure the tilt along two perpendicular
directions.
102

Figure 83. Inclinometer measurement of displacement

3. Single and multiple point rod extensometers are devices to monitor displacements at
various depths employing a rod or multi-rods, Figure 84, which are anchored at one end
of a borehole. Relative movements between the end anchors and the reference tube of the
borehole are measured with either a dial depth gauge or a vibrating wire transducer. The
multi-point magnetic extensometer, Figure 85, contains ring magnets sliding on a central
access tube. The magnets are fixed in the ground at locations where movement is to be
monitored.

Figure 84. Rod extensometer

103

Figure 85. magnetic multiple point extensometers

104

9.2.

Stress Measurements

Deep excavations change the state of stresses in the adjacent soil mass. It may be important to
quantify the effect of deep excavations on the state of stress in soil to estimate the stresses inside
the soil and wall. However, these types of measurements are usually not successful, because the
presence of the soil stress instruments affects the measured stresses and the installation
(ORouke, 1979). A better alternative is to equip the wall with sets of strain gauges for
determining strains and loads in the wall and hence revealing the acting soil pressure on the wall
using back analysis.
9.2.1.

Piezometers

Deep excavation may affect the water pressure inside the soil. In this case the groundwater
requires monitoring. Piezometers are used to measure groundwater pressure. They are installed
in boreholes from the ground surface. Piezometers require some movement of pore water to
activate the measuring unit. The time required for water to flow to or from the piezometer to
create equalization with the porewater pressure is called the time lag.
Time lag is not significant when piezometers are installed in highly pervious soils such as sands;
the piezometers installed in such formation is called Standpipes or Casagrande piezometers.
Standpipes piezometers are the simplest and the cheapest piezometers. They consist of simple
tubes with a porous tip connected to its lower end. Groundwater level corresponds to the height
of the water surface in the standpipe above the piezometer tip, which is measured with a
dipmeter or using Bourdon gauge for artesian water.
If the soil has low permeability (e.g. clays or silts) and consequently high lag times, it is required
to use a vibrating-wire, a pneumatic or a hydraulic type of piezometers. A vibrating wire
piezometer has a tip comprises a porous element integral with a diaphragm type vibrating wire
pressure transducer, installed in boreholes. The pore pressure is transmitted through the porous
element causing a deflection in a diaphragm. The deflection is measured using the vibrating-wire
transducer. The pneumatic piezometer has a tip comprises a porous element integral with a
proven diaphragm transducer, installed either in a borehole or by pushing into shallow depths in
soft soil. Twin tubes connect the transducer to the portable readout unit. Air or nitrogen is forced
down one line and when its pressure equals the pore water pressure it forces open the membrane
valve and flows up the return line to a flow indicator. The balance pressure can then be read
using a Bourdon gauge. The hydraulic piezometer has a porous piezometer tip installed and
sealed above the measuring level. The tip is connected to the readout location by twin tubes. Deaired water is circulated through the tubes until the tubes and the tip fire completely filled; they
remain filled throughout the working life of the installation and can at any time be re-flushed to
remove air or gases that may have accumulated. The pore water pressure at the piezometer tip
can then be measured at the remote end of either water tube, making a correction for the head
difference between the tip and the measuring gauge. The instrument measures porewater
pressures by measuring the head of water using mercury manometers. Figure 86 shows the
features of the different piezometer types.

105

Figure 86. Features of different piezometers (Murray, 1990)

106

9.2.2.

Strain Gauges

Strain gauges are used for measuring the strains in struts and steel reinforcements. The stresses
are calculated from the strains by using theories of elasticity that relate the strain state to the
stress state. Bending moments and thrust forces in the wall are then calculated from the stresses.
Two types of gages are generally used for measuring strains: surface gages and embedded gages.
Strain gages can be further divided into short and long-term depending on the duration of
measurements. Usually short-term strains are best measured by electrical type gages, while some
long-term strains (e.g., creep and shrinkage) can be conveniently measured using detachable
mechanical gages. El-Nahhas (1980) describes three types of vibrating wire strain gauges to be
used in monitoring programs, namely: weldable gauges, surface gauges and embedded gauges.
Figure 87 shows some details of the three types of strain gauges.

Figure 87. Different types of vibrating wire strain gauges (El-Nahhas, 1980)

107

9.3.
9.3.1.

Real Time Monitoring


Robotic total stations (RTS)

One of the major developments in instrumentation systems in the recent years is the real-time
fully automated monitoring of essentially all parameters of interest using Robotic total stations
(RTS) and proprietary software interfaces that allow real-time monitoring of pre-determined
points. RTS are remotely operated theodolites that can deliver continuous 24/7 survey
measurements on reflective prismatic targets. RTS are also referred to as automated total stations
(ATS) and automated motorized total stations (AMTS).
The robotic total station (RTS) may be equipped with a reflectorless distance-meter. The
reflectorless surface points (RSPs) may be located on a flat, homogeneous and planar surface for
which vertical deformation is to be monitored as shown in Figure 88.
Data can be communicated through satellites, cell phones or internet systems and fed into a
database that is made accessible to all involved parties. Measurements can automatically be
compared to pre-set alert and alarm levels and those responsible can be automatically called up
on cell phone when alert and/or alarm levels are reached (Van der Poel et al., 2005; Finno et al.,
2007; Nyren et al., 2012; Tamagnan & Beth, 2012).

Figure 88. A Reflectorless Robotic Total Station (RRTS) measuring RSPs and prisms (Tamagnan & Beth, 2012)

108

An example of the application of real-lime monitoring is the City Creek - Block 76 project in
downtown Salt Lake City, Utah, USA. The construction comprises several residential high rises
with excavations having approximately 3,000 linear feet perimeter. The excavations are adjacent
to streets, light commuter rail, and structures of 10 to 22 floors, some of which are historic and
all of which are on shallow foundations. The excavation depth below the 22-story Gateway
Tower mat foundation will exceed 40 feet in depth. The consequence of shoring failure 40 feet
below the mat foundation of a 22-story high rise made the advantages of an automated real-time
survey monitoring obvious for the project. The automated real-time monitoring survey system
that was chosen for this project offered many advantages over a manual optical survey program
including: instantaneous data reduction and reporting, no instrument set-up for every acquisition
event, much better repeatability than manual monitoring, and internet access to the monitoring
data to all involved parties (Lange & Kippelen, 2008).
9.3.2.

Three-dimensional Laser scanning

Three-dimensional laser scanning is a terrestrial form of Light Detection And Ranging (LiDAR).
Like radar. LiDAR is a form of surveying by reflecting pulses of energy from objects and
measures time-of-flight to determine the distance from the scanner to individual objects as
shown in Figure 89. LiDAR has two well-known forms: Terrestrial Laser Scanning (TLS) &
Airborne Laser Scanning (ALS). Currently ALS has low accuracy so emphasis on using LiDAR
is concentrating on TLS. The scanner can acquires a set of points whose position is known
relative to the scanner to generate a three-dimensional model of the features surrounding a deep
excavation. These features can be used to monitor the excavation and to generate a threedimensional finite element models of the building located nearby the excavation. TLS has been
used in the geotechnical monitoring of deep excavations, tunnels (during construction and post
construction degradation); rockcuts along transportation corridors; construction (piles, shoring,
etc.); landslides; dams; and building deformation. TLS is an extremely fast and adequately
accurate for deep excavations. The high resolution nature of the data enables realistic images and
models for reporting of results (Laefer et al., 2006; Lato, 2012).

Figure 89. Operating schematic of a TLS scanner (Lato, 2012)

109

9.4.

Trigger Levels for Monitoring

Trigger levels (aka, response values or hazard warning values) are defined as pre-defined values
of the measured parameters. If an instrument reading is higher than this a trigger value, then a
pre-defined action is carried out. It is common to use two or more trigger values during
monitoring of construction to denote different levels of response, given the magnitude of the
reading and urgency or significance of the required response. Commonly, the traffic light system
is adopted (viz., Green, Amber and Red trigger levels). The following trigger zones are
commonly defined (Devriendt, 2012):
Green: OK, proceed
Amber (aka, Threshold, Alert, Review, or Warning): Monitor more frequently, review
calculations and start implementing contingency measures if trends indicate the Red
trigger may shortly be reached.
Red (aka, Limit, Maximum, Action, Response, or Tolerable limit): Implement measures
to cease movements and stop work
The above trigger zones are separated by two trigger levels (Amber and Red) which can be
considered as two separate unrelated scales; one related to calculated movements and one
relating to tolerable movements. As such, the values of the triggers for deep excavations can be
defined as (Devriendt, 2012; Patel et al., 2007):
Amber trigger is set close to the calculated displacement from analysis (usually at 75 or
80% of the calculated settlement;
Red trigger is based on a tolerable damage or deformation criteria. It can be considered as
a conservative estimate of when a serviceability limit state is likely to be exceeded.
An example for setting the trigger levels for a deep excavation for monitoring building
deformations and triggering remedies for damage is shown in Figure 90.

Figure 90. Setting trigger levels for a building subject to settlement from a deep excavation

110

9.5.

Monitoring Experience in Egyptian Deep Excavation Projects

Monitoring has become one of the essential elements in the Egyptian deep excavation projects
since its application in the Greater Cairo Metro Line 1 in the late 1980s and early 1990s (ElNahhas, et al., 1988, 1989, 1990, 1994; Shalaby, 1989; Abdel-Rahman, 1993; and others).
Figure 91 shows the location of the two instrumented test sections of the first line of Cairo Metro
and the details of the instrumentation at the Orabi Station (El-Nahhas 2006). Geotechnical
monitoring of the deep excavations was also implemented in Line 2 to document the ground
settlement and the wall lateral deformations associated with deep excavation in the stations.
Abel-Salam, 1995; Ahmed & Abel-Salam, 1996; Assaf, 1997; El-Nahhas et al., 1998; Mostafa,
1998; Fayed, 2002; Ahmed et al., 2005 and El-Nahhas, 2006).

Figure 91. Instrumented Section at Orabi Station, Cairo Metro Line 1 (El-Nahhas, 2006)

111

The hydraulic efficiency of the grouted plug of line 1 was monitored using the readings of a set
of pneumatic and electrical piezometers. The in-situ measurements confirmed the full ability of
the grouted plug to control of groundwater. Figure 92 shows the measured pore water pressure
under the bottom of excavation just before casting the reinforced concrete floor (El-Nahhas,
1994, El-Nahhas, 2006; El-Nahhas et al., 2006).

Figure 92. Measured pore water pressure below the deep excavations of Cairo Metro Line 1

Instrumentation was also implemented in the underground garages and basements such as new
headquarters of the Faisal Islamic Bank in Cairo (El-Nahhas & Morsy, 1996), Tahrir
underground garages (Abdel Rahman, 2007) and a building in Dokki, Cairo (Abdel Rahman &
El-Sayed; 2002a, 2002b & 2003; El-Sayed & Abdel Rahman, 2002; Abdel Rahman & El-Sayed,
2009). Figures 93 & 94 show the instrumentation plans for two of the aformenetioned case
histories in the Greater Cairo.

Figure 93. Layout the monitoring settlement points (Abdel Rahman & El-Sayed; 2002a, 2002b & 2003; El-Sayed &
Abdel Rahman, 2002; Abdel Rahman & El-Sayed, 2009)

112

Figure 94. The monitoring system for Al-Tahrir garage (Abdel-Rahman, 2007).

113

10.

BUILDING DAMAGE CRITERIA

Deformation of the ground may cause noticeable damage to the structure. This damage depends
not only on the ground deformations but also on the structural aspects of the affected building.
The most settlement sensitive buildings to ground deformations are masonry load bearing walls
or frames with masonry in-fill walls. Bearing walls and frames located perpendicular to deep
excavations tend to become distorted with shear strain and lateral strain. Assessing the potential
damage to new or existing structures by ground deformations has been the subject of many
studies that were undertaken since 1940s & 1950s (Meyerhof, 1947; Skempton and MacDonald,
1956; Meyerhof, 1956; Polshin & Tokar, 1957).
A purely theoretical approach to estimating building response to excavation-related deformations
is not possible due to the variability of the many factors that contribute to the response.
Consequently, building response is estimated utilizing simplified structural approximations to
provide limiting criteria/threshold against unacceptable damage. Generally, damaging criteria
due to deep excavation can be divided into superstructure damage criteria and piling overstressing assessment. In the following sections, these aspects are reviewed and the important
traits are highlighted.

10.1.

Superstructure Damage criteria

Burland & Wroth (1974 & 1975) and Boscardin & Cording (1989) proposed the following
deformation measures (Figures 95 & 96):
1. Settlement (S) is the vertical movement of a point. The maximum settlement is denoted
as (Smax).
2. Differential or relative settlement (S) is the difference between two settlement values.
The maximum differential settlement is denoted as (Smax).
3. Rotation or slope () describes the change in gradient of the straight horizontal line
defined by two reference points embedded in the structure with respect to their initial
horizontal orientation. The maximum rotation is denoted as (max).
4. Angular distortion () is an angle that produces sagging (or upward concavity) when it is
a directed downward from the building titled as a rigid body, or hogging (or downward
concavity) when is directed upward from titled rigid body building. The maximum
angular distortion is denoted as (max). The mode of cracking of a distressing building
affected by excessive settlement depends on the mode of deflection (hogging and
sagging) as they induce different damages.
5. Deflection Ratio (DF=/L) is defined as the quotient of relative defection () and the
corresponding length (L).
6. Tilt () describes the rigid body rotation of the whole superstructure or a well-defined
part of it. In certain cases also rigid body tilt can cause substantial damage, although this
is not commonly acknowledged, especially when several rigid bodies are connected (such
as in a row of houses built together or building parts connected by flexible joints). For
very large rigid body tilts (over 1:100) tilt becomes clearly noticeable and can effect
structural stability and functional performance.
7. Average horizontal strain h develops as a change in horizontal length over the
corresponding length; i.e., h = (L2 L1)/L.
114

Figure 95. Definition of the deformations affecting the building based on Burland & Wroth (1974 & 1975) and
Boscardin & Cording (1989)

115

Figure 96. Definition of sagging and hogging deformation modes (modified from Franzius, 2003)

116

10.1.1.

The maximum angular distortion () criterion

This criterion was initially proposed by Skempton and MacDonald (1956) for differential
settlements under gravity loads. It is appropriate only for the analysis of buildings under its own
weight and not under deformations induced by excavations. However, it is commonly adopted in
building damage analysis and hence it is reviewed herein.
Skempton and MacDonald (1956) examined 98 case histories to identify a basis on which to
determine allowable total and differential foundation settlements. Their database mainly
comprised infilled steel or reinforced concrete framed. Of these buildings, 40 were damaged by
settlement and 58 did not have damage. They correlated the damage with angular distortion
and established the following limiting angular distortions for aesthetic and structural damages:
1. Cracking of panels in frame buildings or walls in load bearing wall structures was likely
to occur if exceeded 1/300.
2. Structural damage to columns and beams was likely if exceeded 1/150.
The 1/300 & 1/150 limits correspond to 2 and 4 cm differential settlement, respectively, for
typical 6 m spans.
Skempton and MacDonald (1956) recommended that in excess of 1/500 should be avoided
when possible, and that this should be decreased to 1/1000 to prevent all damage. These values
correspond to 1.2 and 0.6 cm, respectively, for 6 m spans.
It should be noted that Skempton and MacDonalds (1956) categorization of degrees of damage
was limited to the distinction between cracking and structural damage. Additionally, their
recommendations were limited to traditional steel and concrete frame buildings and structures
with load bearing walls. No consideration was given to the relative length to height ratio of the
affected parts of the structures.
Meyerhof (1956 & 1982) suggested more stringent criteria and made a distinction between load
bearing wall structures and frame structures. He recommended limiting angular distortion to the
following:
1. 1/1000 for load bearing walls (sagging mode)
2. 1/2000 for load bearing walls (hogging mode)
3. 1/500 for panel walls of brick and similar unit masonry (infill frames)
4. 1/250 for beams and columns of frames.
Polshin and Tokar (1957) recommended an angular distortion limits that vary from 1/500 for
steel and concrete frame infilled structures to 1/200 where there is no infill or danger of damage
to the cladding.
Bjerrum (1963) presented data relating angular distortion to building performance based on
additional data and the Skempton and MacDonald (1956) data. Bjerrum (1963) suggested more
levels of serviceability damage based on the angular distortion of the building as shown in Figure
97.

117

Figure 97. Damage criteria based on angular distortion (Bjerrum, 1963)

10.1.2.

The maximum deflection ratio (/L) criterion

Polshin and Tokar (1957) studied over 100 buildings in the former Soviet Union, including the
effect of the building geometry based on L/H. where L is the length between two joints in the
building and H is the building height. They considered the deflection ratio (/L) as a structural
criterion is related to the curvature and they used 0.05% as the limiting tensile strain for brick
unreinforced walls using an analytical approach. They concluded the following limits of the
deflection ratio for unreinforced load bearing walls:
Sagging mode - (L/H 3)
(/L)max = 1/3300 to 1/2500
Sagging mode - (L/H 5)
(/L)max = 1/2000 to 1/1400
10.1.3.

The limiting tensile strain criterion

Burland & Wroth (1974 & 1975), Burland et al. (1977), Mair et al. (1996) and others assumed
that the onset of visible cracking in a given material may be linked to a limiting tensile strain
similar to Polshin and Tokar (1957). In this criterion, the building is modeled as a beam
deforming in the same shape as the settlement trough as shown in Figure 98 for the sagging
mode. Cracking may occur due to horizontal tensile strains from bending or diagonal tensile
strains from shear.
118

Figure 98. Deep beam model (Burland and Wroth, 1974 & 1975; Burland et al. 1977; Burland et al., 2001)

Using this simplified model, expressions were derived for the limiting deflection ratio (/L)
considering a maximum bending axial tensile strain b(max) and maximum shear diagonal tensile
strain d(max) as listed in Table 3.
Table 3. Expression for the limiting deflection ratio (Burland & Wroth, 1974 & 1975; Burland et al., 1977)
Limiting deflection ratio
Limiting deflection ratio
Case
(bending)
(shear)
Central point load

3
=[
+
]

12 2 ()

2
= [1 +
]

18 ()

Uniform load

5
3
=[
+
]

48 2 ()

1 5 2
=[ +
]

2 144 ()

Where
is the mid span deflection
L is the length in sagging / hogging
H is the height of the building from foundation to roof
t is the distance of the neutral axis to the edge of the beam. Burland and Wroth (1974 &
1975) assumed the neutral axis of the deep beam model was located at the middle of the
beam for unreinforced bearing walls undergoing settlement sagging mode (t=H/2). For
the hogging, unreinforced bearing walls were assumed to have neutral axis at the bottom
of the building (t=H). For reinforced walls and frame structure, the neural axis is assumed
to be located at the bottom of the wall.
G is the shear modulus of the beam representing the building
E is young's modulus of the beam representing the building
119

E/G equals to 2.6 for masonry walls, 12.5 for frames. Refer to Section 10.3.1 of this state-ofthe-art report.
I is the moment of inertia of the building (H3/12) per unit width for sagging mode and
(H3/3) per unit width for hogging mode
b(max) is the maximum bending tensile strain
d(max) is the maximum diagonal shear strain
The maximum strain in this analysis refers to a limiting strain at the onset of visible cracking.
Polshin and Tokar (1957) reported appearance of visible cracking in brick walls at a tensile strain
of 0.05%. Burland and Wroth (1974 & 1975) chose a strain value equal to 0.075%. Burland et al.
(1977) correlated the limiting tensile strains for unreinforced masonry walls and the crack width
as demonstrated in Table 4. Generally, the maximum strains that cause failure in common
building materials vary widely as a function of material and mode of deformation (Boone, 1996).
Assuming E/G = 2.6. Burland & Wroth (1974 & 1975) expressed the equation of the
concentrated load in Table 3 as follows:
1. Sagging mode (neutral axis is located at the middle of the beam):
a. Bending:

H
= [0.167 + 0.65 ] ()

H
B

(5)

2
= [0.25 ( ) + 1 ] ()

(6)

b. Shear:

2. Sagging mode (neutral axis is located at the middle of the beam)


a. Bending:

H
= [0.083 + 1.3 ] ()

H
B

(7)

2
= [0.064 ( ) + 1 ] ()

(8)

b. Shear:

Figures 99 & 100 show the ratio /(L.c), where c is the limiting strain, as a function of L/H for
sagging and hogging modes, respectively.

120

Table 4. Damage categories according to Burland et al. (1977) & BRE Digest 251 (1995)
Approx.
Category of
Degree of
crack
Strain
Damage description
damage
damage
width
(%)
Ease of repair in italic type
(mm)
Hairline cracks of less
than about 0.1 mm
0. Negligible
0- 0.1
0-0.05
which are classed as
negligible. No action
required.
Fine cracks which can be
treated easily using
normal
decoration.
Damage
generally
0.05restricted to internal wall
1. Very slight
0.1- 1
0.075
finishes; cracks rarely
visible
in
external
brickwork. Typical crack
widths up to 1 mm.
Aesthetic

2. Slight

Functional

3. Moderate

1- 5

5-15
Or a
number of
cracks > 3
mm

0.0750.15

Cracks easily filled.


Recurrent cracks can be
masked
by
suitable
linings.
Cracks
not
necessarily
visible
externally; some external
repointing
may
be
required
to
ensure
weather-tightness. Doors
and windows may stick
slightly and require
easing and adjusting.
Typical crack widths up
to 5 mm.

0.15-0.3

Cracks which require


some opening up and
can be patched by a
mason. Repointing of
external brickwork and
possibly a small amount
of brickwork to be
replaced. Doors and
windows
sticking.
Service
pipes
may
fracture.
Weathertightness often impaired.
Typical crack widths
are 5 to 15 mm, or
several of, say, 3 mm.

121

Examples

Category of
damage

Serviceability

Degree of
damage

4. Severe

Approx.
crack
width
(mm)

Strain
(%)

15-25,
depends
upon
number of
cracks

More
than 0.3

Structural

5. Very severe

Damage description
Ease of repair in italic type

Extensive damage which


requires
breaking-out
and replacing sections of
walls, especially over
doors and windows.
Windows
and
door
frames distorted, floor
sloping
noticeably*.
Walls leaning or bulging
noticeably*, some loss of
bearing
in
beams.
Service pipes disrupted.
Typical crack widths are
15 to 25 mm, but also
depend on number of
cracks.
-------* Local deviation of slope,
from the horizontal or vertical,
of more than 1/100 will
normally be clearly visible.
Overall deviations in excess of
1/150 are undesirable.

Structural damage which


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
25 mm, but depend on
number of cracks.

Usually
more than
25 mm but
it depends
upon
number of
cracks

122

Examples

Figure 99. Threshold of damage for sagging of load bearing walls, E/G = 2.6 (Burland and Wroth, 1974 & 1975)

Figure 100. Threshold of damage for hogging of load bearing walls, E/G = 2.6 (Burland and Wroth, 1974 & 1975)

123

Buildings sited adjacent to excavations are generally less tolerant to excavation-induced


differential settlements than similar structures settling under their own weight. This is attributed
to the lateral strains that develop in response to most excavations. These strains add to the strains
imposed by the vertical movements associated with the excavation (Boscardin & Cording, 1989;
Son & Cording, 2005; and others).
To account for the combined effect of both excavation and gravity loads, Boscardin and Cording
(1989) superimposed the additional horizontal strains induced by excavations to the bending and
shear tensile strains induced by gravity loads. Their database of structural damage and ground
movements was established during the construction of the metro in Washington D.C. and
included data on a number of structures that were instrumented as part of this project. Their data
differed from earlier studies as much of the previously collected data were for structures
deforming under their own weight while ground movements around excavations include
significant lateral in addition to the vertical deformations.
Boscardin and Cording (1989) developed damaging criterion for buildings adjacent to
excavations in form of multi-dimensional relationship between the angular distortion , the
horizontal strain h and the expected tensile strain/degree of severity as shown in Figure 101. The
criterion was based on the state of strain of a simple deep beam with L/H=1, E/G=2.6, and
neutral axis at the bottom of the beam. The critical tensile strains for different damage levels
were determined considering the field observations of damage associated with deep excavations
and tunnels. The lateral strain limits proposed by Boscardin and Cording (1989) match with the
lateral displacement criteria developed by the National Coal Board for deep mines in the United
Kingdom (NCB, 1975).

Figure 101. Relationship of Damage to Angular Distortion and Horizontal Extension Strain
(Boscardin & Cording, 1989)

Son & Cording (2005) provided analysis for the empirical criteria presented by Boscardin and
Cordings (1989) by estimating the principle tensile strain due to angular distortion and lateral
strain, Figure 102, as follows:
c = l (cos max)2 + sin max cos max
tan (2max) = /l

Where max is the direction of crack formation measured from the vertical plane.
124

(9)
(10)

Son and cording interpret the empirical envelopes of the damage zone presented by Boscardin &
Cording (1989) as envelope of constant critical tensile strain (c). They modified Boscardin &
Cordings (1989) envelopes, as shown in Figure 103, based on the presented above simple
analysis. Son and Cording (2005) also investigated the ratio /(/ L) and found that it is ranged
between 2 and 4. For elastic (including frame structures) and for minor cracking conditions, the
ratio /(/ L) is in the range of 2, but for severe cracking in a structure, it increased up to 4.

Figure 102. Tensile strain components due to horizontal strain, angular distortion and tilting for wall with L/H=1 &
E/G =2.6 (Son & Cording, 2005)

Figure 103. Damage zones with different critical tensile strains (Son & Cording, 2005)

Burland (1997) included lateral strain based on the work of Boscardin and Cording and adapted
different values of critical strain to reflect different damage categories, as illustrated by Figure
125

104. However, this approach was also limited to the case of L/H = 1 unless successive graphical
constructions and interpretation are carried out. He also considered that there was no evidence to
suggest that a strain of 0.3% could cause severe damage in spite of data provided by the National
Coal Board and Boscardin and Cording.

Figure 104. Damage criterion according to Burland (1997).

10.1.4.

The crack width criterion

Boone (1996 & 2001) and Boone et al. (1998 & 1999) developed a methodology for calculating
potential tensile strain experienced by structural members and related this strain to anticipated
crack widths. The analysis considers bending strains, elongation of building elements, and lateral
ground movements. Figure 105 summarizes the procedure proposed by Boone and his coworkers
to estimate the cumulative tensile crack width (Ct) and cumulative principal crack width (Cp).
Boone (1996, 2001) assumed that crack widths are appropriate indicators of damage severity. He
correlated the crack widths with empirical observations on damage of buildings and developed a
damage chart predict damage. Figure 106 shows Boone's chart that relates crack widths and the
expected damage.

126

Figure 105. Boones (1996 & 2001) procedure to assess the expected crack widths

Figure 106. Damage criterion according to crack widths (Boone,1996).

127

10.1.5.

The maximum settlement (Smax) maximum rotation (max) criterion

Rankin (1988) classifies the damage risks for buildings subject to settlement using two
deformation characteristics: Maximum settlement (Smax) and the maximum rotation/slope (max)
as shown in Table 5. He did not account for the horizontal deformations associated with deep
excavations and tunnels.
Table 5. Risk categories according to Rankin (1988)
Maximum
Maximum
Risk
No.
category
rotation (max)
Settlement (max)
Negligible
Less than 1/500
> 10 mm
1
2

Slight

1/500-1/200

10 50 mm

Moderate

1/200-1/50

50 75 mm

High

Greater than 1/50

Greater than 75 mm

10.1.6.

Description of the risk


Superficial damage unlikely
Possible superficial damage which is
unlikely to have structural significance
Expected superficial damage and
possible structural damage to building,
possible damage to relatively rigid
pipelines
Expected structural damage to buildings.
Expected damage to rigid pipelines,
possible damage to other pipelines

The Damage Potential Index (DPI( criterion

Schuster et al. (2009) developed a procedure to assess the potential for building damage adjacent
to an excavation. A Damage Potential Index (DPI) is defined as
DPI = c / (1/200) x 100

(11)

Where c is calculated using Equations 9 & 10.


The building serviceability requirements are established in terms of limiting DPIs, the depth of
excavation He and the distance between the building and the excavation d as shown in Table 6.
Table 6. Levels of Building Damage versus DPI thresholds (Schuster et al. 2009)
Level of damage
DPI for sagging
DPI for hogging
caused by
settlements
settlements (d/He
Remedial measures
excavation
(d/He 1.4)
>1.4)
Damage levels 1 and 2 are considered as
0-15
0-10
1. Negligible
tolerable, and no scheme to protect adjacent
buildings is required.
15-25
10-20
2. Slight
At damage level 3, possible damage to adjacent
buildings might be intolerable. A protection
scheme might be required in the design stage. If
25-35
20-30
3. Slight to moderate
not implemented, great caution must be
exercised to monitor the building during the
construction.
Damage levels 4, 5, and 6 are definitely
35-60
30-50
4. Moderate
intolerable. The excavation design should be
reexamined and possibly changed. Or, a proper
protection scheme must be implemented to
60-85
50-80
5. Severe
protect adjacent buildings.
>85
>80
6. Very severe

128

10.2.

Assessment of the Induced Building Damage

Burland (1995), Mair et al. (1996) and Son & Cording (2005) provided a systematic procedure
for damage assessment of buildings. The design approach consists of three stages:
1. Preliminary assessment
2. Second stage assessment
3. Detailed evaluation.
The three phases are shown schematically in Figure 107 and elaborated in the following sections.

Figure 107. Three-phased damage assessment flow chart (Burland, 1995; Mair et al., 1996; and Son & Cording,
2005)

10.2.1.

Primary assessment

In the primary assessment, the greenfield settlement trough is evaluated. Rankin (1988)
recommendations for the negligible damage zone (i.e., 1/500 & Smax of less than 10mm) are
adopted. Buildings which are located within that zone (i.e., the greenfield prediction gives lower
129

values than the above thresholds) are assumed to experience negligible damages. Such buildings
are not considered further in order. This approach is very simple and conservative as only
greenfield settlement is considered. The above values of maximum slope and settlement may
need to be reduced when assessing the risk for structures of higher sensitivity (i.e., building with
stone or glass claddings and important aesthetical features that should be maintained); however,
for most structures the abovementioned damage criterion can be utilized. If the greenfield
settlement associated with a building exceeds the maximum slope and settlement a second stage
assessment has to be carried out.
10.2.2.

Second stage assessment

In this stage of the risk assessment the building is represented as an elastic deep beam whose
foundation is assumed to follow the greenfield movement trough described by the empirical or
semi empirical equations. The appropriate ratio of E/G should be adopted as elaborated in the
Section 10.3.1. The strain within the beam is evaluated following an approach described in more
detail in Sections Error! Reference source not found., 0 & Error! Reference source not
found.. Categories of damage, defined in previous sections can then be obtained from the
magnitude of strain.
Although this approach is more detailed than the preliminary assessment it is still conservative as
the building is assumed to follow the greenfield settlement trough. Burland (1995) indicated that
the category of damage obtained from this assessment shall only be considered for aesthetical
damage (i.e., up to slight damage category with a limiting bending or diagonal tensile strain of
0.15-0.167%).
10.2.3.

Detailed assessment

In this stage details of the building and of the deep excavation should be taken into account using
advanced modeling such as:
Geotechnical conditions, sub-surface profile and groundwater conditions.
The three-dimensional aspects of the deep excavation construction.
The building stiffness (timber, masonry or framed buildings) and weight (number of
stories, live and dead loadings).
The building orientation with respect to the deep excavation.
Building features such as the foundation design and structural continuity as well as any
previous movement a building may have experienced in the past.
Sensitivity and usage of the building such as office, private home, public building, sports
facility etc.
The influence of the building's stiffness is likely to reduce its deformation. Burland (1995)
indicated that because of the conservative assumption of the second stage assessment, the
detailed evaluation will usually predict lower categories of damage than obtained from the
previous stage.
If the risk of damage remains high after the detailed assessment, necessary protective measures
are to be considered in the form of risk mitigation plans as described in Section 12 of this stateof-the-art.

130

10.3.
10.3.1.

Features Affecting Structural Damage


Ratio of the buildings Young modulus to its shear modulus (E/G)

Burland & Wroth (1974 & 1975) and Burland et al. (1977) demonstrated that the building can
sustain higher deflection ratio (/L) with increasing of the ratio (E/G) as shown in Table 3. They
assume that the ratio of E/G is 2.6 for masonry walls and 12.5 for frames. For buildings that have
little or no tensile restraint (i.e., traditional masonry buildings), they recommended that the E/G
ratio should be reduced to 0 .5. The effect of E/G ratio on the maximum deflection ratio is shown
in Figure 108. Finno et al. (2005) deduced more detailed procedure for analysis of framed
structures that accounts for the distribution of the stiffness of the slabs and the columns.

Figure 108. Effects of E/G and neutral axis location for deep beam analysis (Finno et al., 2005)

131

Son & Cording (2005) proposed a methodology to assess the equivalent Young and shear moduli
for masonry walls considering the presence of opening using a finite element model of the wall
only as shown in Figure 109. Son & Cording (2007) utilized the same approach and simulated
the masonry walls using finite elements as brick elements for the masonry blocks connected by
joint elements for the mortar. They reported much higher values of E/G than Burland and his coworkers due to the effects of having openings (such as doors and windows) and if there are poor
mortars at wall joints which slightly affects the bending stiffness but largely effects the shear
stiffness. This implies that the cracking of walls with openings is mostly controlled by the
diagonal shear stresses that start from their concentration locations at doors and windows.

Figure 109. Estimation and of the equivalent wall modulii

132

10.3.2.

Grade beams

Boscardin and Cording (1989) investigated the effect of grade beams to reduce the greenfield
horizontal tensile strain gh to less strain h as shown in Figure 110, where EgA is the stiffness
and area of the grade beam foundation, Es is the soil stiffness, H is the height of excavation or the
length of the section of the foundation being strained, and S is the spacing between grade beams.
A great benefit is noticed by using ground beam. More than 90% of the greenfield direct tensile
strain is eliminated with the presence of a light grade beam; heavy grade beams do not add
substantial reduction to the direct tensile strain induced by deep excavations.

Figure 110. Effect of grade beams for two-story and three-bay structures

10.3.3.

Building-Soil Relative Stiffness

Goh (2010) and Goh & Mair (2011) proposed design curves for the effect of deep excavations on
buildings using modification factors (Ms) defined as the ratio of the deformation affecting
structure to greenfield deformation. Modification factors were proposed for deflection ratio
(DR) and horizontal strain (h) as follows:
Sagging: MDRsag = DRsag/DRgsag
Hogging: MDRhog = DRhog/DRghog
Compressive horizontal strain: Mhc = hc/ghc
Tensile horizontal strain: Mht = ht/ght

133

The modification factors are shown in Figures 111 & 112 in terms of the following relative
stiffness ratios (Goh, 2010; Goh & Mair, 2011):
A. Bending relative stiffness ratio
(13)

B. Axial relative stiffness ratio


(14)

Where E is the building Youngs modulus, I is the moment of inertia of the building as a deep
beam, Es is the soil modulus, L is the length of the building (in either hogging or sagging based
on the greenfield settlement trough), A is the cross-sectional area of the building as a deep beam,
and B is the total length of the building.

Figure 111. The modification factor for the deflection ratio (Goh, 2010, Goh and Mair, 2011)

134

Figure 112. The modification factor for the horizontal strain (Goh, 2010, Goh and Mair, 2011)

Dimmock and Mair (2008) recommended the following adaptions to the similar modification
factors proposed by Potts and Addenbrooke (1997) for tunnels to improve the results:
In hogging: use the bending stiffness of the foundation only
In sagging: reduce the calculated bending stiffness by about one order of magnitude to
include the effect of window and door openings.
The same adaptations would be required in Goh (2010) and Goh & Mairs (2011) modification
factors for deep excavations (Korff, 2012).

135

10.4.

Damage Assessment for Deep Foundations

Foundation piles in or adjacent to deep excavations may be damaged during the excavation. This
can be caused by horizontal deformation, especially in very soft soil. Recently, some studies
were developed to address the effect of excavations on piles using theoretical methods (e.g.,
Poulos & Chen, 1996a, 1996b & 1997; Poulos, 2005). Generally, there is a lack of field data to
verify these theoretical methods as it is practically impossible to instrument existing piles in the
field.
Poulos & Chen (1997) developed design charts for the preliminary assessment of a pile response
to the deformation induced by a nearby deep excavation for a standard problem shown in Figure
113. The maximum pile deflection and maximum bending moment for this case, for a single pile
in a homogeneous clay layer. The non-linear behavior of the pile has been allowed for by
imposing lateral pile-soil pressures with a limiting value equal to 9su, where su is the undrained
shear strength of the clay.
The results of Poulos & Chens (1997) analyses have then been expressed in terms of basic
solutions for the standard problem, with correction factors for each of the parameters varied.
Figures 114 & 115 show the maximum moment and deflection, respectively, for the standard
case. The following approximate expressions are derived for the maximum pile deflection, and
the maximum bending moment, Mmax:
= b . kcu. kd. kNc. kEI. kk ks
M = Mb . kcu. kd. kNc. kEI. kk ks

(15)
(16)

Where b is the basic maximum deflection, Mb is the basic maximum bending moment, kcu &
k'cu are the correction factors for the undrained shear strength, kd & k'd are the correction factors
for pile diameter, kNc & kNc are the correction factors for depth of excavation (depending on Nc
=h/su, where is the average soil unit weight, su is the undrained shear strength, h is the
excavation height), kEI & kEI are the correction factors for wall stiffness, kk & kk are the
correction factors for strut (or support) stiffness, and ks.& ks are the correction factors for strut
(or support) spacing.
Figures 116 & 117 show the various correction factors for the case of a free-head unrestrained
pile. The moment and lateral deformations estimated by Equations 15 & 16 should be
superimposed with the existing bending moment and deformations.
Leung et al. (2000) used the centrifuge model test to investigate the behavior of piles subject to
deformation induced by unstrutted deep excavations in dry dense sand. It was found that, for the
case of a pile situated within relatively stable ground, that is, either behind a stable retaining wall
or outside the soil failure zone of a failed wall, the theoretical predictions were generally in good
agreement with the measured values. For the case of a pile located within the soil failure, the
analytical approaches do not give accurate results of the pile deformation and moments.

136

Figure 113. Basic problem for the pile response (Poulos & Chen, 1997).

Figure 114. Basic bending versus distance from excavation face (Poulos & Chen, 1997).

Figure 115. Basic movement versus distance from excavation face (Poulos & Chen, 1997).

137

Figure 116. Correction factors for bending moment (Poulos & Chen, 1997).

138

Figure 117. Correction factors for lateral pile movement (Poulos & Chen, 1997).

139

10.5.

Design Deep Excavations for Admissible Structural Deformations

Clough et al. (1989), Clough and ORourke (1990) showed that the system stiffness affects the
maximum lateral deformation of the wall and hence it affects the deformations affects nearby
buildings.
Zapata-Medina (2007) presented a revised system relative stiffness factor that is more related to
the deformations induced by excavation as demonstrated in the Section 3 of this-state-of-the-art.
He also introduced a methodology for designing deep excavation support to achieve specified
deformations to avoid distressing nearby buildings. The flow chart for this approach is shown in
Figure 118.

Figure 118. Iterative designing of excavation support systems for admissible deformations (Zapata-Medina, 2007)

140

11.

SURVEY OF DAMAGE DUE TO INDUCED GROUND MOVEMENT

A well-documented damage survey may be needed to assess the damage of a building located
nearby a deep excavation. BRE Digest 251 (1995) required a thorough survey showing records
of damage and distortion reasonably consistent with foundation movement as demonstrated in
Figure 119.

Figure 119. Crack patterns associated with different modes of ground settlements

Generally, cracks induced by ground movement are not widely distributed throughout the
building, but tend to be concentrated in areas where maximum structural distortion and structural
weak points coincide.
The following procedure was recommended by BRE Digest 251 (1995) in assessment of
building damage induced by ground movements:
1. Sketch each damaged walls, draw the position and direction of any cracks. Both external
and internal crack patterns are to be plotted on the same elevation so that the mode of
distortion and cause of movement can be better understood.
141

2. Distinguish where possible between tensile cracks, compressive cracks (indicated by


small flakes of brick squeezed from the surface and by localized crushing) and shear
cracks (indicated by relative movement along a crack of points on opposite side of it).
3. Note the direction of any crack taper, crack widths, and the frequency of cracks if they
are too numerous to record individually. An example of crack plotting is shown in Figure
120.

Figure 120. Crack plotting on building elevation

4. Take photographs of cracked zones as they provide a useful record of crack patterns and
density.
5. Try to determine the approximate age of the cracks by questioning the occupants on the
date of discovery and by examining the fracture surfaces. Additionally, recent cracks in
brickwork have a clean appearance, whereas older cracks show signs of dirt
accumulation.
6. Where possible, measure or estimate the magnitude of any distortion and movement of
the building. Examples are tilt and bulge of walls, slope of floors and slip on damp-proof
course.
7. Describe the impairment to the serviceability of the building. For example: jamming of
doors and windows, cracking of window, penetration of draughts and rainwater through
cracks, and fracturing of service pipes.
8. Give a thorough description of the materials of walls and finishes and their condition to
enable the identification of causes other than ground movement (e.g., shrinkage of
concrete products, or differential thermal expansion of dissimilar materials) as well as to
assist in the selection of suitable methods of structural repair.
9. Record details of the construction and the basic information about the affected structure
as these details can affect significantly the response of the structure to foundation
movement, the way in which damage may occur and the form of remedial works.

142

12.

RISK MANAGEMENT AND MITIGATIONS

Many sources of risks are associated with the construction of deep excavations including:
Ground movements, groundwater control, and improper quality of construction. Some of the
commonly-acknowledged risk categories are shown in Table 7. The major sources for the
aforementioned risks are the uncertainness in the soil properties and the construction procedure.
Table 7. Examples of uncertainty in the geotechnical works (Patel et al., 2007)
Example
No.
Geotechnical Uncertainty
Geological
Complex geology & hydrogeology
1
Parameter and modeling
Undrained soil verses drained behavior
2
Ground treatment
Grouting, dewatering
3
Construction
Complex temporary work
4

Risk management in deep excavations can be performed by identifying the different risk sources
and carrying out risk analyses using the following procedure (Ahuja, 1994; Abdel-Rahman,
2007; Lee at al., 2007):
1. Estimating the probability of occurrence of the undesirable event;
2. Estimating the magnitude of consequences;
3. Identifying options to accommodate the risks, including:
a. Reducing the probability of the cause;
b. Mitigating the consequence; and
c. Reducing the escalation from cause to consequence.
4. Prioritize risk management efforts based on:
a. Level of risk (probability and consequence);
b. Status of risk control and risk management activities; and
c. Optimum timescale for risk control action.
Risk control could be always ensured through the following:
1. Incorporating a design with adequate safety factor and reasonable ground movements that
could be safely tolerated by the surrounding structures.
2. Incorporating an inclusive quality control program during construction.
3. Performing a pre-construction dilapidation survey to verify the conditions of the
surrounding structures and their safety conditions when subjected to the predicted ground
movements.
4. Adopting an elaborate monitoring system that suit the risk sources associated with the
execution of the deep excavation.
Contingency plans are used in the event of emergency response, back-up operations, and disaster
recovery for construction projects which carry a large element of risk. The contingency plan
shall therefore focus upon ways in which certain events identified through completion of project
risk assessments can be militated against using a set of pre-identified procedures. The plan shall
be fit-for-purpose and undergo the following key tests prior to its release:
1. Is the plan achievable in reality, should this be required?
2. Are the trigger mechanisms for actual activation of the plan clear and realistic?
3. Does the plan address anticipated situations in a timely, affordable, effective, consistent
manner?
143

Puller (2003) listed the following contingency measures to reduce the deformation induced with
the deep excavation and hence reduce the risks of affecting nearby buildings:
1. Use of construction methods such as the top-down system or preloading of temporary
struts may achieve reductions in settlements below nearby buildings.
2. Strengthening the ground by means of cement or chemical grout injection, mix-in-place
or pin piles. In extreme cases, freezing of the subsoil may prove an effective solution in
granular, water-bearing soils.
3. Temporary or permanent strengthening the affected building by means of vertical support
and horizontal ties to resist horizontal tensile strains imposed by the soil deformation.
Shear stiffness of the building may also be improved by temporarily filling window and
door openings in facades and cross-walls with brickwork or blockwork of requisite
strength.
4. Structural jacking applied progressively as the deep excavation is made to counteract
vertical settlement, possibly with improvements of temporary strengthening to the
structure.
5. Compensation and fracture grouting may be applied progressively as deep excavation is
made. Compaction grouting applied to both granular and cohesive subsoils can provide a
means of lifting structures to counteract the effects of vertical settlements. Successive
injections of compensation and fracture grouting may be carried out from tubes-amanchette drilled in arrays from positions both inside and outside the affected structure.
Abdel-Rahman (2007) illustrated the applicability of the risk management approach to mitigate
the risks of affecting structures nearby a deep excavation for a multi-story underground garage in
Al-Tahrir square, Cairo, Egypt (Figure 121). He studied in details the risks associated with the
deep excavation in this project and as follows:
1. Risk situations developed during the excavation of the diaphragm wall trenches,
especially at the panels that were close to Omar Makram minaret (6.0 m away).
a. Any failure to one of the diaphragm wall panels would cause a collapse to the
minaret due to its proximity to the diaphragm wall trench. The footing width (5.0
m) is very comparable to the length of the diaphragm wall panels (2.80 m). Also,
the relatively high contact stress (about 200 kN/m2) elevates the risk as well.
b. The effect of any tilting that might develop due to the excavation process would
result in an obvious tilt to the minaret because of its height (50 m).
2. Risks developed from the effect of anticipated ground settlement and lateral deformations
on the structure of Al-Sadat station:
a. Possible instability of the diaphragm wall trenches during excavation as a result of
an improper quality of excavation, or a particular ground anomaly.
b. Effect of the induced ground movements due to trenching and strutted excavation
on the structural safety of the metro station as well as its shallow-founded
entrance.
c. Possible instability/disfunctioning of the base grout plug might affect the safety of
the metro station.
d. Possible ground loss due to migration of soils, if water leakage between the
diaphragm wall panels occurred during water pumping from within the
construction pit.

144

(b) Monitoring of the settlement for the


Minaret of Omar Makram Mosque
(a) Plan of the site and the monitoring system

(c) Section A-A: The garage boundary relevant to the tunnel of


the Greater Cairo Metro Line-2

(e) Section C-C: The garage boundary with respect to Al-Sadat


Station

(d) Section B-B: The garage boundary


relevant to Al-Sadat Station.

(f) Elevation Reference Points (ERPs) and


crack indicators on Omar Makram Mosque

(g) Crack indicators on Omar Makram Mosque


Figure 121. Al-Tahrir garage, its surrounding structures and monitoring system (Abdel-Rahman, 2007).

145

Abdel-Rahman (2007) also outlined the risk mitigation plan that was undertaken to manage the
aforementioned risks after analysis of the foundations of the surrounding structures as follows:
1. A pre-construction dilapidation survey was carried out for all surrounding structures to
locate any structural anomaly. This survey is aimed at recording the existing conditions
of the surrounding structures so that any updates. Elevation reference points (ERP) were
installed in order to monitor the settlement of the structures. Crack indicators are also
installed where cracks or structural joints exist as shown in Figure 121.
2. The effect of wind on the tilting of Omar Makram minaret (50 m height) was also
recorded prior to construction for a period of two months at two points.
3. A nonlinear geotechnical two-dimensional analysis was carried out, Figure 122, which
predicts the maximum lateral displacement as 20 mm at the top; while the maximum
predicted settlement at the location of Al-Sadat station at the tip of its diaphragm wall is 2
mm.

Figure 121. Predicted lateral displacement of the diaphragm wall while advancing the construction stages (AbdelRahman, 2007)

146

4. Serviceability criteria were chosen based on the geotechnical and structural analyses as
follows:
a. Al-Sadat Station: The carried-out structural analyses indicated that the maximum
safe settlement is 5mm. It was anticipated from the analyses that this limit will not
be reached as the predicted value is only 2mm.
b. Omar Makram Mosque: Based on literature and reported values, a limit of 1/700
was considered for Omar Makram Mosque to avoid any distresses in its structure.
c. Omar Makram Minaret: The maximum tilting along the height of the minaret
from the analyses was determined by geotechnical and structural analyses to be
1/1500.
5. Large scale field pumping tests were performed before developing the excavation within
the entire site in order to detect unforeseen sources of risks encountered due to
construction problems such as leaking within the tanking system. A high variation in the
water table was noticed within two adjacent compartment indicating construction
problems in the cut-off between these two compartments. This reflected the presence of a
major opening that allowed water seepage between the two compartments during
pumping. This location was traced through the piezometers readings and sealed by
injecting jet-grout columns.
6. During the construction stages, the monitoring reports of the different variables (surface
and deep settlement, lateral movements, drawdown, etc.) were compared with the limits
set during the design stage.
7. A contingency plan of actions was prepared to meet unforeseen conditions as summarizes
in Table 8.

Table 8. Contingency plans for deep excavation (Abdel-Rahman, 2007)


Risk source
Contingency plan of action
Excessive lateral movement of the wall and Increase the number of lateral supports
ground settlement
Instability of the grout plug
Refill the excavation pit with water up to the level that
adequately re-stabilize the situation, or perform heavy
dewatering to lower the water table as needed.
Insufficient drawdown to the water below Increase the number of wells
excavation level
Lateral leaking from the support system
Inject grout columns behind the leaking locations

147

13.

SUMMARY OF THE PRESENTED WORKS

Deep excavations occasionally cause failures of adjacent structures yet they often produce
serviceability problems to nearby buildings in form of wall cracks, tilting and impairments to
windows and doors due to the ground deformations associated with deep excavations. With the
increasing demands for deep excavation in urban areas having soft deltaic soils such as the
Greater Cairo, it becomes increasingly important to have well-designed support systems for deep
excavations that do not only ensure the stability of the excavation itself but also warrant that the
excavation will not cause damage to the adjacent buildings and utilities due to potentially
excessive ground deformations.
The optimal design for a supporting system for deep excavations must include assessment of the
induced deformations using empirical/semi-empirical approaches and analytical methods.
Numerous empirical methods were envisaged to estimate the deformations associated with deep
excavations based on the worldwide and local databases of the monitoring observations in deep
excavation projects. The monitoring factual data may be augmented using numerical or
analytical approaches to reliably produce semi-empirical correlations. As the deformations
associated with deep excavation depend on diverse aspects (e.g., the soil strength and stiffness,
the time-dependent soil behavior, the wall stiffness, the method and sequence of construction, the
prestressing of the struts and anchors, the presence of nearby structures, etc.), engineers should
utilize analytical approaches methods such as the finite element method and the Artificial Neural
Network to validate the findings of the empirical and semi-empirical approach.
Generally, empirical/semi-empirical methods focus on the assessment of the maximum
settlements behind the wall, maximum lateral deformations of the walls, the widths of the
settlement troughs and distributions of settlements and lateral deformations behind the walls. The
main findings of the commonly-acknowledged empirical/semi-empirical methods presented in
this state-of-the-art are listed in Table 9.
Numerical analysis of the braced excavation is considered as one of the most complicated soilstructure interaction problems to be undertaken by a geotechnical engineer. Many different
aspects of soil are to be incorporated into numerical models; these aspects vary with regards to
their complexities. Simple models do not apprehend all aspects of the soil behavior while
complex models have many parameters that either do not have a clear physical meaning or
cannot even be determined using routine tests. Additionally, advanced numerical threedimensional time-dependent modeling usually require huge amount of input parameters and
computational resources. The following aspects are to be considered while considering numerical
analysis of deep excavations (Abdel-Rahman, 1993; Bentler, 1998; Strom & Ebeling, 2001;
Gutierrez et al., 2002; Kung, 2010):
1. Accuracy of finite element analysis of excavation depends generally on the performance
of soil constitutive model incorporated with its required parameters.
2. Realistic modeling of such problems should consider the simulation of the interface
medium between the ground and the wall.
3. Rational idealizations of the construction processes, as well as the workmanship quality
are musts although workmanship can only be judgmentally considered into the finite
element.
148

Table 9. Summary of the findings of the common empirical/semi-empirical methods for average workmanship in
terms of the excavation depth (He) or the Trench/pile Depth (D)
Max. wall Max. surficial
Study
Supp.
Const.
Trough
Soil Type
lateral
ground
Reference(s)
type* Activities**
width
deform.
settlement
Sands; medium to hard
1% He
2 He
F
A
clays
Peck (1969a)
Soft clay (limited depth)
F
A
2% He
3.5 He
Soft clay (extended)
F
A
>2% He
>3.5 He
Dense sands; stiff clays
F
A
0.3% He
2 He
ORourke
Soft clay (limited depth)
F
A
2% He
1.5 He
(1976)
Soft clay (extended)
F
A
>2% He
>2.2 He
Sands, gravels, very stiff
M
A
0.171% He
2 He
Goldberg et al.
to hard clays
0.5-2*sett.
(1976)
Soft clays
M
A
1.22% He
2.5 He
Sands,
M
P
Max. 0.3% He
2 He
0.2% He
Aver. 0.15%
Clough and
Stiff clays
M
P
3 He
He
ORourke
0.5-1 * wall
(1990)
Soft clays
M
P
Figure 9
lateral
2 He
deformation
Different soils
W
0.15% D
2D
Aver.
Aver. 0.22%
Sand, hard clay
M
A
0.19% He
He
Bentler (1998)
Aver.
Aver. 0.55%
Soft to stiff clay
M
A
0.45%He
He
Moorman
Aver.
Soft clay
M
A
Aver. 1.1% He
2 He
(2004)
0.87% He
Konstantakos
Soft Clay
S
A
0.2% He
0.2-0.4% He
(2008)
El-Nahhas et al.
(1988, 1989,
1990, 1994,
2006); Abdel
0.2-0.22%
Rahman (1993); Nile alluviums
S
A
0.1-0.15% He
1.5-2 He
He
Ahmed &
Abdel-Salam
(1996) and
others
3.8 He for
0.11% He for
shallow
shallow found.
found.
Abdel-Rahman

S
P
2.2 He for
& El-Sayed
Nile alluviums
0.03% He for
deep
(2009)
deep found.
found.
W
0.045% D
2D
* F: flexible support (sheet-pile walls, soldier beam and lagging); S: stiff support (secant pile wall, tangent pile wall,
diaphragm walls); M: mixed support stiffness
** A: All activities including dewatering and consolidation occurred during construction and monitoring, P: Pit
excavation only, W: Wall installation.
No massive dewatering due to utilization of a bottom grouted plug

149

Artificial Neural Networks (ANNs) are a form of Artificial Intelligence (AI) that attempts to
imitate the human brain and nervous system. The concept of ANN is to capture the relationship
between the model input variables and the corresponding outputs through learning by examples
and without need to set the form such relationship before the analysis. As such, they are
considered as an ideal tool to utilize the monitoring data and the numerical data to obtain quick
yet adequately accurate assessment of the deformation and refined soil and construction
modeling parameters. They have been utilized in assessment of the settlement associated with the
construction of the stations in the Greater Cairo Metro Line 2 (Fayed, 2002) and in the
constitutive modeling and inverse analysis of deep excavations (Hashash et al., 2003, 2004 &
2010).
Bolton and his co-workers (Osman & Bolton, 2004a, 2006 & 2007; Bolton et al., 1989, 1990a,
1990b, 2008 & 2010; Bolton, 2011) developed an analytical approach (Mobilizable Strength
Method) for analysis of deep excavations in which limit analysis for the undrained finite
deformations in clays is used to derive the mobilized shear stresses under working loads. This
approach needs a kinematically admissible deformation field that satisfies the boundary
condition in order to be utilized in conjunction with the Principle of Virtual Work.
Although they are not very common, centrifugal and physical models may be utilized to assess
the deformations induced by deep excavations. Centrifugal modeling can account for the scale
effect which can be considered in normal physical models. The main difficulty in using this type
of modeling is the realistic simulation of the excavation sequence. The recent development is this
type of modeling is the utilization of a small scale robotic excavator to remove the soil in-flight
in the centrifuge (Lam, 2010; Bolton, 2011). This new development makes centrifugal modeling
a valuable tool in studying the deformations associated with sequentially constructed deep
excavations.
The deformations patterns associated with deep excavations depend of the mode of the wall
deformations. Two basic patterns of the settlement troughs are commonly acknowledged:
spandrel settlement trough (associated with the wall cantilever deformations) and concave
settlement trough (associated with the wall bulging deformations). The cumulative settlement is
a function of the relative ratio between the wall bulging and cantilever deformations (Clough &
ORourke, 1990; Hsieh & Ou, 1998).
Many damage criteria have been set to assess the effect of the ground deformations induced by
deep excavations on building (Boscardin and Cording, 1989; Boone, 1996 & 2001; Boone et al.,
1998 & 1999; Burland, 1997; Son & Cording, 2005 & 2007). The main common aspect of these
approaches is the inclusion of the effect of the horizontal deformation caused by deep
excavations. Three-stage damage assessment is commonly applied to buildings located within
the influence zone of the deep excavation (Mair et al., 1996; Son & Cording, 2005). It is
anticipated that the consideration of the building stiffness reduces the deformations particularly if
the building is rigid (Goh, 2010; Goh & Mair, 2011).

Pile foundations are affected by deep excavations. It was noted by Abdel-Rahman & El-Sayed
(2002b & 2009) and Korff (2012) that most of the deformations affecting piles occur during the
150

wall installation. Poulos & Chen (1997) developed design charts for the assessment of a pile
response to the deformation induced by a nearby deep excavation in Clays. Analysis of
deformation effect on piles using analytical models such as Poulos & Chen (1997) approach
produces good results if there is an adequately factor of safety for the excavation; in failure and
near failure case, advanced nonlinear numerical analysis of the effect of deep excavations on
piles must be performed (Leung et al., 2000).
Monitoring programs and risk management are powerful tools in the observational approach to
allow construction to proceed smoothly in the face of the abundant risks associated with deep
excavation projects, particularly the risks associated with unforeseen geotechnical conditions or
construction problems. A proper prepared risk mitigation plans with well-set monitoring trigger
levels become a necessity in deep excavations especially in urban areas (El-Nahhas, 1992;
Abdel-Rahman, 2007). Recent advances in monitoring allow continuous monitoring of deep
excavations using Robotic Total Stations (RTS) and three-dimensional Terrestrial Laser
Scanning (TLS). The results of the monitoring are made available to all concerned parties
through modern communication means such as the Internet and cell phones (Finno et al., 2007).

151

14.

REFERENCES

1.

Abdel-Rahman, A.H. 1993. Numerical modeling of concrete diaphragm walls. M. Sc. Thesis, Ain Shams
University, Cairo, Egypt.

2.

Abdel-Rahman, A.H. 2007. Construction risk management of deep braced excavations in Cairo Australian
Journal of Basic and Applied Sciences, 1(4): 506-518.

3.

Abdel-Rahman, A.H. & El-Sayed, S.M. 2002a. Settlement Trough Associated with Diaphragm Wall
Construction in Greater Cairo. The Journal of the Egyptian Geotechnical Society, Cairo, Egypt.

4.

Abdel-Rahman, A.H. & El-Sayed, S.M. 2002b. Building subsidence associated with cut-and-cover
excavations in alluvial soils. Faculty of Engineering Scientific Bulletin, Ain Shams University, Vol. 37, No.
4, Cairo, Egypt.

5.

Abdel Rahman, A.H. & El-Sayed, S.M. 2009. Foundation subsidence due to trenching of diaphragm walls
and deep braced excavations in alluvium soils. 17th International Conference of ISSMGE, Alexandria, Egypt.

6.

Abdel-Salam, N. 1995. In-situ performance of a subway station in Cairo. M.Sc. Thesis, Ain Shams
University, Cairo, Egypt.

7.

Ahmed, A.A. & Abel-Salam, N. 1996. In-situ performance of subway stations in Cairo. Proc. of Seventh
International Colloquium on Structural and Geotechnical Engineering, Ain Shams University, Cairo, Vol. 1,
pp. 447-460.

8.

Ahmed, A.A.; Ali, H.E. & Fayed, A.L. 2005. Advanced numerical prediction of movements associated with
braced excavation. Ain Shams University, Faculty of Engineering, Scientific Bulletin, Vol. 40, No. 1, pp.
181-199.

9.

Ahuja, H. 1994. Project management techniques in planning and controlling construction projects. J. Wiley
& Sons, USA.

10.

Anderson, W.F.; Hanna, T.H.; Ponniah, D.A., & Shah, S.A. 1982. Laboratory-scale tests on anchored
retaining walls supporting backfill with surface loading. Canadian Geotechnical Journal, 19: 213-224.

11.

Assaf, O.M. 1997. Analysis of interaction of in-situ walls with the ground. M.Sc. Thesis, Ain Shams
University, Cairo, Egypt.

12.

Aye, Z.Z.; Karki, D. & Schulz, C. 2006. Ground Movement Prediction and Building Damage RiskAssessment for the Deep Excavations and Tunneling Works in Bangkok Subsoil. International Symposium on
Underground Excavation and Tunnelling, Thailand. pp. 281-297.

13.

Azevedo, R.F. 1983. Centrifuge and analytical modelling of excavation in sand. Ph.D. thesis, University of
Colorado, Bounder.

14.

Balay, J. 1985. Recommandations pour le choix des paramtres de calcul des crans de soutnement par la
method aux modules de reaction. Note dinformation technique, Paris, LCPC, 24 pp. Quoted from Delattre
(2001).

15.

Balay J.; Frank R. & Harfouche L. 1982. Programme DENEBOLA pour le calcul des soutnements par la
mthode des modules de reaction. Bulletin de liaison des Laboratoires des Ponts et Chausses, 120, juilletaot, pp. 3-12. Quoted from Delattre (2001).

16.

Bentler, D.J. 1998. Finite Element Analysis of Deep Excavations. PhD thesis, Virginia Polytechnic Institute
and State University.

17.

Benz, T. 2007. Small-strain stiffness of soils and its numerical consequences. Ph.D Thesis, Universitat of
Sttutgart. 193 pp.

18.

Bjerrum, L. 1963. Discussion. Proc. of the European Conference on Soil Mechanics and Foundation
Engineering, Vol. III, Wiesbaden, pp. 135.

19.

Bjerrum, L. & Eide, O. 1956. Stability of Strutted excavation in clay. Geotechnique, 6: pp. 32-47.

152

20.

Bolton, M.D. & Powrie, W. 1987. The collapse of diaphragm walls retaining clay. Geotechnique 37 (3), pp.
335- 353.

21.

Bolton, M.D. 2008. Geotechnical design to control ground movements. 7th Jennings Memorial Lecture.

22.

Bolton, M.D. 2011. Learning from reality: lessons from centrifuge models. International Symposium on
Backward Problems in Geotechnical Engineering TC302-Osaka 2011.

23.

Bolton, M.D.; Lam, S.Y. & Vardanega, P.J. 2010. Predicting and controlling ground movements around deep
excavations. Keynote Lecture in Geotechnical Challenges in Urban Regeneration, 11th Int. Conf. DFIEFFC,
London.

24.

Bolton, M.D.; Lam, S.Y. & Osman, A.S. 2008. Supporting excavations in clay from analysis to decision
making. TC 28 Conference Geotechnical Aspects of Underground Construction in Soft Ground, Shanghai.

25.

Boone, S.J. 1996. Ground Movement Related Building Damage. Journal of Geotechnical Engineering, 1996,
ASCE, 122(11), pp. 886 - 896.

26.

Boone, S.J. 2001. Assessing construction and settlement-induced building damage: a return to fundamental
principles. Proceedings, Underground Construction, Institution of Mining and Metallurgy, London, 559
570.

27.

Boone, S.J. 2003. Design of Deep Excavations in Urban Environments. Ph.D. Thesis. Toronto: University of
Toronto.

28.

Boone, S.J. & Westland, J. 2005. Estimating Displacements Associated with Deep Excavations. Geotechnical
Aspects of Underground Construction in Soft Ground: Proceedings of the 5 th International Symposium TC28.
Amsterdam, the Netherlands, 15-17 June 2005, Bakker, K.J.; Bezuijen, A.; Broere, W.; Kwast, E.A. (eds.),
pp. 817-822.

29.

Boone, S.J.; Garrod, B. & Branco, P. 1998. Building and utility damage assessments, risk and construction
settlement control. Tunnels and Metropolises, Balkema, pp. 243 - 248.

30.

Boone, S.J.; Westland, J. & Nusink, R. 1999. Comparative evaluation of building responses to an adjacent
braced excavation. Canadian Geotech. J., Vol. 36, 210-223.

31.

Borja, R.I.; Lee, S.R. & Seed, R.B. 1989. Numerical simulation of excavation in elastoplastic soils.
International Journal of Numerical and Analytical Methods in Geomechanics 13, pp. 231-249.

32.

Boscardin, M.D. & Cording, E.J. 1989. Building Response to Excavation-Induced Settlement. J. of Geotech.
Eng., ASCE, 115( 1), pp. 1-21.

33.

Boscardin, M.D.; Cording, E.J. & ORourke, T.D. 1979. Case studies of building behavior in response to
adjacent excavation. Report No. UMTA-IL-06-0043-78-2, U.S. Department of Transportation.

34.

Boudier J.; Colin C. & Mastikian L. 1970. Calcul de stabilit des parois sur ordinateur Exemples
dapplication. Travaux, 429, pp. 40-45. Quoted from Delattre (2001).

35.

Bowles, J.E. 1988. Foundation analysis and design. McGraw-Hill, Inc. New York.

36.

Brown, P.T. & Booker, J. R. 1986. Finite element analysis of excavations. Research Report No. 532, School
of Civil and Mining Engineering. The University of Sydney, Australia.

37.

Budge-Reid, A.J.; Cater, R.W. & Storey, F.G. 1984. Geotechnical and construction aspects of the Hong
Kong Mass Transit Railway system. Proceedings of the Second Conference on Mass Transportation in Asia,
Singapore .

38.

Building Research Establishment (BRE). 1995. Assessment of damage in low-rise buildings. BRE Digest 251
Concise reviews of building technology, BRE.

39.

Burd, H.J.; Houlsby, G.T.; Augarde, C.E., & Liu, G. 2000. Modelling tunnelling-induced settlement of
masonry buildings. Proc. Instn. Civ. Engrs. Geotech. Engineering, 143, pp. 17- 29.

40.

Burland, J.B. 1989. Ninth Laurits Bjerrum memorial lecture: Small is beautiful - the stiffness of soils at small
strain. Canadian Geotechnical Journal, Vol. 26, 499516.

153

41.

Burland, J.B. 1995. Assessment of risk of damage to buildings due to tunneling and excavation. Proc., 1st Int.
Conf on Earthquake Geotechnical Engineering, Tokyo.

42.

Burland, J.B. 1997. Assessment of risk of damage to buildings due to tunnelling and excavation. Earthquake
Geotechnical Engineering, Ishihara (ed.), 1997, Balkema, Rotterdam, pp. 1189 - 1201.

43.

Burland, J.B. & Wroth, C.P. 1974. Settlement of buildings and associated damage. Proceeding of a
Conference on Settlement of Structures, Cambridge, pp. 611-654.

44.

Burland, J.B. & Wroth, C.P. 1975. Settlement of buildings and associated damage. Building Research
Establishment Current Paper, Building Research Establishment (BRE) , Watford.

45.

Burland, J.B.; Broms, B.B. & Demello, V.F.B. 1977. Behavior of Foundations and Structures: State of the
Art Report. Proc. of the 9th Int. Conf. on Soil Mech. and Found. Eng., 1977, Tokyo, pp. 495-546.

46.

Caspe, M.S. 1966. Surface settlement adjacent to braced open cuts. Journal of Soil Mechanics and
Foundation Division, ASCE, 92(SM4): 51-59.

47.

Chai, J.; Shen, S.; Ding, W.; Zhu, H. & Carter, J. 2014. Numerical investigation of the failure of a building in
Shanghai, China. Computers and Geotechnics, 55, 482-493.

48.

Chandrasekaran, V.S. & King, G.J.W. 1974. Simulation of excavation using finite elements. Journal of
Geotechnical Engineering 100, pp. 1086-1089.

49.

Chang, C.V. 1969. Finite element analysis of soil movements caused by deep excavation and dewatering.
Ph.D. thesis, University of California, Berkeley.

50.

Christian, J.T. & Wong, I.H. 1973. Errors in simulating excavation in elastic media by finite elements. Soils
and Foundations 12(1), 1-10.

51.

Clough, G.W. 1975. Deep Excavations and retaining structures. Proceedings of the Short Course- Seminar
on Analysis and Design of Building Foundations, Bethlehem, Pennsylvania, pp 4 7-465.

52.

Clough, G.W. & Duncan, J. M. 1969. Finite element analyses of Port Allen and Old River Locks. Contract
Report S-69-6, U.S. Army Engineer Waterways Experiment Station, Vicksburg, MS.

53.

Clough, G.W. & Duncan, J.M. 1971. Finite element analysis of retaining wall behavior. Journal of the Soil
Mechanics and Foundations Division, 97(SM12), pp. 1657-1673.

54.

Clough, G.W. & Mana, A.I. 1976. Lessons learned in finite element analysis of temporary excavations.
Proceedings, 2nd International Conference, Numerical Methods in Geomechanics. ASCE, Blacksburg, VA.

55.

Clough, G.W. & ORourke, T.D. 1990. Construction induced movements of insitu walls. Design and
performance of earth retaining structures, Geotech. Special Publication No. 25, ASCE, Lambe and L.A.
Hansen, eds., 1990, pp. 439 - 470.

56.

Clough, G.W.; Weber, P.R. & Lamont, J. 1972. Design and observation of a tied-back wall. Proc., ASCE
Spec. Conf on Perf. of Earth and Earth-Supported Struct., ASCE, New York, N.Y., Vol. 1, pp. 1367-1390.

57.

Clough, G.W.; Smith, E.M. & Sweeney, B.P. 1989. Movement control of excavation support systems by
iterative design. Proc., ASCE, Foundation Engineering: Current Principles and Practices, Vol. 2, 869-884.

58.

Cording, E.J.; ORourke, T.D. & Boscardin. M.D. 1978. Ground movements and damage to structures. Proc.,
Int. Conf. on Evaluation and Prediction of Subsidence, Florida, 1978, 516-537.

59.

Cowland, J.W. & Thorley, C.B.B. 1985. Ground and building settlement associated with adjacent slurry
trench excavation. Ground Movements and Structures Proc., Third Int. Conf., University of Wales Institute
of Science and Technology, J. D. Geddes, ed., Pentech Press, London, England, pp. 723-738.

60.

Dawkins, W.P. 1994a. User's guide: Computer program for Winkler soil-structure interaction analysis of
sheet-pile walls (CWALSSI). Instruction Report ITL-94-5, U.S. Army Engineer Waterways Experiment
Station, Vicksburg, MS.

154

61.

Dawkins, W.P. 1994b. User's guide: Computer program for analysis of beam-column structures with
nonlinear supports (CBEAMC). Instruction Report ITL-94-6, U.S. Army Engineer Waterways Experiment
Station, Vicksburg, MS.

62.

Delattre, L. 2001. A century of design methods for retaining wallsThe French point of view. I. Calculationbased approaches conventional and subgrade reaction methods. Bulletin des laboratoires des Ponts et
Chausses, BLPC - 2001/234, pp. 33-52.

63.

Desai, C. S. & Abel, J.F. 1972. Introduction to the finite element method. Van Nostrand Reinhold Co., NY,
USA.

64.

Devriendt, M. 2012. Trigger levels for displacement monitoring. Geotechnical Instrumentation News (GIN),
March 2012, pp. 23-25.

65.

Dimmock, P.S & Mair, R.J. 2008. Effect of building stiffness on tunnelling-induced ground movement.
Tunnelling and Underground Space Technology, 23(4): pp 438450.

66.

Duncan, J.M. & Chang, C.Y. 1970. Nonlinear analysis of stress and strain in soils. Journal of Soil Mechanics
and Foundations Division 96(SM5), pp. 1629-1653.

67.

Duncan, J.M. & Clough, G.W. 1971. Finite element analyses of Port Allen Lock. Journal of Soil Mechanics
and Foundations Division 97(SM8), pp. 1053- 1068.

68.

Dunlop, P.; Duncan, J.M., & Seed, H.B. 1968. Finite element analyses of slopes in soil. Report TE 68-3,
Office of Research Services, University of California.

69.

El-Nahhas, F. 1980. The behaviour of tunnels in stiff soils. Ph. D. Thesis, Alberta University, Edmonton,
Canada

70.

El-Nahhas, F.M. 1992. Construction monitoring of urban tunnels and subway stations. Tunnelling and
Underground Space Technology, Pergamon Press, Vol. 7, No.4, pp. 425-439.

71.

El-Nahhas, F.M. 2003. Geotechnical aspects of controlling groundwater levels in urban areas. A Keynote
Paper, Proc. of the Tenth Int. Colloquium on Structural and Geotechnical Eng., Ain Shams Univ., Cairo, Vol.
6, KL4.

72.

El-Nahhas, F.M. 2006. Tunnelling and supported deep excavations in the Greater Cairo. Keynote Paper at
the Int. Symposium on Utilization of Underground Space in Urban Areas. International Tunnelling
Association & Egyptian Tunnelling Society, Sharm El-Sheikh, Egypt, pp. 27-56.

73.

El-Nahhas, F.M. & Morsy, M.M. 2002. Comparison of the measured and computed performance of propped
diaphragm retaining wall in Egypt. Proc. of the Int. Symposium on Numerical Models in Geomechanics,
Rome, Italy, pp. 579-586.

74.

El-Nahhas, F.M. & El-Mossallamy. 2009. The role of small strain constitutive model for predicting
differential settlement above tunnels. 13th International Conference on Structural and Geotechnical
Engineering. Ain Shams University, Faculty of Engineering, Department of Structural Engineering, pp. 144158.

75.

El-Nahhas, F.M.; Eisenstein, Z.; Shalaby, A.; Salam, M.; & El-Bedaiwy, S. 1988. Geotechnical performance
of diaphragm walls during construction of Cairo Metro tunnel. Proc. Int. Congress on Tunnels and Water,
ITA, Madrid, Spain, Vol.1, pp. 259-266.

76.

El-Nahhas, F.M.; Eisenstein, Z.; & Shalaby, A. 1989. Behaviour of diaphragm walls during construction of
Cairo Metro. Proc. Twelfth Int. Conf. on Soil Mechanics and Foundation Engineering, Rio de Janeiro, Brazil,
pp. 1483-1486.

77.

El-Nahhas, F.M., Eisenstein, Z., & Shalaby, A. 1990. In-situ behaviour of Orabi subway stations during
construction. Proc. of 1st Alexandria Conference on Structural and Geotechnical Engineering, Alexandria,
Egypt, Vol. 1, pp. 189-198.

78.

El-Nahhas, F.M.; Ahmed, A.A.; El-Gammal, A. & Abdel-Rahman, A.H. 1994. Modelling of braced
excavation for subway stations. Proc. of Int. Congress on Tunnelling and Ground Conditions. ITA, Cairo,
Published by Balkema, Netherlands, pp. 105-111.

155

79.

El-Nahhas, F.M.; Ahmed, A.; & Assaf, O. 1998. Analysis of in-situ walls interaction with the ground. Proc.
of Eighth Int. Colloquium on Structural and Geotechnical Eng., Ain Shams Univ., Cairo, Vol. 3, pp. 211-220.

80.

El-Nahhas, F.M.; Abdel-Rahman, M.T. & Iskander, G. M. 2006. Utilization of grouting techniques for
construction of underground structures in urban areas. Int. Symposium on Utilization of Underground Space
in Urban Areas, ETS & ITA, Sharm El-Sheikh

81.

El-Ramli, A.H. 1992. Geology and geotechnics in some areas in Egypt. Proc. of Int. Symposium on Current
Experiences in Tunnelling, National Authority for Tunnels and International Tunnelling Association, Cairo,
pp. 93-117.

82.

El-Sayed, S.M. 2001. Elasto-plastic three-dimensional analysis of shielded tunnels, with special emphasis on
Greater Cairo Metro. Ph.D. Thesis. Ain Shams University.

83.

El-Sayed, S.M. & Abdel-Rahman, A.H. 2002. Spatial stress-deformation analysis for installation of a
diaphragm wall. Faculty of Engineering Scientific Bulletin, Ain Shams University, Vol. 37, No. 3, Cairo,
Egypt.

84.

Elshafie, M. 2008. Effect of building stiffness on excavation-induced displacements. Ph.D thesis, University
of Cambridge.

85.

El-Sohby, M.A. & Mazen, O. 1985. Geology aspects in Cairo Subsurface Development. Proceedings of the
11th ICSMFE, San Francisco, Vol. 3, pp. 2401-2405.

86.

Fages R. & Bouyat C. 1971a. Calcul de rideaux de parois moules ou de palplanches Modle
mathmatique integrant le comportement irrversible du sol en tat lasto-plastique. Travaux, 439, 1971a,
pp. 49-51. Quoted from Delattre (2001).

87.

Fages R. & Bouyat C. 1971b. Calcul de rideaux de parois moules ou de palplanches Modle
mathmatique integrant le comportement irrversible du sol en tat lasto-plastique Exemple dapplication
Etude de linfluence de paramtres. Travaux, 441, 1971b, pp. 38-46. Quoted from Delattre (2001).

88.

Fausett, L.V. 1994. Fundamentals neural networks: Architecture, algorithms, and applications. PrenticeHall, Inc.,

89.

Fayed, A.L. 2002. Numerical modelling of underground subway stations. Ph.D. Thesis, Ain Shams
University, Cairo, Egypt.

90.

Finno, R.J. & Harahap, I.S. 1991. Finite element analysis of HDR-4 excavation. Journal of Geotechnical and
Geoenvironmental Engineering, Vol. 117, No. 10, 15901609.

91.

Finno, R.J. & Bryson, L.S. 2002. Response of building adjacent to stiff excavation support system in soft clay.
J. Perfom. Constr. Facil., 16(1), pp. 10-20.

92.

Finno, R.J.; Calvello, M. & Bryson, S.L. 2002. Analysis and performance of the excavation for the ChicagoState Subway Renovation Project and its effects on adjacent structures. Department of Civil Engineering,
Northwestern University. U.S. Department of Transportation.

93.

Finno, R.J.; Voss Jr, F.T.; Rossow, E. & Blackburn, J.T. 2005. Evaluating damage potential in buildings
affected by excavations. Journal of geotechnical and geoenvironmental engineering, 131(10), 1199-1210.

94.

Finno, R.J.; Langousis, M.; Winter, D.G.; Smith, M.W. & Chin, K.H. 2007. Real time monitoring at the Olive
8 Excavation. 7th International Symposium on Field Measurements in Geomechanics, 12 pp.

95.

Franzius, J.N. 2003. Behaviour of buildings due to tunnel induced subsidence. Ph.D. Thesis. University of
London. 358 pp.

96.

Franzius, J.N.; Potts, D. M.; Addenbrooke, T.I. & Burland, J.B. 2004. The influence of building weight on
tunnelling-induced ground and building deformation. Soils and Foundations 44(1): pp 25-38.

97.

Frischmann, W.W.; Hellings, J.E. & Snowdon, C. 1994. Protection of the Mansion House against damage
caused by ground movements due to the Docklands Light Railway Extension. Proc. of the Inst. of Civ. Eng.,
Geotech. Eng., 1994, Vol. 107, 65-76.

156

98.

Gaba, A.R.; Simpson, B.; Powrie, W. & Beadman, D.R. 2003. Embedded retaining walls guidance for
economic design. Report C580. CIRIA London, UK.

99.

Ghaboussi, J. & Pecknold, D.A. 1984. Incremental finite element analysis of geometrically altered structures.
International Journal of Numerical Methods in Engineering 20, 205I-64.

100. Gill, S. & Lukas, R. 1990. Ground movement adjacent to braced cuts. Design and performance of earth
retaining structures. ASCE Geotechnical Special Publications 25.
101. Goh, A.T.C.; Wong, K.S. & Broms, B.B. 1995. Estimation of lateral wall movements in braced excavation
using neural networks. Canadian Geotech. J., 32, pp. 1059-1064.
102. Goh, K.H. 2010. Response of ground and buildings to deep excavations and tunnelling. Ph.D thesis,
Cambridge University, UK.
103. Goh, K.H. & Mair, R.J. 2011. The response of buildings to movements induced by deep excavations.
Geotechnical Aspects of Underground Construction in Soft Ground - Proceedings of the 7th International
Symposium on Geotechnical Aspects of Underground Construction in Soft Ground, Rome. pp 903-910.
104. Goldberg, D.T.; Jaworski, W.E. and Gordon, M.D., 1976. Lateral support systems and underpinning. Report
FHWA-RD-75-128, Vol. 1, Federal Highway Administration, Washington D.C., p. 312.
105. Goodman, L. E.; Taylor, R. L. & Brekke, T.L. 1968. A model for the mechanics of jointed rock. Journal of the
Soil Mechanics and Foundations Division 94(SM3), pp.637-668.
106. Gourvenec, S. M. & Powrie, W. 1999. Three-dimensional finite-element analysis of diaphragm wall
installation. Gotechnique, Vol. 49(6), pp. 801-823.
107. Gowland, J.W. & Thorley, C.B.B. 1984. Ground and building settlement associated with adjacent slurry
trench excavation. Proceedings of the Third International Conference on Ground Movements and Structures,
Cardiff, pp 723-738.
108. Gutierrez, M.S.; Filz, G.M. & Ebeling, R.M. 2002. Numerical methods to model excavation of soil adjacent
to retaining structures. Report No. ERDC/ITL TR-02-7. Virginia Polytechnic Institute and State University;
U.S. Army Engineer Research and Development Center, Information Technology Laboratory. 105pp.
109. Haigh, S.K.; Houghton, N.E.; Lam, S.Y., Li, Z. & Wallbridge, P.W. 2010. Development of a double-axis
servo-actuator for novel centrifuge modeling. 7th Int. Conf. Physical Modelling in Geotechnics, Zurich, Vol.1,
239-244.
110. Haliburton, T.A. 1968. Numerical analysis of flexible retaining structures. Proc. ASCE, Journal of the Soil
Mechanics and Foundations Division, Vol. 94, SM6, pp. 1233-1251.
111. Haliburton, T.A. 1987. Soil structure interaction: Numerical analysis of beams and beam-columns. Technical
Publication No. 14, School of Civil Engineering, Oklahoma State University, Stillwater.
112. Hashash, Y.M.A. & Whittle, A.J. 1996. Ground movement prediction for deep excavations in soft clay.
Journal of the Geotechnical Engineering, Vol. 122, No. 6, pp. 474486.
113. Hashash, Y.M.A. and Whittle, A.J. 2002. Mechanism of load transfer and arching for braced excavations in
clay. J. Geotech. And Geoenviron. Engrg, Vol. 128, No. 3, March 1, 2002, pp.187-197.
114. Hashash, Y.M.A.; Marulanda, C.; Ghaboussi, J., & Jung, S. 2003. Systematic update of a deep excavation
model using field performance data. Computers and Geotechnics, 30(6), pp. 477-488.
115. Hashash, Y.M.A.; Jung, S.; & Ghaboussi, J. 2004. Numerical implementation of a neural network based
material model in finite element analysis. International Journal for numerical methods in engineering, 59(7),
pp. 989-1005.
116. Hashash, Y.M.A.; Levasseur, S.; Osouli, A.; Finno, R. & Malecot, Y. 2010. Comparison of two inverse
analysis techniques for learning deep excavation response. Computers and geotechnics, 37(3), 323-333.
117. Hashash, Y.M.A.; Osouli, A. & Marulanda, C. 2008. Central Artery/ Tunnel Project Excavation Induced
ground deformations. ASCE Journal of Geotechnical and Geoenvironmental Engineering, 134(9), pp. 13991406.

157

118. Hsiao, C.L. 2007. Wall and ground movements in a braced excavation in clays and serviceability reliability
of adjacent buildings. Ph.D. Thesis, Clemson University. 152 pp.
119. Hsieh, P-G. & Ou, C-Y. 1998. Shape of ground surface settlement profiles caused by excavation. Canadian
Geotechnical Journal, Vol. 35(6), pp. 1004-1017.
120. Hwang, H.; Bhhm-Soo, C.; Soo, K. & Sang-Seom, Jeong. 1996. Sequential analysis of earth retaining
structures using p-y curves for subgrade reaction, Journal of KGS, Vol. (12), No. 3, pp. 149-161.
121. Jardine, R.J.; Potts, D.M.; Fourie, A.B. & Burland, J.B. 1986. Studies of the influence of non-linear stress
strain characteristics in soilstructure interaction. Geotechnique, 36(3), 377-396.
122. Juang, C.H.; Schuster, M.; Ou, C-Y; & Phoon, K.K. 2011. Fully probabilistic framework for evaluating
excavation-induced damage potential of adjacent buildings. Journal of Geotechnical and Geoenvironmental
Engineering, Vol. 137, No. 2, pp. 130-139.
123. Karlsrud, K. 1997a. Some aspects of design and construction of deep supported excavation, Discussion
leader's contribution. Proc. 14th Int. Conf. on Soil Mech. Found. Eng. Hamburg. Vol. 4, pp. 2315-2320.
124. Karlsrud, K. 1997b. Panel contribution: Comments on use of performance monitoring for underground
works. International Conference on Soil Mechanics and Foundation Engineering, 14. Hamburg 1997.
Proceedings, Vol. 4, pp. 2413-2415.
125. Kimura, T.; Takemura, J.; Hiro-pka; A. Suemasa, N. & Kouda, N. 1993. Stability of unsupported and
supported vertical cuts in soft clay. 11th SEAGC, Singapore, pp 61-70.
126. Konstantakos, D.C. 2008. Online database of deep excavation performance and prediction. 6 th international
conference on case histories and geotechnical engineering. Arlington, paper 5.16 (1-12).
127. Korff, M. 2012. Response of piled buildings to the construction of deep excavations. Ph.D. Thesis. University
of Cambridge. IOS Press BV.
128. Korff, M. & Tol, A.F. 2012. Failure cost analysis of 50 deep excavations in The Netherlands.
http://www.ice.org.uk/ICE_Web_Portal/media/Events/Failure-cost-analysis-of-50-deep-excavations-in-TheNetherlands.pdf.
129. Kung, G.T.C. 2003. Surface settlement induced by excavation with consideration of small strain behavior of
Taipei silty clay. PhD Dissertation, Department of Construction Engineering, National Taiwan University of
Science and Technology, Taipei, Taiwan.
130. Kung, G.T.C.; Juang, C.H.; Hsiao, E.C.L. & Hashash, Y.M.A. 2007a. A simplified model for wall deflection
and ground surface settlement caused by braced excavation in clays. ASCE Journal of Geotechnical and
Geoenvironmental Engineering. 133(6): pp. 1-17.
131. Kung, G.T.C.; Hsiao, E.C.L. & Juang, C.H. 2007b. Evaluation of a simplified small strain soil model for
estimation of excavation-induced movements. Canadian Geotechnical Journal, Vol. 44, 726736, ISSN:12086010.
132. Kung, G.T.C.; Ou, C.Y. & Juang, C.H. 2009. Modeling small-strain behaviour of Taipei clays for finite
element analysis of braced excavations. Computers and Geotechnics, Vol. 36, No. 1-2, pp. 304-319.
133. Kung, G.T-C. 2010. Finite element analysis of wall deflection and ground movements caused by braced
excavations. In Finite Element Analysis. Moratal, D. (ed.). Publisher: Sciyo. 688 pp.
134. Laefer, D.F. 2001. Prediction and assessment of ground movement and building damage induced by adjacent
excavation. Ph.D. thesis, University of Illinois at Urbana-Champaign, Urbana,
135. Laefer, D.F.; Carr, H.; Morrish, S. & Kalkan, E. 2006. Opportunities and impediments to the use of threedimensional Laser scanning for adjacent excavations. Geocongress 2006, pp. 1-6.
136. Lam, S.S.Y. 2010. Ground movements due to excavation in clay: Physical and analytical models. Ph.D.
Thesis, University of Cambridge. 305 pp.

158

137. Lam, S.Y. & Bolton, M.D. 2011. Energy conservation as a principle underlying mobilizable strength design
for deep excavations. Journal of Geotechnical and Geoenvironmental Engineering, Vol. 137, No. 11, pp.
1062-1074.
138. Lam S.Y.; Elshafie, M.Z.E.B.; Haigh, S.K. & Bolton M.D. 2011. Development of a new apparatus for
modeling deep excavations in a geotechnical centrifuge. International Journal of Physical Modeling in
Geotechnics.
139. Lambe, T.W. 1970. Braced excavations. Proceedings of the ASCE Specialty Conference on Lateral Stresses
in the Ground and Design of Earth Retaining Structures, Cornell Univ. , Ithaca, N. Y., pp. 149-218.
140. Lange, C. & Kippelen, D. 2008. Real-time survey monitoring for support of excavation. Geostrata.
January/February, pp. 16-18.
141. Lato, M.J. 2012. Remote monitoring of deformation using Terrestrial Laser Scanning (TLS or Terrestrial
LiDAR). Geotechnical Instrumentation News (GIN), March 2012, pp. 27-27.
142. Lee, F.H. 2008. Application of large three-dimensional finite element analyses to practical problems. The
12th International Conference of International Association for Computer Methods and Advances in
Geomechanics (IACMAG) 1-6 October, 2008 Goa, India, pp. 125-132.
143. Lee, S.J.; Song, T.W.; Lee, Y.S.; Song, Y.H. & Kim, J.K. 2007. A case study of damage risk assessment due
to the multi-propped deep excavation in deep soft soil. In Proceedings of the 4th International Conference on
Soft Soil Engineering. Chan, D. & Law, K.T. (editors). Vancouver. Taylor and Francis, pp. 281-289.
144. Leung, C. F.; Chow, Y. K. & Shen, R. F. 2000. Behavior of pile subject to excavation-induced soil movement.
Journal of Geotechnical and Geoenvironmental Engineering, Vol. 126, No. 11, pp. 947-954.
145. Loh, C.K.; Tan, T.S. & Lee, F.H. 1998. Three dimensional excavation tests in the centrifuge. Centrifuge 98,
Tokyo.
146. Long, M. 2001. Database for retaining wall and ground movements due to deep excavations. Journal of
Geotechnical and Environmental Engineering 127(3): 203-224.
147. Lyndon, A. & Schofield, A.N. 1970. Centrifuge model test of short term failure in London clay.
Geotechnique 20, No.4, pp.440-442.
148. Mair, R.J. 2011. Tunnelling and deep excavations - Ground movements and their effects. Proceedings of the
XV European Conference in Soil Mechanics and Geotechnical Engineering (ECSMGE-Athens), pp. 39 70.
149. Mair, R.J.; Taylor, R.N. & Burland, J.B. 1996. Prediction of ground movements and assessment of risk of
building damage due to bored tunnelling. Geotechnical Aspects of Underground Construction in Soft
Ground. Proceedings of the International Symposium. Mair, R.J. and Taylor, R.N. (ed.). pp. 713718,
Rotterdam, Balkema, pp. 713-718.
150. Mana, A.I. & Clough, G.W. 1981. Prediction of movements for braced cuts in clay. Journal of Geotechnical
Engineering, American Society of Civil Engineers, vol. 107, pp. 759-777.
151. Maren, A.; Harston, C. & Pap, R. 1990. Handbook of neural computing applications, Academic Press, Inc.,
San Diego, California.
152. Mayne P.W. & Kulhawy, F.H. 1982. Ko-OCR relationships in soil. Journal of Geotechnical Engineering,
ASCE, 108 (GT6), pp. 851-872.
153. Meyerhof, G.G. 1947. The settlement analysis of building frames. The Structural Engineer, Vol. 25, 369.
154. Meyerhof, G.G. 1956. Some recent foundation research and its application to design. The Structural
Engineer, 1956, Vol. 31, pp. 151-167.
155. Meyerhof, G.G. 1982. Limit states design in geotechnical engineering. Structural Safety Journal, 1:6771,
1982.
156. Mitchell, J. K. & Soga, K. 2005. Fundamentals of soil behavior. 3rd edn, John Wiley and Sons.

159

157. Moh, Z.C.; Kong, S.K. & Hwang, R.N. 1999. Protecting adjacent buildings during underground
construction. Proceedings of Symposium on Innovative Solutions in Structural and Geotechnical Engineering
In Honor of Professor Seng-Lip Lee, May 14-15, 1999, Bangkok, Thailand.
158. Moormann, C. 2004. Analysis of wall and ground movement due to deep excavation in soft soil based on a
new worldwide database. Soils and Foundations, Vol. 44, No. 1, pp.87- 98.
159. Moormann, C. & Moormann, H.R. 2002. Study of wall and ground movements due to deep excavation in soft
soil based on worldwide experiences. Geotechnical Aspects of Underground Construction in Soft Ground,
Toulouse, Spcifique, Lyon.
160. Morrison, C.S. 1995. Development of modular finite element program for analysis of soil-structure
interaction. Ph.D. Thesis, Virginia Polytechnic Institute and State University, Blacksburg, Virginia.
161. Morton, K.; Cater, R.W. & Linney, L. 1980. Observed settlements of buildings adjacent to stations
constructed for the modified initial system of the Mass Transit Railway, Hong Kong. Proceedings of the Sixth
Southeast Asian Conference on Soil Engineering, Taipe1, vol. 1, pp 415-429.
162. Murray, R. T. 1990. Rapporteur's paper: Geotechnical Instrumentation in practice. Proceedings of the
conference of geotechnical instrumentation in civil engineering projects, Thomas Telford, London, England,
pp 75-85.
163. National Coal Board (NCB). 1975. Subsidence engineers handbook. National Coal Board Productions
Department, London.
164. Naylor, D.J.; Pande, G.N. & Tabb, R. 1981. Finite elements in geotechnical engineering. Pineridge Press.
245 pp.
165. Negro, Jr. A.; Karlsrud, K. & Vorster, E. 2009. Prediction, monitoring and evaluation of performance of
geotechnical structures.17th International Conference of ISSMGE, Alexandria, Egypt.
166. Ng, C.W.W. 1992. An evaluation of soil-structure interaction associated with a multi-propped excavation.
Ph.D. Thesis, University of Bristol, UK.
167. Ng, C.W.W. & Yan, W.M. 1999. A true three-dimensional numerical analysis of diaphragm walling.
Geotechnique, Vol. 49, No. 6, pp. 825834.
168. Nicholson, D.; Tse, C-M & Penny, C. 1999. The observational method in ground engineering: principles and
applications. CIRIA Report 185, London, 214 pp.
169. Nyren, R.; Drefus & R. & Johnson, S. 2012. Remote monitoring of deformation using Robotic Total Stations
(RTS). Geotechnical Instrumentation News (GIN), March 2012, pp. 29-29.
170. ORouke, J.E. 1979. Soil stress measurement experiences. J. of Geotech. Eng., ASCE, Vol. 104, No. GT12,
pp. 1501-1514.
171. ORourke, T.D. 1981. Ground movements caused by braced excavations. Journal of Geotechnical
Engineering, American Society of Civil Engineers, val. 107, pp 1159-1178.
172. ORourke, T.D. 1993. Base stability and ground movement prediction for excavations in soft clay. Retaining
Structures. Thomas Telford, London, pp. 131-139.
173. ORourke, T.D.; Cording, E.J. and Boscardin, M. 1976. The ground movements related to braced excavation
and their influence on adjacent buildings. U.S. Department of Transportation, Report no. DOT-TST 76, T-23.
174. Obrzud, R.F. 2010. On the use of the hardening soil small strain model in geotechnical practice. Numerics in
Geotechnics and Structures. Zimmermann Th., Truty, A. & Podkes, K. (eds).
175. Osaimi, A.E. & Clough, G.W. 1979. Pore-pressure dissipation during excavation. Journal of the
Geotechnical Engineering Division, 105(4), pp. 481-498.
176. Osman, A.S. & Bolton, M.D. 2004a. A new design method for retaining walls in clay. Can. Geotech. J. 41:
pp. 451466.

160

177. Osman, A.S. & Bolton, M.D. 2004b. A new approach to the estimation of undrained settlement of shallow
foundations on soft clay. Engineering Practice and Performance of Soft Deposits, IS-OSAKA, pp. 93-98.
178. Osman, A.S. & Bolton, M.D. 2004c. Estimation of undrained settlement of shallow foundations on London
Clay. International Conference on Structural and Foundation Failures August 2-4, Singapore, pp. 443-454.
179. Osman, A.S. & Bolton, M.D. 2006. Ground movement predictions for braced excavations in undrained clay.
J. Geotech. and Geo-env. Eng., ASCE, 132 (4), 465-477.
180. Osman, A.S. & Bolton, M.D. 2007. Back analysis of three case histories of braced excavations in Boston
Blue Clay using MSD method. 4th Int. Conf. Soft Soil Eng. Vancouver, London, Taylor Francis, pp 755-764.
181. Ou, C.Y. & Lai, C.H. 1994. Finite-element analysis of deep excavation in layered sandy and clayey soil
deposits. Canadian geotechnical journal, 31(2), pp. 204-214.
182. Ou, C.Y., Chiou, D.C. & Wu, T.S. 1996. Three dimensional finite element analysis of deep excavations.
Journal of Geotechnical Engineering. ASCE, Vol. 122, No. 5, pp. 337-345.
183. Ou, C.Y.; Hsieh, P.G & Chiou, D.C., 1993. Characteristics of ground surface settlement during excavation.
Canadian Geotechnical Journal, 30, pp. 758-767.
184. Ou, C.Y.; Shiau, B Y. & Wang, I.W. 2000. Three-dimensional deformation behavior of the Taipei National
Enterprise Center (TNEC) excavation case history. Canadian Geotechnical Journal, Vol. 37(2), pp. 438-448.
185. Patel, D., Nicholson, D., Huybrechts, N., & Maertens, J. 2007. The observational method in geotechnics.
Proceedings of the XIV European Conf. on Soil Mechanics and Geotechnical Engineering, Madrid, pp. 2427.
186. Peck, R.B. 1969a. State-of-the-art: Deep excavation and tunneling in soft ground. Proceedings of the Seventh
International Conference on Soil Mechanics and Foundation Engineering, Universidad Nacional Autonoma
de Mexico Instituto de Ingenira, Mexico City, Mexico, Vol. 3, pp. 225-290
187. Peck, R.B. 1969b. Advantages and limitations of the observational method in applied soil mechanics.
Geotechnique, Vol. 19, No. 2, pp. 171-187.
188. Pfister, P.; Evers, G.; Guillaud, M. & Davidson, R. 1982. Permanent ground anchors-Soletanche design
criteria. Report FHWA-RD-81-150, Federal Highway Administration, McLean, VA.
189. Poh, T.Y. & Wong, I. H. 1998. Effects of construction of diaphragm wall panels on adjacent ground: Field
trial. J. Geotech. and Geoenvir. Engrg., ASCE, 124(8), pp. 749756.
190. Poh, T.Y.; Goh, A.T.C. & Wong, I.H. 2001. Ground movements associated with wall construction: Case
Histories. J. Geotech. and Geoenvir. Engrg., ASCE, 127(12), pp. 1061-1069.
191. Polshin, D.E. and Tokar, R.A. 1957. Maximum allowable non-uniform settlement of structures. Proc. of the
4th Int. Conf. on Soil Mech. and Found. Eng., Vol. 1, pp. 402-405.
192. Potts, D.M. & Addenbrooke, T.I. 1997. A structure's influence on tunnelling-induced ground movements.
Proc. Instn. Civ. Engrs. Geotech. Engineering, 125, 109{125.
193. Potts, D.M., & Zdravkovi, L. 1999. Finite element analysis in geotechnical engineering: Theory. Thomas
Telford. 500 pp.
194. Potts, D.M. & Zdravkovic, L. 2001. Finite element analysis in geotechnical engineering: Application.
Thomas Telford. 448 pp.
195. Poulos, H.G. 2005. The influence of construction side effects on existing pile foundations. Geotechnical
Engineering, 36(1), pp. 51-68
196. Poulos, H.G. & Chen, L.T. 1996a. Pile response due to excavation-induced lateral soil movement. Jnl. Geot.
Eng., ASCE, Vol. 123,2,94-99.
197. Poulos, H.G. & Chen, L.T. l996b. Pile response due to unsupported excavation - induced lateral soil
movement. Can. Geot. Jnl., Vol. 33, 670-677.

161

198. Poulos, H.G. & Chen, L.T. 1997. Pile response due to excavation-induced lateral soil movement. Journal of
Geotechnical and Geoenvironmental Engineering, ASCE, Vol.123, No.2, 94-99.
199. Powderham, A.J. 1994. An overview of the observational method: development in cut and cover bored
tunnelling projects. Gotechnique, 44 (4), pp. 619-636.
200. Powderham, A.J. 1998. The observational method application through progressive modification. Civil
Engineering Practice: Journal of the Boston Society of Civil Engineers Section/ASCE, 13 (2), pp. 87-110.
201. Powderham, A.J. 2002. The observational method learning from projects. Proceedings of the Institution of
Civil Engineers: Geotechnical Engineering, 155 (1), pp. 59-69.
202. Powderham, A.J. & Nicholson, D.P. 1996. The way forward. In: The observational method in geotechnical
engineering. Nicholson, D.P. & Powderham, A.J. (ed.). Thomas Telford, London.
203. Powrie, W. & Li, E.S.F. 1991. Finite element analyses of an in situ wall propped at formation level.
Geotechnique, Vol. 41, No. 4, pp. 499-514.
204. Powrie, W. & Kantartzi, C. 1996 Ground response during diaphragm wall installation in clay: Centrifuge
model tests. Geotechnique, 46(4), pp. 725739.
205. Puller, M. 2003. Deep excavations: A practical manual. 2nd edition. Thomas Telford Books.
206. Rankin, W.J. 1988. Ground movements resulting from urban tunnelling: predictions and effects. Engineering
Geology of Underground Movements, F.G. Bell et al. (eds), Geological Society, London, pp. 79 - 92.
207. Rossignol P. & Genin M.-J. 1973. Calculs de rideaux de parois moules avec le programme Paroi: exemples
dapplication. Travaux, 465, 1973, pp. 65-67. Quoted from Delattre (2001).
208. Rowson, J., 2009. Cologne: groundwater extraction method probed, New
http://www.nce.co.uk/cologne-groundwater-extraction-method-probed/1995535.article.

Civil

Engineer.

209. Said, R. 1981. The geological evolution of the River Nile. Springler,Verlag.
210. Schuster, M.J., Kung, G.T.C., Juang, C.H.& Hashash, Y.M.A. 2009. Simplified model for evaluating damage
potential of buildings adjacent to a braced excavation. J. Geotech. Geoenviron. Eng., 135(12), pp. 1823
1835.
211. Seok, J.W.; Kim, O.Y.; Chung, C.K. & Kim, M.M. 2001. Evaluation of ground and building settlement near
braced excavation sites model testing. Can. Geotech. J. Vol (38). pp 1127-1133.
212. Shahin, M. A., Jaksa, M. B., & Maier, H. R. 2001. Artificial neural network applications in geotechnical
engineering. Australian Geomechanics, 36(1), pp. 49-62.
213. Shahin, M. A., Jaksa, M. B., & Maier, H. R. 2008. State of the art of artificial neural networks in
geotechnical engineering. Electronic Journal of Geotechnical Engineering, 8, 1-26.
214. Shalaby, A.M. 1989. Geotechnical behaviour of diaphragm walls. Ph. D. Thesis, Geotechnical Engineering,
University of Alberta, Canada.
215. Skempton, A.W. & Macdonald, D.H. 1956. The allowable settlements of buildings. Proc., Inst. of Civ. Engrs.,
1956, Part III, 5, pp. 727-768.
216. Son, M. 2003. The response of buildings to excavation-induced ground movements. Ph.D. thesis, Univ. of
Illinois at Urbana-Champaign, Urbana,
217. Son, M. & Cording, E.J. 2005. Estimation of building damage due to excavation-induced ground movements.
Journal of Geotechnical and Geoenvironmental Engineering, Vol. 131, No.2, pp. 162-177.
218. Son, M. & Cording, E.J. 2007. Evaluation of building stiffness for building response analysis to excavationinduced ground movements. Journal of Geotechnical and Geoenvironmental Engineering (133): 995-1002.
219. Sowers, G.F. 1993. Human factors in civil and geotechnical engineering failures. J. Geotech. Eng., 119(2),
pp. 238256.

162

220. Stallebrass, S.E. & Taylor, R.N. 1997. The development and evaluation of a constitutive model for the
prediction of ground movements in overconsolidation clay. Geotechnique, Vol. 47, No. 2, pp. 235253.
221. Strom, R.W. & Ebeling, R.M. 2001. State of the practice in the design of tall, stiff, and flexible tieback
retaining walls. Report ERDC/ITR-0l-1. U.S. Army Engineer Research and Development Center,
Information Technology Laboratory, Vicksburg, MS, USA. 251 pp.
222. Tamagnan, D. & Beth, M. 2012. Remote monitoring of surface deformation with Robotic Total Stations using
reflectorless measurements (RRTS). Geotechnical Instrumentation News (GIN), March 2012, pp. 30-30.
223. Terzaghi, K. 1943. Theoretical soil mechanics. John Wiley & Sons Inc..
224. Terzaghi, K. 1955. Evaluation of coefficient of subgrade reaction. Geotechnique 5, pp. 297-326.
225. Terzaghi, K. & Peck, R. B. 1967. Soil Mechanics in engineering practice. John Wiley and Sons Inc., 729 pp.
226. Thorley, C.B.B. & Forth, R A. 2002. Settlement due to Diaphragm Wall Construction in Reclaimed Land in
Hong Kong. Journal of Geotechnical and Geoenvironmental Engineering, ASCE, Vol. 128(6), pp. 473-478.
227. Turabi, D.A. & Balla A. 1968. Distribution of earth pressure on sheet-pile walls. Proc. ASCE, Journal of the
Soil Mechanics and Foundations Division, Vol. 94, SM6, pp. 1271-1301.
228. Ukritchon, B.; Whittle, A.J.; & Sloan, S. W. 2003. Undrained stability of braced excavations in clay. Journal
of Geotechnical and Geoenvironmental Engineering, ASCE, Vol. 129(8), pp. 738-755.
229. Van Baars, S. 2011. Causes of major geotechnical disasters. 3rd International Symposium on Geotechnical
Safety and Risk. Munchen, Germany.
230. Van der Poel, J.T.; Gastine, E & Kaalberg, F.J. 2005. Monitoring for construction of a new North/South
metro line in Amsterdam, the Netherlands. Proc. 5th Int. of. Of TC28 ICSNGE. Geotechnical Aspects of
Underground Construction in Soft Ground, pp 8745-751.
231. Van Tol, A.F. 2010. Case study: Amsterdam Metro North-South Line an update on the data obtained and
lessons learned. GE & NCE Basements and Underground Structures Conference 2010, London
232. Vatovec, M.; Kelley, P.; Brainerd, M. & Russo, C.2010. Mitigation of damage to buildings adjacent to
construction sites in urban environments. STRUCTURE magazine, September 2010, pp. 10-12.
233. Weatherby, D. E., Chung, M., Kim, N-K., and Briaud, J-L. 1998. Summary report of research on permanent
ground anchor walls; II, Full-scale wall tests and a soil-structure interaction model. Report FHWA-RD-98066, Federal Highway Administration, McLean, VA.
234. Whittle, A.J. & Davies, R.V. 2006. Nicoll Highway Collapse: Evaluation of geotechnical factors affecting
design of excavation support system. International Conference on Deep Excavations, 28-30 June 2006,
Singapore.
235. Whittle, A.J.; Hashash, Y.M.A. & Whitman, R.V. 1993. Analysis of deep excavation in Boston. Journal of the
Geotechnical Engineering, Vol. 119, No. 1, pp. 6990.
236. Winkler, E. 1867. Die lehrevon elasizital and testigeit (On elasticity and fixity). Prague, Czechoslovakia.
237. Woods, R.I. 2003. The application of finite element analysis to the design of embedded retaining walls. Ph.D.
Thesis. University of Surrey.
238. Yong, K.Y.; Lee, F.H.; Parnploy, U. & Lee, S.L. 1989. Elasto-plastic consolidation analysis for strutted
excavation in clay. Computers and Geotechnics, 8(4), pp. 311-328.
239. Yu, D. & Gang, Z. 2008. Numerical analysis of effect of friction between diaphragm wall and soil on braced
excavation. J. Cent. South Univ. Technol. 15(s2), pp. 81-86.
240. Zapata-Medina, D.G. 2007. Semi-empirical method for designing excavation support systems based on
deformation control. M.Sc. Thesis. University of Kentucky. 468 pp.
241. Zienkiewicz, O.C. & Morice, P. B. 1971. The finite element method in engineering science. McGraw-hill. 521
pp.

163

164

You might also like