You are on page 1of 221

SEISMIC SLOPE SAFETY - DETERMINATION OF CRITICAL SLIP SURFACE USING ACCEPTABILITY CRITERIA

A thesis submitted to the University of London in partial fulfillment for the degree of Doctor of Philosophy and Diploma of Imperial College London by Ding Tan

Department of Civil and Environmental Engineering Imperial College University of London November 2006

Abstract

For the design of earth dams and embankments under earthquake loading, the seismicdisplacement approach provides better criteria than the load-based approach. In order to apply complex models in seismic displacement analysis, it is essential to find the critical slip surface, the critical acceleration and the kinematically acceptable slip mechanism. The objective of current research is to obtain this information. A new procedure is developed here, based on pseudo-static analysis within the limit equilibrium framework, to obtain the slip surface with an acceptable stress field within the surface. The procedure uses stress acceptability as a prerequisite to derive a system of nonlinear equations to determine the slip surface slice by slice upward. An iterative approach is used to overcome the divergence near the corner point of the slope and a slip-path approach is employed to account for the variety of potential failures at the boundaries of different soil layers. In the homogeneous cases without pore water pressures, the equations can be expressed analytically; however for nonhomogeneous slopes or when pore water pressures are considered, the equations can only be solved numerically. The procedure is extensively validated by a series of simulations performed using the Imperial College Finite Element Program with an elastic perfectly plastic Mohr-Coulomb model. In those simulations different rigidity parameters and different initial stress fields are assumed to check the independency of these conditions. If the fully associated condition is adopted, good agreements are achieved between the proposed procedure and finite element analysis, in terms of the critical acceleration, the critical slip surface and the stress distribution along the critical slip surface, both for homogenous and non-homogenous slopes.

Acknowledgements

Firstly I would like to express my sincere gratitude to Dr Sarada K. Sarma for his consistent support, guidance and understanding over the past three years. Without his innovative thought this research would not have been possible. Particular thanks are owed to Prof. David Potts for his Imperial College Finite Element Program and to Dr Lidija Zdravkovic for her assistance and guidance on the finite element analysis conducted for this thesis. I am grateful to Dr John Douglas for his comments during the preparation of the thesis and his kind help during these three years. The financial support provided by the Overseas Research Students Awards Scheme from Universities UK is gratefully acknowledged. I am grateful to all academic staff in the Geotechnics Section of Imperial College London, particularly to Prof. John Burland, Dr. Jamie Standing, Dr. Catherine O'Sullivan and Prof. Julian Bommer for their interesting lectures. Other academic staff and researchers in the department have helped me with many aspects of my study and given life at Imperial a sense of joy, these colleagues are numerous and only a few of them are listed here for brevity: Dr Leroy Gardner, Dr Jon Hancock, Dr Fleur Strasser, Miss Rishmila Mendis, Mr Guillermo Aldama Bustos, Dr Stavroula Kontoe, Miss Victoria Potts, Mr Dave Edwards, Mr Minh Nguyen, Mr William Mannion and Mr Satoshi Nishimura. I would like to thank my parents for their support and encouragement. It is difficult to explain my gratitude with words but I wish to delicate this thesis to them with love from their son.

Table of Contents

Abstract........................................................................................................................................... 2 Acknowledgements ........................................................................................................................ 3 Table of contents ............................................................................................................................ 4 List of figures.................................................................................................................................. 7 List of tables ................................................................................................................................. 10 List of notation ............................................................................................................................. 11 1. Introduction.............................................................................................................................. 14 1.1 background .......................................................................................................................... 14 1.2 factor of safety and critical acceleration in slope stability analysis..................................... 15 1.3 seismic slope safety ............................................................................................................. 16 1.3.1 load-based seismic slope analysis ................................................................................ 16 1.3.2 displacement-based seismic slope analysis .................................................................. 16 1.4 outline of this thesis............................................................................................................. 20 2. Methods used for slope stability analysis............................................................................... 23 2.1 scenario................................................................................................................................ 23 2.2 limit equilibrium method ..................................................................................................... 23 2.2.1 unified frameworks for methods of slices .................................................................... 27 2.2.2 acceptability criteria ..................................................................................................... 30 2.3 method of characteristics ..................................................................................................... 36 2.4 limit analysis........................................................................................................................ 37 2.4.1 rotation mechanism (for example, chen, giger and fang, 1969 and chen and giger, 1971) ............................................................................................................................................... 37 2.4.2 translational mechanism (for example michalowski,1995).......................................... 39 2.5 finite element method .......................................................................................................... 42 3. Optimization techniques used in slope stability analysis to search for the critical slip surface ........................................................................................................................................... 43 3.1 scenario................................................................................................................................ 43 3.2 pattern search scheme.......................................................................................................... 43 3.2.1 circular slip surface (e.g. Bishop and morgenstern, 1960) ........................................... 44 3.2.2 log-spiral slip surface (e.g. Prater, 1979; lighthall, 1979) ............................................ 45 3.2.3 comments on the pattern search scheme ...................................................................... 48 3.3 calculus of variations ........................................................................................................... 49 3.3.1 the application of calculus of variations in slope stability analysis.............................. 49 3.3.2 comments on variation of calculus............................................................................... 51 3.4 calculus-based optimization methods .................................................................................. 52 3.4.1 dynamic programming ................................................................................................. 52 3.4.2 simplex analysis ........................................................................................................... 54

3.4.3 alternating variable method .......................................................................................... 56 3.4.4 mixed optimization techniques..................................................................................... 58 3.4.5 comments on calculus-based optimization................................................................... 60 3.5 random search (monte carlo) techniques ............................................................................. 60 3.5.1 the application of random search in slope stability analysis ........................................ 61 3.5.2 comments on random search technique........................................................................ 68 3.6 new optimization methods................................................................................................... 70 3.7 comments on the optimization techniques used in slope stability analysis ......................... 73 4. Acceptability criteria and formulation of the new procedure ............................................. 74 4.1 scenario................................................................................................................................ 74 4.2 the requirement of acceptability criteria - the basics of limit equilibrium technique for the critical acceleration analysis...................................................................................................... 74 4.3 new procedure and general assumptions ............................................................................. 80 4.3.1 new procedure .............................................................................................................. 80 4.3.2 general assumptions ..................................................................................................... 81 4.4 the analysis of a standard slice ............................................................................................ 81 4.4.1 unknowns of a standard slice ....................................................................................... 82 4.4.2 the acceptability criteria ............................................................................................... 85 4.4.3 the internal plane and m in the standard slice ............................................................. 86 4.4.4 effective normal stress distribution .............................................................................. 89 4.4.5 application of acceptability criterion............................................................................ 91 4.4.6 the definition of pore water pressure............................................................................ 95 4.4.7 line of thrust and moment equilibrium ......................................................................... 99 4.5 the analysis of the first slice .............................................................................................. 100 4.6 the analysis of the last two slices....................................................................................... 102 4.6.1 the geometry and forces acting on the last slice and initial conditions ...................... 102 4.6.2 the application of acceptability criteria into the last two slices.................................. 104 5. The application of acceptability criteria to homogenous slopes ........................................ 107 5.1 scenario.............................................................................................................................. 107 5.2 the first slice ...................................................................................................................... 107 5.3 methods to solve the non-linear equations......................................................................... 110 5.3.1 the determination of the jacobian matrix.................................................................... 111 5.3.2 validation of stress acceptability ................................................................................ 114 5.3.3 the initial approximation of i and i+1 ....................................................................... 114 5.4 acceptable slip surface ....................................................................................................... 117 5.4.1 infinite slopes ............................................................................................................. 117 5.4.2 finite height slopes ..................................................................................................... 118 5.4.3 acceptable critical acceleration................................................................................... 119 5.4.4 initial value of the critical acceleration ...................................................................... 121 5.5 the effect of increment x ................................................................................................. 121 5.5.1 the effect of increment x on the solution of i and i+1 for a single slice ................. 121 5.5.2 the effect of increment x on the critical acceleration of a whole slope.................... 123 5.5 line of thrust....................................................................................................................... 126 5.6 global critical slip surface.................................................................................................. 129 6. The application of acceptability criteria to non-homogenous slopes................................. 132 6.1 scenario.............................................................................................................................. 132 6.2 stress acceptability for a standard slice in non-homogenous slopes .................................. 132 6.2.1 m in a standard slice in non-homogenous slopes ...................................................... 133

6.2.2 validation of stress acceptability and further constraints in non-homogenous slopes 137 6.2.3 method to find the maximum value of m .................................................................. 142 6.3 the increment of x in non-homogenous slopes ................................................................ 147 6.3.1 variable x used to make the slip surface of a single slice stay within the same soil layer..................................................................................................................................... 147 6.3.2 variable x to satisfy kinematical acceptability ......................................................... 150 6.4 accepted slip surface in non-homogenous slopes .............................................................. 150 6.5 slip surface path in non-homogenous slopes ..................................................................... 151 6.6 methodology to determine the critical slip surface............................................................ 153 6.7 practical application........................................................................................................... 154 7. Comparison of the new procedure with finite element method for determining the critical slip surface .................................................................................................................................. 160 7.1 scenario.............................................................................................................................. 160 7.2 interpretation of results from finite element slope stability analysis ................................. 160 7.2.1 interpretation of the results from finite element analysis in terms of factor of safety 161 7.2.2 interpretation of the results from finite element analysis in terms of critical acceleration ......................................................................................................................... 167 7.3 finite element model used to determinate the critical acceleration and the critical slip surface...................................................................................................................................... 167 7.3.1 description of the model............................................................................................. 168 7.3.2 definition of failure..................................................................................................... 169 7.3.3 soil model ................................................................................................................... 169 7.3.4 the effect of the initial stress condition on the computed critical acceleration........... 172 7.3.5 the effect of stiffness on the computed critical acceleration ...................................... 177 7.3.6 the effect of the angle of dilation on the computed critical acceleration.................... 179 7.4 comparison between the new procedure and finite element method in homogenous slopes ................................................................................................................................................. 181 7.4.1 comparison of the computed critical acceleration ...................................................... 181 7.4.2 comparison of the critical slip surface........................................................................ 185 7.4.3 comparison of normal and shear stress distributions along the critical slip surface... 189 7.5 comparison between the new procedure and finite element method in non-homogenous slopes ....................................................................................................................................... 192 8. Conclusions and recommendations ...................................................................................... 203 8.1 conclusions ........................................................................................................................ 203 8.2 recommendations for future study ..................................................................................... 204 References................................................................................................................................... 205 Appendix a. Parabolic effective normal stress distribution ................................................... 212 Appendix b. Analysis for the first slice .................................................................................... 215 B.1 geometry and forces acting on the first slice and initial conditions .................................. 215 B.2 the application of acceptability criterion in the first slice ................................................. 215 Appendix c. Mohr circle analysis for passive wedge............................................................... 219

List of Figures

Figure 1.1 Sliding block model used for the assessment of post-seismic displacements (after Ambraseys and Srbulov, 1995) ............................................................................................. 19 Figure 1.2 Geometry changes and mass transfer in a three-block system with straight initial ground surface (AD). New ground surface = A'F''H'E''E'J'D' (after Chlimintzas, 2002) ...... 20 Figure 2.1 A standard slice in limit equilibrium methods (Fredlund and Krahn, 1977)................ 25 Figure 2.2 Velocity compatibility between adjacent blocks. The left wedge moves upward. (a) Velocities of the blocks. (b) Velocity hodograph. (Donald and Chen, 1997) ....................... 34 Figure 2.3 Directions of normal and shear forces between adjacent slices.................................... 35 Figure 2.4 Geometry of the slope and forces acting on the slope (Sokolovski,1960) ................... 36 Figure 2.5 Logarithmic rotational failure mechanism (Chen and Giger, 1971)............................. 39 Figure 2.6 (a) Translational failure mechanism; (b) single block; (c) velocity vectors; (d) hodograph (Michalowski, 1995) ........................................................................................... 40 Figure 2.7 (a) Hodograph for a homogeneous slope; (b) hodograph when the internal friction on inter-faces between blocks is neglected (Michalowski, 1995).............................................. 41 Figure 3.1 Specification of parameters (Bishop and Morgenstern, 1960) ..................................... 44 Figure 3.2 Detailed contours for a typical case (Bishop and Morgenstern, 1960)......................... 45 Figure 3.3 The logarithmic spiral surface for toe failure (Prater, 1979) ........................................ 46 Figure 3.4 The geometry of log spiral slip surface for non-toe failure (Lighthall, 1979).............. 47 Figure 3.5 Typical contours of kc values (Lighthall, 1979)............................................................ 48 Figure 3.6 Diagram used for Janbus method (Revilla and Castillo, 1977)................................... 50 Figure 3.7 Stages and states used in DP (Baker, 1980) ................................................................. 53 Figure 3.8 The discretization scheme (Baker, 1980) ..................................................................... 55 Figure 3.9An example of Simplex reflection for three dimensional space (Nguyen, 1985).......... 55 Figure 3.10 Shifting points on the slip surface (Celestino and Duncan,1981)............................... 57 Figure 3.11 Description of non-circular slip surface by nodal points (Li and White, 1987) ......... 58 Figure 3.12 Discretization patterns (Chen and Shao, 1988) .......................................................... 60 Figure 3.13Selecting initial line segment (Siegel et al., 1981) ...................................................... 62 Figure 3.14 Direction limits for successive line segments of an irregular surface (Siegel et al., 1981) ..................................................................................................................................... 62 Figure 3.15Sliding generator using more than two boxes (Siegel et al., 1981) ............................. 63 Figure 3.16 The search and confidence areas (Chen, 1992) .......................................................... 64 Figure 3.17Generation of the first slip surface (Husein Malkawi et al., 2001 a)........................... 67 Figure 3.18 Moving point A to the next stage (Husein Malkawi et al., 2001 a)............................ 67 Figure 3.19 Step-by-Step search procedure for rotating three-segment example to minimize factor of safety (Husein Malkawi et al. (2001 b)............................................................................. 69 Figure 3.20 Generation of a non-circular failure surface (Cheng, 2003)....................................... 71 Figure 3.21 Non-circular failure surface used in the Genetic Algorithm (Zolfaghari et al., 2005)72 Figure 4.1 a schematic diagram for a slope and the forces and their acting points in one slice .... 75 Figure 4.2 An example slope and its soil properties ...................................................................... 78 Figure 4.3 (a) The variations of E resulting from different values of kc (b) The variations of local factors of safety resulting from different values of kc ........................................................... 79 Figure 4.4 Forces acting on an inclined slice (The total forces N and E are shown, which are the sums of the effective force and the force due to pore water). ............................................... 83 Figure 4.5 The internal plane and forces acting on it..................................................................... 86

Figure 4.6 Geometry of the segment ABA in the slip surface side................................................ 87 Figure 4.7 Geometry of segment BDD ......................................................................................... 89 Figure 4.8 Forces acting on the first slice. The total forces N and E are shown, which are the sums of the effective force and the force due to pore water......................................................... 101 Figure 4.9 Forces acting on the last two slices. The total forces N and E are shown, which are the sums of the effective force and the force due to pore water................................................ 102 Figure 5.1 The geometry of a passive wedge on the sloping ground........................................... 108 Figure 5.2 Two sets of angles 1 and 2 corresponding to different values of H, N represents the solution from the new procedure and M represents the solution from the Mohr-circle solutions .............................................................................................................................. 110 Figure 5.3 A general slice within an example slope .................................................................... 112 Figure 5.4 The variation of m in the twentieth slice in Example 5.2.......................................... 114 Figure 5.5 The variation of i and i+1 in the first twenty slices in the Example 5.2 ................... 115 Figure 5.6 Solutions from two different initial approximations of i and i+1 in the second slice116 Figure 5.7 The relationship between vertical distance and desired critical acceleration in the example infinite slope ......................................................................................................... 117 Figure 5.8 The different slip surfaces produced by different values of kc ................................... 119 Figure 5.9 The values of kc and kgend at each iteration................................................................. 120 Figure 5.10 The resulting point B by assuming different x....................................................... 123 Figure 5.11 Lines of thrust resulting from different assumed stress distribution along the interslice boundaries................................................................................................................... 127 Figure 5.12 Points of application of Ni and Ei+1 resulting from linear stress distribution.......... 128 Figure 5.13 Points of application of Ni and Ei+1 resulting from parabolic stress distribution .... 128 Figure 5.144 Flowchart of determining the critical slip surface in homogenous slopes.............. 130 Figure 5.155 Local critical slip surfaces and the global slip surface ........................................... 131 Figure 6.1 A general slice in two soil layers (a stronger soil layer overlaying a weaker soil layer) ............................................................................................................................................. 138 Figure 6.2 A general slice in two soil layers (a weaker soil layer overlaying a stronger soil layer) ............................................................................................................................................. 139 Figure 6.3 The variation of m in the slice in Example 6.1 ......................................................... 140 Figure 6.4 The variation of m in the slice in Example 6.2 ......................................................... 141 Figure 6.5 The algorithm of solving a single slice within non-homogenous slopes.................... 143 Figure 6.6 Variable x to make the slip surface of a single slice stay within the same soil layer148 Figure 6.7 Slip surface path. Note that different surfaces encounter the first soil layer boundary at three different points near B but the differences are too small to be seen........................... 152 Figure 6.8 (a) and (b) Comparison of slip surfaces of four-layer slope given by the new procedure and that obtained by Zolfaghari et al. (2005) using a genetic algorithm approach ............. 155 Figure 6.9 Comparison of slip surfaces of a four-layer slope estimated using different methods158 Figure 6.10 Internal factors of safety on vertical planes within surfaces obtained: (a) by the new procedure; (b) by Yamagami & Ueta (1988b); and (c) by Greco (1996)............................ 159 Figure 7.1 Model used for homogenous slopes ........................................................................... 168 Figure 7.2 The horizontal (a) and vertical stress (b) distributions after the application of gravity ............................................................................................................................................. 174 Figure 7.3 Mesh for the excavation from level ground (1:3 slope) ............................................. 174 Figure 7.4 The horizontal (a) and vertical (b) stress distributions after excavation from level ground (K0 = 1.0) ................................................................................................................ 175 Figure 7.5 The effect of initial stress on the computed critical acceleration ............................... 176 Figure 7.6 Comparison of critical slip surfaces between the new procedure and finite element method under different initial stresses: (a) gravity turned on (b) K0= 0.658 and (c) K0 = 1.5 ............................................................................................................................................. 178

Figure 7.7 The effect of stiffness on the computed critical acceleration ..................................... 180 Figure 7.8 Effect of angle of dilation on the computed critical acceleration ............................... 181 Figure 7.9 Comparison of critical acceleration between the new procedure and finite element method for slope angles: (a) 1:5 (b) 1:3 and (c) 1:1.5......................................................... 183 Figure 7.10 Comparison of critical slip surface between finite element analysis and the new procedure for slope 3: (a) = (b) = 0.5 (c) = 0 ................................................... 187 Figure 7.11 Comparison of critical slip surface between finite element analysis and the new procedure for slope 12: (a) = (b) = 0.5 (c) = 0 ................................................. 188 Figure 7.12 Comparison of normal stress (a) and shear stress (b) distributions for slope 3 ........ 190 Figure 7.13 Comparison of normal stress (a) and shear stress (b) distributions for slope 13 ...... 191 Figure 7.14 Slope profile used in Example 7.4............................................................................ 193 Figure 7.15 Mesh of the slope after excavation ........................................................................... 193 Figure 7.16 Comparison of slip surfaces in Example 7.4 ............................................................ 194 Figure 7.17 Comparison of the normal stresses (a) and shear stresses (b) on the slip surface in Example 7.4......................................................................................................................... 195 Figure 7.18 Slope profile used in Example 7.5............................................................................ 196 Figure 7.19 Comparison of slip surfaces in example 7.5............................................................. 197 Figure 7.20 Comparison of the normal stresses (a) and shear stresses (b) on the slip surface in Example 7.5......................................................................................................................... 198 Figure 7.21 (a) The mesh before, and (b) after excavation.......................................................... 200 Figure 7.22 Comparison of slip surfaces in Example 7.6 ............................................................ 201 Figure 7.23 Comparison of the normal stresses (a) and shear stresses (b) on the slip surface in Example 7.6......................................................................................................................... 202 Figure C.1 Mohr circle analysis for the first slice........................................................................ 221

List of Tables

Table 2.1 Summary of equations and unknowns associated with limit equilibrium methods ....... 25 Table 2.2 Equilibrium conditions and assumptions in various limit equilibrium methods............ 26 Table 5.1 Comparison of results of each iteration between analytical Jacobian matrix and finite difference Jacobian matrix .................................................................................................. 113 Table 5.2 The effect of increment x on the values of i and i+1 for an arbitrary slice.............. 122 Table 5.3 The effect of increment x on the computed critical acceleration in three example slopes................................................................................................................................... 125 Table 6.1 (a) Soil properties (b) Initial conditions and details of solved slice in Example 6.2 ... 139 Table 6.2 (a) Soil properties (b) Initial conditions and details of solved slice in Example 6.3 ... 140 Table 6.3 Geometry of the slice, weighted average soil parameters and maximum value of m in each iteration when is taken as: (a) 20, (b) 50 and (c) 100 .............................................. 144 Table 6.4 Soil properties in Example 6.3..................................................................................... 148 Table 6.5 Geometry of the new slice assuming different x ....................................................... 149 Table 6.6 Soil properties in Example 6.4..................................................................................... 151 Table 6.7 Soil properties in Example 6.6..................................................................................... 154 Table 6.8 Soil properties in Example 6.7..................................................................................... 156 Table 7.1 Homogenous slopes analysed by the finite element method ....................................... 184 Table 7.2 Soil properties in Example 7.4..................................................................................... 192 Table 7.3 Soil properties in Example 7.5..................................................................................... 196

10

List of Notation

c'

cohesion in terms of effective stresses on slip surface average cohesion in terms of effective stresses on the inter-slice boundary average cohesion in terms of effective stresses on the internal plane length of the internal plane in the inter-slice boundary side normal force acting on the inter-slice in terms of effective stresses implied normal force in terms of total stresses acting on the internal plane within inter-

c' c'
D E' E

slice boundary side EL normal force in terms of total stresses acting on the small section on the inter-slice

boundary F Fv kc L li l_pos mS mB factor of safety factor of safety on the vertical inter-slices critical acceleration coefficient length of the internal plane in the slip surface side the point of application of Ni measured along the slip surface the percentage of the application points Ni of the corresponding slip plane of the slice the reciprocal of the factor of safety of the internal plane within slip surface side the reciprocal of the factor of safety of the internal plane within inter-slice

boundary side N' N side PW force due to water pressure on the inter-slice boundary normal force on the base of slice in terms of effective stress implied normal force in terms of total stresses on the internal plane within slip surface

11

PW PWL Ru T T U U U Wi W X X XL zi z_pos axis

force due to pore water pressure on the internal plane within inter-slice boundary side force due to pore water acting on the small section on the slice boundary pore pressure ratio shear force on the base of slice implied shear force on the internal plane within slip surface side force due to water pressure on the base of slice force due to pore water pressure on the internal plane within slip surface side pore water pressure weight of i th slice weight of the small segment inside the slice shear force on the inter-slice boundary implied shear force on the internal plane within inter-slice boundary side shear forces on the small section on the slice boundary the point of application of Ei measured along the inter-slice boundary the percentage of the application point Ei of the corresponding inter-slice plane angle made by slip surface with horizontal, positive anticlockwise from the positive x

tan kc angle made by inter-slice boundary with vertical, positive clockwise measured from the

positive y axis. h horizontal displacement of the corner point of the slope under the application of

horizontal acceleraion

' ' '

friction angle in terms of effective stresses on slip surface average friction angle in terms of effective stresses on the inter-slice boundary average friction angle in terms of effective stresses on the internal plane

12

angle made by the internal plane with the slip surface or the inter-slice boundary unit weight of soil convergence number normal stress in the intersection between i th side of slice and slip surface normal stress in the intersection between i th side of slice and ground surface shear stress two definitions: the inclination angle of a slope in the new procedure and angle of

'i 'i0

dilation in the finite element method

increment in the slope surface

13

1. Introduction

1.1 BACKGROUND
Landslides have been one of the major causes of damage and casualties in earthquakes. For example, landslides caused more than half of the economic losses in the great 1964 earthquake (M=9.2) in Alaska (Keefer, 1984) and killed more than one hundred thousand people during an earthquake (M=7.8) in China in 1920 (Wang and Xu, 1984). The 1994 Northridge earthquake (M=6.7) is a well documented earthquake to perform a detailed analysis of the factors related to the seismic triggering of landslides. This earthquake caused widespread damage and huge economic losses. Though of moderate magnitude, this is the most costly earthquake in US history, with losses estimated at more than $30 billion (Parise and Jibson, 2000). More recently, the 1999 Chi-Chi earthquake (M=7.3) triggered more than 10,000 landslides of various types in the steep mountainous terrain of Central Taiwan, throughout an area of about 11,000 km2 (Khazai and Sitar, 2004). Understanding the failure mechanism of landslides during earthquakes is important for reducing damage and loss of life in future earthquakes. However, the design, or evaluation, of new, or existing, dams, embankments and slopes to safely resist the destructive effects of earthquakes constitutes a complex problem. Strong ground motions during earthquakes induce large inertia forces in earth dams and embankments, which together with the initial stresses may bring them to failure. It is also known that pore water pressures within the soil mass of an earth dam or embankment may increase during earthquake loading. This increase of pore water pressure will reduce the strength of the material, which may lead to failure. In the extreme situation, this increase of pore water pressure may produce liquefaction within the soil mass. Therefore, a simple and rational procedure is requirement for practicing engineers to obtain qualitative solutions for a large number of analyses to judge whether the slope is stable under earthquake conditions.

14

1.2 FACTOR OF SAFETY AND CRITICAL ACCELERATION IN SLOPE STABILITY ANALYSIS


The factor of safety of a slope is defined as the ratio between the available strength and the strength required for a state of incipient failure along a possible slip surface. In terms of the limit equilibrium principle, a factor of safety less than one represents failure in a slope. This means that when the factor of safety is less than one, a section of the dam or embankment will slide along the failure surface and only come to rest again at a place where the new stresses do not exceed the available strength, due to the change of the geometry and soil properties. It is therefore obvious that a factor of safety less than one cannot be permitted under static conditions. In seismic slope stability analysis, a conventional procedure, pseudo-static slope stability analysis, can be adopted, where acceleration is applied to the slope: it is assumed to be constant over the whole slope and to act in a horizontal direction. The vertical component of earthquake acceleration can be easily accommodated by simply changing the unit weight of materials to take care of the additional vertical acceleration, then finding the horizontal critical acceleration for the modified unit weight and then the resulting critical acceleration can be adjusted for the modified unit weight. If water exists in the slope, the unit weight of water should also be modified accordingly. Under the pseudo-static condition, a factor of safety can be defined in the same way as in the static case. However, a better parameter to describe the seismic safety of a slope is the critical acceleration. The critical acceleration is defined as the acceleration that, when applied to the mass between the slip and the slope surfaces, produces a state of incipient failure along that surface. The idea of the critical acceleration in seismic slope stability analysis is that if the applied acceleration is larger than the critical, then the mass of soil will move along the slip surface. It is noted that the factor of safety (or critical acceleration) is evaluated for a given slip surface while the minimum factor of safety (or critical acceleration) is a characteristic of the slope.

15

1.3 SEISMIC SLOPE SAFETY

1.3.1 Load-based seismic slope analysis One of the earliest procedures for seismic slope stability analysis is the load-based procedure, in which the earthquake loading is represented by a horizontal static force, equal to the soil weight multiplied by a coefficient, which can be estimated by empirical guidelines or codes (e.g., Seed, 1977). The pseudo-static force is then integrated in a conventional limit equilibrium slope stability analysis and the factor of safety is computed. The computed factor of safety provides an indication of the possible magnitude of seismically induced displacement (Makdisi and Seed, 1978). However, load-based seismic slope analysis produced factors of safety well above 1 for a number of dams that later failed during earthquakes, such as Sheffield Dam and Tailings dam (Kramer, 1996). It shows that the load-based design has not reached an acceptable level of competence in seismic slope stability analysis.

1.3.2 Displacement-based seismic slope analysis An alternative approach is to assess the exact amount of movement that the slope would undergo under a specific seismic load. Dynamic analysis should be used to simulate the movement of the slope in the time domain during or immediately after the imposed ground motion. This approach forms the concept of displacement-based analysis, which can also be implemented in simplified methods (Chlimintzas 2003). Advantages of displacement-based analysis As stated by Chlimintzas (2002), when compared to load-based analysis, displacement-based techniques constitute a more realistic way of estimating likely hazard for the following reasons: (1) Failure and damage can be easily quantified within a model using displacements; such quantities provide a much more comprehensive representation of the real case than does the use of the critical acceleration or factor of safety alone.

16

(2) The identification of damage in the field is related to the induced movements, unlike factor of safety (or critical acceleration) which can not be observed and evaluated in-situ. (3) Depending on the nature and requirements of a specific earth-structure, different displacement limits may be regarded as acceptable; and by employing a displacement-based method of analysis, specific tolerance for each particular situation can be employed; (4) During strong earthquake motions, transient drops of factor of safety below unity may occur without necessarily inducing severe damage or failure. The overall displacement at the end of the earthquake may be extremely small and tolerable and the final factor of safety of the slope may be well above unity.

Types of earthquake-induced displacements


With the development of more sophisticated analysis, deformations caused by earthquakes in natural and man-made slopes are identified in three consequent stages (Ambraseys and Srbulov, 1995). In the first stage, which is co-seismic, gravity as well as seismic forces can generate a failure surface or activate a pre-existing slip surface causing permanent displacements of the slope. Coseismic displacements are usually small. The second stage, which is post-seismic, follows immediately after the earthquake if the fast residual undrained shear strength on the slip surface is less than that required to maintain static equilibrium. In this case, downward displacements will continue with an outward movement of the toe. Consequently, the mass comes to rest at a time when the velocity of the movement becomes zero and in a new position where its factor of safety will be greater than one. Gravity is the only driving force during this second stage, while resisting forces will depend on the residual undrained strength available on the slip surface generated during the first stage, and on the resistance due to the effect of the toe. In the third stage, further movements may develop as a result of creep and consolidation, as well as from destabilising hydrostatic forces if tension cracks produced by the earthquake are filled with water. Additional movements may continue to occur, which will be slow and associated with progressive failure and with the drained strength of the soil.

17

Development of displacement-based analysis


A brief review of the development of displacement-based analysis is presented below. Newmark (1965) Newmark (1965) firstly proposed a relatively simple rigid sliding block analysis. In this approach the displacement of a mass of soil above a slip surface is represented as a rigid block sliding on a plane surface. When the ground motion exceeds the critical acceleration, the block begins to slip along the plane. Movements continue until the velocity of the block relative to the underlying mass becomes zero. The block will slide again if the ground motion exceeds the critical acceleration. To compute displacements, the acceleration in excess of the critical acceleration is integrated once to obtain the velocity and a second time to compute the displacement. In this analysis the probable displacement can be estimated if the expected ground motion and the critical acceleration are known. The accurate identification of the minimum critical acceleration is essential for the application of the Newmark method (1965) for the computation of co-seismic displacements. Charts of critical accelerations, such as those by Pratter (1979) and Lighthall (1979), have been developed for homogeneous slopes similar to those used to derive the stability charts for the factor of safety. A more robust procedure for the determination of the minimum critical acceleration, both for homogenous slopes and non-homogenous slopes, is needed. The single block model is directly applicable in cases of planar translational failures (using Newmark's formulation) or circular rotational failures (Sarma, 1981). One of the difficulties in the application of the single block model in real situations is that the failure surface is generally non-planar and non-circular. Also, since the change of geometry is not taken into account in single block slipping, there is no change in the critical acceleration (other than that due to the change in pore water pressure), the computed large displacements are conservative. Ambraseys and Srbulov (1995) A two-block sliding model is employed by Ambraseys and Srbulov (1995) for the calculation of displacements in the post-seismic stage, immediately after the earthquake. It simulates the motion

18

of a slide that can be approximated by two plane shear surfaces with minimal internal disruption. In this stage movements will begin with a static factor of safety less than unity, and a section of the slope will move under gravity to a new position of equilibrium. It is assumed that the slope has a finite length and that the mass slips on a surface produced by the earthquake during the first stage. It is also assumed that the slip surface consists of two slip planes: one of length L with a depth h parallel to the slope and inclined to the horizontal at an angle , and a second plane of initial length b, which is inclined at an angle (Figure 1.1). This is a model of two continuous blocks, free to slide on two plane surfaces, separated by an inter-slice plane at an angle to the horizontal. This model takes into account the internal deformations and uses the concept of mass transfer between blocks.

Figure 1.1 Sliding block model used for the assessment of post-seismic displacements (after Ambraseys and Srbulov, 1995)

Chlimintzas (2002) There are certain limitations in Ambraseys and Srbulovs (1995) model: it is limited to two blocks; only the post-seismic movement are analysed while the co-seismic sliding is not considered; simplified geometry is used to represent the failure surface; and it does not explicitly

19

consider a geometry transformation rule during sliding. Therefore, a more general multi-block dynamic model was developed by Chlimintzas (2002) for the calculation of sliding displacements using an approach similar to, but more general, than that of Ambraseys and Srbulov (1995). The model relies on the assumption that the pseudo-static limiting equilibrium conditions could provide the failure mechanism on which to construct and configure the dynamic system. A failure mechanism is needed for the application of the multi-block dynamic model and if the slope starts to move, the failure mechanism has to be kinematically acceptable. The main geometry changes that are accommodated by the model are: first, mass transfer from one slice to another; second, it results in changes of the cohesion forces due to variations in interslice boundaries. The mass transfer and internal deformation that result from the displacement of the slope can be thought of as constituting a geometry transformation. The transformation rule used in the model to produce the deformed geometry of the slope is briefly described in Figure 1.2 for a simple three block system.

c J b E F" F' A u1 B H' F H E C C C E I u3 D

D u3

G B u2

Figure 1.2 Geometry changes and mass transfer in a three-block system with straight initial ground surface (AD). New ground surface = A'F''H'E''E'J'D' (after Chlimintzas, 2002)

1.4 OUTLINE OF THIS THESIS


It is shown above that to apply the multi-block sliding model, a kinematically acceptable failure mechanism needs to be known, including the critical slip surface and kinematically acceptable inter-slice boundaries. Optimization techniques can be used to determine the most likely of the

20

admissible mechanisms, i.e. the one that has the lowest critical acceleration. However, there are two criteria that must be satisfied if thus solution obtained is acceptable. The first is a criterion in terms of stresses and the second is a criterion in terms of kinematics. In existing slope stability analysis methods within the limit equilibrium framework, the stress acceptability criterion is not integral to the method, rather it is only sometimes checked after the solution is obtained, and quite often it is found to be violated for some parts of the slip surface and the inter-slice boundaries. It is usually left to the experience of the engineer to declare whether the solutions are acceptable or not. A new method of determining the critical slip surface in slope stability analysis is developed in this thesis, based on the limit equilibrium technique with added stress acceptability criterion. The procedure uses stress acceptability as a prerequisite to derive a system of non-linear equations to determine the slip surface slice by slice upwards; no prior assumption of the shape of the surface is needed. Throughout the thesis, the new procedure refers to the procedure proposed here. The seismic analysis of slopes is an undrained problem. The problem can be analysed using a total stress analysis with undrained strength parameters, or using effective stress analysis with c and . The method developed here can be applied in either total or effective stress analysis. However, the examples given in this thesis are formulated in terms of drained (effective stress) parameters. The practical application of the method depends on incorporating the excess pore water pressure generated under seismic loading. The method proposed by Sarma (1975) can be taken into the current framework to estimate the excess pore water pressure. However, the excess pore water pressure generated under seismic loading is beyond the scope of the current research. Chapter 2 reviews some of the existing slope stability analysis methods for the evaluation of factor of safety or critical acceleration, including the limit equilibrium method, limit analysis, method of characteristics and finite element method. Chapter 3 discusses some of the existing optimization techniques, which are used to identify the minimum factor of safety (or critical acceleration) and the associated critical slip surface within the limit equilibrium framework. Chapter 4 provides the formulation of the proposed procedure. This chapter starts by showing that for a given possible slip surface, inter-slice force distributions exist that will produce any desired

21

critical acceleration as a solution, while they may or may not be acceptable. This is followed by the presentation of acceptability criteria and their application to a general slice. The special considerations for the first slice and the last two slices are presented separately. Chapter 5 demonstrates the application of the new procedure to homogenous slopes. Chapter 6 presents the application of the new procedure to non-homogenous slopes. Chapter 7 begins with a review of some existing research applying finite element analysis to slope stability problems. Finite element analysis is then used as an independent tool to validate the new procedure, in terms of the critical acceleration, the failure mechanism and the stress distribution along the critical slip surface for both homogenous and non-homogenous slopes. Chapter 8 concludes and summarizes the findings from this study and gives recommendations for future study.

22

2. Methods Used for Slope Stability Analysis

2.1 SCENARIO
A quantitative assessment of the stability of a slope is important when a judgement is needed about whether the slope is stable. Once soil shear strengths, pore water pressures, slope geometries and other soil and slope properties have been investigated, an appropriate method has to be selected to perform slope stability calculations, obtaining the factor of safety or critical acceleration. In the literature, there are a number of methods available to perform slope stability analysis: the limit equilibrium method, the method of characteristics, the limit analysis and the finite element method. Their basic features, advantages and limitations are reviewed in this chapter.

2.2 LIMIT EQUILIBRIUM METHOD


Limit equilibrium methods have been the most widely used methods for slope stability analysis (Duncan, 1996). The limit equilibrium methods generally satisfy force and moment equilibrium and boundary conditions and the failure criterion along the slip surface. Within the limit equilibrium framework, methods of slices are extensively used for dealing with complex slope geometry, variable soil properties and the existence of pore water pressure. A standard slice is presented in Figure 2.1 and the unknowns and available equations for a whole slope are listed in Table 2.1. It is shown in Table 2.1 that the limit equilibrium method is a statically indeterminate problem (except for one wedge analysis). During the last century, more than ten methods of slices have been developed: Ordinary method (1936), Simplified Bishop method (1955), Simplified Janbu method (1956), Corps of Engineers method (1967), Rigorous Janbu method (1954, 1973), Spencer method (1967), Morgenstern-Price method (1965), Sarma vertical-slice method (1973), Sarma inclined-slice method (1979) and

23

others. These methods differ in the formulations employed in deriving the factor of safety (or critical acceleration) equations but their main differences are due to the assumptions used (Table 2.2). There are two kinds of solutions. The first is a simplified solution where the conditions of static equilibrium are not rigorously satisfied. Assumptions are made to obtain the solution in a simple form. The second is a rigorous solution where the equilibrium conditions are completely satisfied (Sarma, 1979). This is not meant to be an exhaustive literature review on limit equilibrium methods since only the most commonly used methods are discussed. All methods use vertical slices except for Sarma method (1979), which uses inclined non-parallel slices. The basic features and the assumptions of the above methods can easily be found in a textbook on slope stability and therefore, are not presented in this thesis. It is noted that methods originally developed for the computation of factors of safety can be used for the calculation of critical accelerations and vice versa. The common features of limit equilibrium methods are as follows (Zhu et al., 2003) 1) The sliding body above an assumed slip surface is divided into a number of vertical (or inclined) slices. 2) The strength of the slip surface is mobilized by the same factor of safety, where the cohesion component and the friction component of the strength are reduced equally. There are methods in the literature based on variable factors of safety (e.g., Chugh, 1986) but they are beyond the scope of the current study. 3) Assumptions regarding inter-slice forces are employed to render the problem determinate. Note that not all limit equilibrium methods satisfy all equations of equilibrium (Table 2.2). 4) The factor of safety or critical acceleration is derived from force or/and moment equilibrium equations. In their original formulation, the Ordinary and Simplified Bishop methods are derived only for circular slip surfaces but they can be extended to arbitrary slip surfaces. Some methods are derived based on infinitesimal widths and then the factor of safety equations take the form of definite integrals (Janbu, 1954; Morgenstern-Price, 1965). However, it is difficult to write such equations for variable geometries and soil and pore water conditions for practical problems. Most other methods of slices derive factor of safety equations based on slices with discrete widths. The

24

difference in formulation is merely one of mathematical preference. The new procedure presented in this thesis is based on Sarma inclined-slice method (1979) and slices with discrete widths

Figure 2.1 A standard slice in limit equilibrium methods (Fredlund and Krahn, 1977) Table 2.1 Summary of equations and unknowns associated with limit equilibrium methods assuming that the number of slices be n Number of equations n n n n 4n Number of unknowns n n n-1 n-1 n-1 n 1 6n-2 Type of equations Horizontal force equilibrium Vertical force equilibrium Moment equilibrium Mohr-Coulomb failure criterion at the base of a slice Total number of equations Type of unknowns Total normal force at the base of a slice, P Shear force at the base of a slice, Sm Inter-slice total normal force, E Inter-slice shear force, X Point of application of the inter-slice total normal force Point of application of the total normal force at the base of a slice Factor of safety or critical acceleration Total number of unknowns

25

Table 2.2 Equilibrium conditions and assumptions in various limit equilibrium methods Horizontal force equilibrium No No Yes Yes Yes Yes Yes

Methods Ordinary method Simplified Bishop Method Simplified Janbu method Corps of Engineers Method Rigorous Janbu Method Spencer method Morgenstern-Price Method

Vertical force equilibrium Yes Yes Yes Yes Yes Yes Yes

Moment equilibrium Yes Yes No No Partially** Yes Yes

FF

FM Yes Yes

Assumption Inter-slice forces are neglected Resultant inter-slice forces are horizontal (i.e. there are no shear forces between slices) Resultant inter-slice forces are horizontal (i.e. there are no shear forces between slices); an empirical correction factor, f0, is used to account for inter-slice shear forces. Direction of the resultant inter-slice force is equal to the average slope from the beginning to the ground surface. Location of the inter-slice normal force is defined by an assumed line of thrust. Resultant inter-slice forces are at a constant angle throughout the sliding mass. The shear force between slices are related to the normal forces as X = f(x)E. The proportion of the function, , required to satisfy moment and force equilibrium is computed. Inter-slice shear forces is related to the inter-slice shear strength, Sv, by X = f(x)Sv; inter-slice shear strength depends on shear strength parameters, pore water pressure, and the horizontal component of inter-slice forces. The shear strength is assumed to be mobilized on the boundaries of all slices. The inclination of slice interfaces is varied to produce a critical condition.

Yes Yes Yes Yes Yes Yes Yes

Sarma (1973)

Yes

Yes

Yes

Yes

Yes

Sarma (1979)
*

Yes

Yes

Yes

Yes

**

FF means that factor of safety is derived from overall force equilibrium and FM means that factor of safety is derived from overall moment equilibrium The method is not technically a rigorous solution; moment equilibrium on the last slice is not satisfied.

26

2.2.1 Unified frameworks for methods of slices For a given slope to be analysed, different values of factor of safety (or critical acceleration) can be computed, depending on which method of slices is adopted and what assumption is being made. In the past, extensive research has been carried out to identify these differences and then ascertain which method is more appropriate in different circumstances. Later on, some unified formulations have been proposed to incorporate different methods into one framework. Some of this work is reviewed in the following section. Fredlund and Krahn (1977) Figure 2.1 shows an arbitrary slice within a general slope with the associated variables. A unified framework has been proposed by Fredlund and Krahn (1977), recognizing six methods of slices in the literature whether moment and/or force equilibriums are explicitly satisfied. According to Fredlund and Krahn (1977), all methods of slices satisfying overall moment equilibrium can be written in the same form as:

Fm =

clR + (P ul )R tan Wx Pf + kWe + A a A a


L L

R R

+ Ld

(2.1)

All methods satisfying overall force equilibrium have the following form for the factor of safety equation:

Ff =

cl cos + (P ul ) tan cos P sin + kW + A A L cos


L R

(2.2)

Therefore, it is possible to view the six methods (given below) in terms of factor of safety based on moment equilibrium, factor of safety based on force equilibrium or factor of safety based on both equilibriums. The normal force at the base of each slice is derived either from summation of forces perpendicular to the base or from the summation of forces in the vertical and horizontal directions. The difference of the normal force, P, in different methods comes from their basic assumption (Table 2.2).

27

For the Ordinary Method:

P = W cos kW sin
For the Simplified Bishop Method:

(2.3)

c l sin ul tan sin + P = W / m F F


where

(2.4)

m = cos + (sin tan ) / F


For Spencers Method:

(2.5)

c l sin ul tan sin P = W (E R E L ) tan + / m F F


For Janbus Simplified Method:

(2.6)

c l cos ul tan sin P = W + / m F F


For Janbus Rigorous Method:

(2.7)

c l sin ul tan sin P = W ( X R X L ) + / m F F


For Morgenstern and Price Method:
c l sin ul tan sin P = W ( X R X L ) + / m F F

(2.8)

(2.9)

28

Then it is possible to view the analytical aspects of slope stability in terms of one factor of safety satisfying overall moment equilibrium and another satisfying overall force equilibrium. The rigorous method iterates the procedure to obtain an identical factor of safety, which satisfies overall moment equilibrium and force equilibrium. Fredlund and Krahn (1977) conclude that the factor of safety with respect to moment equilibrium is relatively insensitive to the inter-slice force assumption. Therefore, the factors of safety computed by the Spencer and Morgenstern-Price method are generally similar to those obtained by the simplified Bishop method. However, the factors of safety based on overall force equilibrium are far more sensitive to the inter-slice force assumption. No acceptability criteria, in terms of stress and kinematical acceptability, are introduced into the proposed unified framework. Espinoza, Bourdeau and Muhunthan (1994) A unified formulation is proposed by Espinoza et al. (1994) and eight existing limit equilibrium methods have been grouped into three categories based on the hypotheses used to describe shear forces along the inter-slice boundaries: (1) the direction of the internal forces; (2) the height of the line of thrust; (3) the shape of the distribution function of the internal shear forces. Based on several case studies, Espinoza et al. (1994) conclude that for circular slip surfaces, the value of factor of safety is insensitive to the choice of a particular assumption, however, for noncircular slip surfaces, depending on the shape of the slip surface, the assumption chosen may have a significant influence on the computed factor of safety. However, they emphasise that the comparison made is for chosen slip surfaces and not for critical slip surfaces. Again, no acceptability criteria are employed within their unified framework. Zhu, Lee and Jiang (2003) Zhu et al. (2003) propose a general framework for almost all the existing limit equilibrium methods for slope stability analyses with general slip surfaces. All the methods of slices in Table 2.2 have been included in their work. In this framework, the equilibrium equations are derived in terms of factors of safety and the initially assumed normal stress distribution over the slip surface

29

is multiplied by a modification function involving two auxiliary unknowns. The various assumptions regarding the inter-slice forces can be transformed into the unified form of the expression for the normal stress distribution on the slip surface. Zhu et al. (2003) conclude that the factors of safety associated with the non-rigorous methods of slices are less justifiable than those associated with the rigorous methods. All rigorous methods give almost identical results both for circular slip surfaces and general slip surfaces. Zhu et al. (2003) believe that for circular slip surfaces, although the simplified Bishop method is not rigorous, it is still applicable owing to its simplicity and since it gives nearly identical solutions to rigorous methods. For non-circular slip surfaces, the Morgenstern & Price method and the Spencer method are most applicable since they provide consistent factors of safety and involve very few numerical difficulties. However, because only vertical slices are used for the comparison of different methods, their conclusion that Sarma inclined-slice method (1979) gives a slightly higher factor of safety than other rigorous methods is not appropriate. Zhu et al. (2003) state that for the same slope, different methods of slices could yield different locations for the critical slip surfaces. They also conclude that the factors of safety with unreasonable internal stresses often differ to those with reasonable internal stresses, provided complete equilibrium conditions are satisfied. The acceptability criteria have been used and are detailed in the following section. The moment equilibrium condition requires the location of application of the normal force on the slip surface. The usual assumption of middle of the slice is a good and reasonable one. However, this is still an assumption and one can use any other reasonable point. The factor of safety will depend on this assumption as well, which is not discussed in any of the above unified frameworks. It is important to note that with reasonable internal stresses, the differences are small but with unreasonable internal stresses, the differences may be large. This shows the importance of the acceptable internal stresses.

2.2.2 Acceptability criteria


The results obtained from limit equilibrium methods generally satisfy the equilibrium, boundary conditions and failure criterion along the slip surface. However, to be physically acceptable,

30

stress acceptability and kinematical acceptability have also to be satisfied. However, stress acceptability is not an inherent requirement in the existing limit equilibrium methods. Stress acceptability The stress acceptability criterion exists because the slice is made up of soil, which has a limited strength. The forces on the slice imply stresses inside the soil mass. Because of the limited strength of soil, the implied stresses cannot exceed its limited strength. In other words, the factor of safety in any plane within the slice must be greater than or equal to one. With the limit equilibrium method with slices, for a given potential slip surface, we can compute the normal and shear forces in the inter-slice plane and compute the local factors of safety. Strictly speaking, all the local factors of safety in inter-slice planes should be larger than unity in order to satisfy stress acceptability. If not, the potential slip surface is deemed to be physically unacceptable and a new slip surface should be generated and evaluated. However, for a given potential slip surface, different sets of normal and shear forces on the interslice planes can be computed based on the assumed f(x) ( f(x) defines the ratio between the normal force and shear force on the inter-slice boundaries as shown in Table 2.2 and can take any prescribed form in principle) for some limit equilibrium methods, such as Morgenstern and Price method and Sarma method (1973). The function f(x) may be varied to obtain acceptable normal and shear forces on the inter-slice planes for a specified potential slip surface. However, there is no standard procedure to obtain physically acceptable solution by changing f(x) and the success depends on trial and error. There is no guarantee that a satisfactory f(x) can always be found. The acceptability criterion has been used by different researchers in the literature to judge the computed solution from limit equilibrium methods. The basic features of some attempts are reviewed in the following section. Morgenstern and Price (1965) Morgenstern and Price state that the implied state of stress within the soil mass must be physically possible. In particular, the failure criterion within the soil mass above the slip surface must not be violated. This is perhaps the first time that an internal factor of safety is mentioned as

31

a criterion for an acceptable solution. They admit that for an arbitrarily chosen slip surface it is not possible to ensure that all these conditions are satisfied. Therefore, it is necessary to calculate the normal and shear forces and line of thrust in order to inspect the acceptability. Based on the example provide by Morgenstern and Price (1965), the normal and shear forces on the vertical inter-slice plane are relatively insensitive to the function f(x), provided that the function is reasonable. Hence, for an arbitrarily chosen slip surface it may not be possible to obtain a physically admissible slip surface. Whitman and Bailey (1967) In the work of Whitman and Bailey (1967), the Morgenstern and Price method is used to check the acceptability of an arbitrary slip surface. They conclude that some f(x)s may lead to an unreasonable distribution of stresses along the failure. Some f(x)s will lead to the line of thrust falling outside the failure mass and some f(x)s will result in values of shear forces that exceed the shear resistance available along the vertical inter-slice planes. Three examples are given in their paper, including homogenous and non-homogenous slopes. In each example, different assumptions of f(x) are used to inspect the acceptability of the solution. Sarma (1973) Sarma method (1973) is originally derived for the calculation of critical acceleration. He states that for a given slip surface, infinite numbers of solutions exist by assuming infinite variations in the shape of the distribution of the shear forces on the inter-slice boundaries. Of this infinite number, only those solutions that do not violate the failure criterion of the soil mass above the slip surface and which at the same time do not cause tension in the material are acceptable. Janbu (1973) Janbu (1973) mentions that the average factor of safety, Fv, along the vertical inter-slice boundary should be greater than F, where F is the factor of safety on the slip surface.

32

Zhu , Lee and Jiang (2003) According to Zhu et al. (2003), the solution by limit equilibrium method can be regarded as acceptable if the associated internal forces are statically reasonable. The following four criteria have been used to check the acceptability of the internal forces: (a) the effective normal stress along the slip surface is non-negative; (b) the line of thrust lies within the sliding mass (note that this criterion is not exactly sufficient to guarantee compressive stress on the inter-slice boundary, however, it is considered sufficient); (c) the effective normal internal forces are non-negative; (d) the local factors of safety along the vertical inter-slices are not less than those along the sliding surface. The satisfaction of stress acceptability is not an inherent requirement of limit equilibrium method and it is usually inspected after the solution is obtained. The validation of stress acceptability overburdens the computation of the traditional limit equilibrium method. Therefore, when an optimization procedure is used to search for the critical slip surface, which involves analysing a large number of potential slip surfaces, the check of stress acceptability is seldom carried Moreover, the above procedures only guarantee that the failure criterion is not violated on the vertical inter-slice plane. However, the failure criterion should be satisfied everywhere within the slice. It seems extremely difficult to obtain such a solution based on available limit equilibrium techniques by trial and error. Kinematical acceptability To examine the kinematical acceptability, let us look at two adjoining wedges shown in Figure 2.2. The left and right wedges move with absolute velocities Vl and Vr inclined at angle el and er to their bases or l and r to the x axis in a counter clockwise direction, where el and er are the angles of dilation. The relative velocity of the left wedge with respect to the right one along the inter-slice plane is represented as Vj, inclined at an angle j to the x axis. To satisfy kinematical acceptability, the two adjoining wedges must not move to cause overlap or separation. This implies that the velocity hodograph must be closed, therefore:

33

Vr + V j = Vl

(2.10)

From Figure 2.2 we can obtain (Donald and Chen, 1997):

V r = Vl

( sin (

sin l j
r

) )

(2.11)

V j = Vl

sin ( r l ) sin r j

(2.12)

Figure 2.2 Velocity compatibility between adjacent blocks. The left wedge moves upward. (a) Velocities of the blocks. (b) Velocity hodograph. (Donald and Chen, 1997)

From Figure 2.2 it is obvious that r > j and l > j. Therefore, the magnitude of Vr is positive if the magnitude of Vl is positive. To make the magnitude of Vj positive, it is necessary to have

r > l, where

r = i +1 + er

(2.13)

l = i + el

(2.14)

34

In homogenous slopes, the angle of dilation is constant so if r > l, then r > l. This means that to satisfy kinematical acceptability, the inclination angle of the base line of the next slice has to be larger than the inclination angle of the base line of the current slice. It is noted that the above conclusion is true only in homogenous slopes. In non-homogenous slopes, if er < el, r may be larger than l even if r < l. It applies that if the slip surface develops from a soil layer with a higher angle of dilation into a soil layer with a smaller angle of dilation, the inclination angle of the base line of next slice may be larger than the inclination angle of the base line of the current slice while the kinematical acceptability is still satisfied. However, this is too complex, therefore only the criterion, r > l is adopted as kinematical acceptability requirement in the new procedure. Kinematical acceptability can be satisfied by specifying the potential slip surface to be concave upward (r > l). Also, the inter-slice shear forces (Xi+1 in Figure 2.3) should show directions implying downward movement of the slice along the inter-slice boundaries. The direction of the shear force on the slip surface (Ti and Ti+1 in Figure 2.3) is obviously showing downhill movement. The effective normal forces (Ni, Ni+1 and Ei+1 in Figure 2.3) should be compressive and may be zero but not tensile.

Xi+1 Ei+1 Ti+1 A Ni+1

Ti Ni

Figure 2.3 Directions of normal and shear forces between adjacent slices

35

2.3 METHOD OF CHARACTERISTICS


Sokolovski (1960) assumes that the entire soil mass in the slope is in limiting equilibrium. It is assumed that a normal stress x = p(y) acts on the horizontal semi-infinite boundary, the y-axis, and then the contour of the slope is determined (Figure 2.4). The whole medium is assumed to be in a critical state and violation of equilibrium will induce a slide downwards. Along the positive y-axis, the normal and shear components of stress are defined by x = p(y) and

xy = 0. Along the contour of the slope, the normal and tangential components of stress are zero.
In the zone which has passed into the plastic state Sokolovski proves that there exists a system of two curvilinear families of surfaces of slip and derives two functions which change along these lines. By following a numerical technique, he solves the two functions to obtain characteristic lines and thereby determines the contour of the slope. It is interesting to note that the curved slope line is a result of having body forces and limiting equilibrium everywhere and is not due to variable surface load on the crest. This implies that if the slope surface is linear, then the whole of soil mass will not be in limiting equilibrium.

Figure 2.4 Geometry of the slope and forces acting on the slope (Sokolovski,1960)

36

2.4 LIMIT ANALYSIS


As stated by Chen (1975), in contrast to slip-line and limit equilibrium methods, the limit analysis method considers the stress-strain relationship of a soil in an idealized manner. This idealization, named as normality, establishes the limit theorems on which the analysis is based. The conditions required to establish an upper or lower solution are briefly discussed below. Lower bound theorem (Chen, 1975) The loads, determined from a distribution of stress alone, which satisfy: (a) the equilibrium equations; (b) stress boundary conditions; and (c) nowhere violates the yield criterion; are not greater than the actual collapse load. The distribution of stress is termed as a statically admissible stress field. The lower bound technique considers only equilibrium and yield. Upper bound theorem (Chen, 1975) The loads, determined by equating the external rate of work to the internal rate of dissipation in an assumed deformation mode that satisfy: (a) velocity boundary conditions; and (b) strain and velocity compatibility conditions; are not less than the actual collapse load. A velocity field satisfying the above conditions has been termed a kinematically admissible velocity field. The upper bound technique considers only velocity or failure modes and energy dissipation.

2.4.1 Rotation mechanism (for example, Chen, Giger and Fang, 1969 and Chen and Giger, 1971)
The upper bound theorem of limit analysis is applied to obtain numerical solutions for the critical height of an embankment. The work by Chen at el. (1969) considers only the log-spiral shape of slip surface passing through the toe of the slope and is extended by Chen and Giger (1971) for the same shape of slip surface passing below the toe. By equating external and internal energies for such failure mechanism it gives an upper bound on the critical height of the slope.

37

For more general use, the slip surface passing below the toe point is shown in Figure 2.5. The rate of external work done by region ABC'CA can easily be obtained by first finding rates of work W1, W2, W3 and W4 due to the soil weight in regions OBC'O, OABO, OAC'O and ACC'A, respectively. Then, the total rate of external work due to the weight of the soil in region ABC'CA is

r03 ( f 1 f 2 f 3 f 4 )
where:

(2.15)

f1 =

1 { ( 3 tan cos h + sin h ) exp [ 3( h 0 ) tan ] 3 1 + 9 tan 2 (3 tan cos 0 + sin 0 ) }

(2.16)

f2 =

L 6r0

L sin ( 0 + ) 2 cos 0 r cos 0

(2.17)

L 1 f 3 = exp[ ( h 0 ) tan ]sin ( h 0 ) sin ( h + ) r 6 0 L cos 0 cos + cos h exp[ ( h 0 ) tan ] r0

(2.18)

H f4 = r 0

L sin ( ) 1H r cos 3 r 2 sin sin cos 0 0 0


2

( ) + cot cot

(2.19)

The total internal dissipation of energy is found by integration over the whole surface as

h
0

c (V cos )

cr 2 rd = 0 { exp[ 2( h 0 ) tan ] 1} cos 2 tan

(2.20)

By equating external rate of work with internal rate of energy dissipation yields

38

H=

f ( h , 0 , )

(2.21)

An upper bound to the critical height of an embankment can be obtained by minimizing Function 2.21.

Figure 2.5 Logarithmic rotational failure mechanism (Chen and Giger, 1971)

2.4.2 Translational mechanism (for example Michalowski,1995)


If a rigid rotation mechanism is assumed for limit analysis, a log-spiral slip surface is always used. Some researchers rare extend the limit analysis for translational mechanisms, for example, Karal (1977) and Izbicki (1981) developed the technique earlier. More recently, Michalowski (1995) extended the translational mechanism for limit analysis with slices. his work is reviewed in the following section. The translational failure mechanism is shown in Figure 2.6a. The entire failing mass is divided into n blocks and a single block is shown in Figure 2.6b. If an associate flow rule is assumed, it requires that the velocity of block k-1 is inclined to the base line at angle (Figure 2.6c). For the same reason, the velocity of adjacent block must move with a direction inclined at angle to its

39

base line. The two blocks then move with different velocities, [V] is the velocity jump between the two blocks, which can be found by using hodograph in Figure 2.6d.

Figure 2.6 (a) Translational failure mechanism; (b) single block; (c) velocity vectors; (d) hodograph (Michalowski, 1995)

The hodograph for the entire failure mechanism is shown in Figure 2.7. The respective velocities for each block and the inter-slice velocity jump are:

Vk = Vk 1

cos k 1 k 1 k cos k + k k

(2.22)

[V ]k

= Vk

sin ( k k 1 k + k 1 ) cos k 1 k 1 k

(2.23)

where k is the angle of resistance at the base of the slice andk is the angle of resistance at the inter-slice between block k-1 and k.

40

Figure 2.7 (a) Hodograph for a homogeneous slope; (b) hodograph when the internal friction on inter-faces between blocks is neglected (Michalowski, 1995)

If the weight of block k is denoted Gk and any additional vertical load on the block k is denoted Qk, the rate of external work can be written as:
n

W=

(G
k =1

+ Qk )Vk sin ( k k )

(2.24)

The energy dissipation rate in the entire mechanism becomes:

D=

[ l c V
n k k k =1

cos k + t k ck [V ]k cos k

(2.25)

where tk and lk are shown in Figure 2.5b. Then the factor of safety of the slope can be expressed as:

F=

[ l c V
n k k k =1

cos dk + t k ck [V ]k cos dk + Qk ) Vk sin ( k dk )

]
(2.26)

(G
k =1

where dk = tan-1(tan k / F) anddk = tan-1(tank / F). Angles dk anddk also need to be used in Equations 2.22 and 2.23. Solutions for F are found by iteration.

41

2.5 FINITE ELEMENT METHOD


The finite element method is used as an independent tool to validate the new procedure proposed in this thesis. The basic formulation of finite element method and its application in slope stability analysis is reviewed in Chapter 7. In Pottss Rankine lecture (2003), he concentrated on the debate that Numerical methods of analysis have reached the stage where they are superior to conventional approaches and can replace them in the geotechnical design processes. However, as shown in this thesis, the traditional method can be enhanced and thus achieve similar results as the finite element method. The traditional methods are more easily understood by design engineers and more cost-effective, compared to the finite element methods.

42

3. Optimization Techniques Used in Slope Stability Analysis to Search for the Critical Slip Surface

3.1 SCENARIO
The analysis of slope stability using the limit equilibrium method involves: first, the calculation of the factor of safety (or critical acceleration) for a given slip surface and, second, a search for the slip surface with the minimum factor of safety (or critical acceleration) for the slope. Note that, for a given slope, the critical slip surface associated with the minimum factor of safety may be different to the critical slip surface associated with the critical acceleration (Sarma & Bhave, 1974). The critical slip surface in slope stability analysis is the surface that produces the minimum factor of safety in static analysis or the minimum critical acceleration in pseudo-static analysis. Since the process of finding the critical slip surface is linked to the technique for finding the minimum factor of safety (or critical acceleration), it is natural to consider using an optimization method. There are many available methods to determine the critical slip surface; some of these methods and their basic features are reviewed in this chapter.

3.2 PATTERN SEARCH SCHEME


In Pattern Search Scheme (PSS), a specified shape of slip surfaces has to be used. Theoretically, any shape of the slip surface can be used, however, circular or log-spiral slip surfaces are the most commonly used in the literature. The advantage of using circular or log-spiral slip surfaces is that to define the position and shape of the surface, only three parameters are needed: two to define the centre of the surface and the third to specify the radius. To carry out the PSS analysis, an area where the centre of the surface may exist is defined by the user and a grid is established by specifying increments in both the x and y directions. For each point in the grid, the radius of the surface is varied to obtain the minimum factor of safety (or critical acceleration) for that point.

43

By comparing the associated factor of safety for each centre point, the minimum factor of safety can be found.

3.2.1 Circular slip surface (e.g. Bishop and Morgenstern, 1960)


The PSS is frequently used in the literature for the formulation of stability charts of slopes. For example, a simple slope geometry is used by Bishop and Morgenstern (1960). Its geometry is presented in Figure 3.1, among which D is termed a depth factor. For instance, for an earth dam founded on bedrock the depth factor would be unity. The simplified Bishop Method is used for the calculation of the factor of safety.

Figure 3.1 Specification of parameters (Bishop and Morgenstern, 1960)

A set of contours of equal values of factor of safety based on forty-nine circles is shown in Figure 3.2 for a homogenous slope, and the geometry and the soil parameters are shown in the same figure. Some conclusions on the position of the critical circle have been drawn, for example, the change of factor of safety is relatively sensitive to a change in the y coordinate of the slip circle centre as compared to a change in the x coordinate. For the formulation of stability charts, approximately twenty circles were used by Bishop and Morgenstern (1960) to identify a minimum in each case.

44

Bishop and Morgenstern (1960) conclude that for a given value of dimensionless number, c' / H, the factor of safety depends on the geometry of the slope, on the water pore pressure ratio and on the angle of shearing resistance. For a simple homogenous slope, to a close approximation, the factor of safety varies linearly with the value of the pore water pressure ratio, ru, which is defined as:

ru =

u H

where u is the pore-water pressure at any point and H is the depth of that point below the slope surface.

Figure 3.2 Detailed contours for a typical case (Bishop and Morgenstern, 1960)

3.2.2 Log-spiral slip surface (e.g. Prater, 1979; Lighthall, 1979)


Prater (1979) The log-spiral slip surface is used by Prater (1979) to find the critical slip surface. The expression of the log-spiral surface is shown in Figure 3.3. This formulation is only valid for the slip surface

45

passing through the toe point and two angles are needed to define the position and the shape of the slip surface. An initial value of angle t (Figure 3.3) is assumed to be less than slope angle i and a relatively small angle z is assumed. Both angles t and z increase within possible limits and the critical accelerations are computed for each combination. The values of increments for both angles depend on the precision required. Then the critical value can be identified by simply comparing all the computed critical accelerations.

Figure 3.3 The logarithmic spiral surface for toe failure (Prater, 1979)

Lighthall (1979) In Lighthalls analysis (1979), log-spiral slip surfaces are also used to find the critical accelerations. The formulation of the slip surface is shown in Figure 3.4. The method is suitable for a general slope. The slip surface may or may not pass through the toe. Although there is a difference in the formulation of critical acceleration equations between Prater (1979) and Lighthall (1979), the values passing through the toe point calculated by the two methods are identical. The technique of searching for the critical pole is as follows: coordinates x0 and y0 are chosen by the user as the lower corner of the grid so that the critical centre is expected to be right and above the starting point (Figure 3.5). The increments, dx and dy, are defined by the user. Calculation of

46

critical acceleration is made on successive grid points in each row until an increase in the value of critical acceleration is encountered. The lowest value is stored for that row and the same procedure is carried out for the next row. Successive minimum values for each row are compared until an increase of lowest value is encountered. This value is termed the minimum critical acceleration. It should be noted that the above procedure is only valid for homogenous slopes with simple geometry. In such a case, the contours of critical accelerations form a uniform field and no local minimum exists.

Figure 3.4 The geometry of log spiral slip surface for non-toe failure (Lighthall, 1979)

The effect of varying toe intercept is investigated by Lighthall (1979). It is concluded that for stable slopes, which have high critical accelerations, differences between the general critical slip surface and the critical surface passing through the toe can be expected to be no more than 2%. For a slope with marginal stability (kc 0), slip surfaces passing through the toe are always critical.

47

Figure 3.5 Typical contours of kc values (Lighthall, 1979)

3.2.3 Comments on the Pattern Search Scheme


There are two obvious shortcomings in the PSS. Firstly for a complicated slope, to find the critical value the grid spacing becomes important. Secondly a large number of calculations are needed to avoid finding a local minimum: however, with the development of fast computers, this is a less important issue. The main drawback of this procedure is that a specified shape of the slip surface has to be used, e.g. circular or log-spiral. For non-homogenous slopes, from both practical experience and theoretical analysis it is seen that the critical surface is not of a simple shape. On the other hand, one advantage of PSS with grids is that local minima are found which may be important, particularly in seismic analysis, if the global minimum gives a surface that is not important, such as a superficial slide.

48

3.3 CALCULUS OF VARIATIONS


The calculus of variations is a method to deal with the problem of maxima and minima, studying maxima and minima of functionals instead of functions (Revilla and Castillo, 1977). A functional is an application of a set of functions on the set of real numbers. In the case of the stability of slopes, slip surfaces can be expressed as the functions and the real numbers corresponding to the functions are their associated factors of safety (or critical accelerations).

3.3.1 The application of calculus of variations in slope stability analysis


Revilla and Castillo (1977) The Simplified Janbus method (1956) without correction was used by Revilla and Castillo (1977) to formulate the equations for the calculation of factor of safety for simple homogenous slopes. The analysis is presented for cohesive frictionless soil. In their article, the factor of safety is expressed as a quotient of two integrals:
x x F (x, y, y )dx S= x x G (x, y, y )dx
1 0 1 0

(3.1)

where F is a function of the available strength, G is a function of the shear stress obtained from the equilibrium and y represents the slip surface. However, as stated by Revilla and Castillo (1977), the analysis of this kind of functional is not generally studied in the existing calculus of variations literature. Therefore the equation of the quotient of the two integrals is transferred to an integral-differential equation shown in Equation 3.2, which is used as the Eulers equation for the problem.

F d F x0 y dx y = x1 G d G x0 G (x, y, y )dx y dx y
x1

F (x, y , y )dx

(3.2)

49

Then, transversality conditions together with appropriate boundary conditions are applied to solve the problem.

Figure 3.6 Diagram used for Janbus method (Revilla and Castillo, 1977)

The first observation based on the above method is that the computed factors of safety are significantly lower than the values from the upper bound solution given by the friction-circle method (Taylor, 1937). Friedli and Giger (1978) comment that an empirical correction factor should be applied to the results to account for the influence of shear forces along the vertical boundaries within the simplified Janbus method. This brings the comparisons into closer agreement. Another conclusion made by Revilla and Castillo is that the slip surface passing through the toe may not be the critical slip surface. Baker and Garber (1978) Baker and Garber (1978) present a calculus of variations approach for a soil with both cohesive and frictional components. The equation is formulated for the calculation of the factor of safety in terms of two unknown functions, one is the equation of the potential slip surface and the other is the normal stress distribution along this surface. The problem is expressed as the minimization of horizontal force equilibrium equation by factor of safety, while vertical force and moment

50

equilibrium are applied as constraints. Baker and Garber (1978) conclude that taking into account the rotational mode of failure, a log-spiral surface is the first set of possible slip surfaces for homogenous slopes, while if the translational mode of failure is considered, a straight line is the next set of possible slip surfaces for homogenous cases. However, the factor of safety based on Baker and Garbers analysis is independent of the normal stress distribution along the critical slip surface. This critical defect of their work has been addressed by several researchers and the details are given in Section 3.3.2. A detailed computation scheme is presented for the determination of the minimum factor of safety and corresponding critical slip surfaces, however, no results are given. In addition, they propose the application to layered slopes, however, because of the critical defect in the analysis technique noted above, the success of its application for non-homogenous slopes is not guaranteed. De Josselin De Jong (1980) A calculus of variations approach is applied to the stability of vertical cut offs in cohesive frictionless soils by De Josselin De Jong (1980). In this article he attempts to show that the calculus of variations method is not appropriate for slope stability problems. In the article, the total equilibrium equations agree with the equations used by Baker and Garber (1978) where the angle of shearing resistance is set to zero. Additional local equilibriums along stress characteristics are also added to the formulation. According to De Josselin De Jong (1980), three additional investigations have to be made for establishing the character of the solution, which are the Legendre condition, indicating whether the solution represents a maximum or a minimum; the Jacobi condition, indicating whether there exists an extremum at all and the Weierstrass function, showing whether the extremum is strong or weak. He concludes that the calculus of variation may lead to a weak maximum (or no extremum at all) and therefore when the application to the slope stability problems is involved, it may produce unsafe predictions.

3.3.2 Comments on variation of calculus


As shown by De Josselin De Jong (1981), the essence of the defect of Baker and Garbers work is that the functional is degenerated, because the auxiliary function used by Baker and Garber (1978) is independent of
y and is linear in , where is the normal stress along the slip surface and x x

51

y is the function of the slip surface. Because of the absence of

the functional has its x

extremum only for particular distributions of . In that case there is only an extremum for that functional (maximum or minimum) if is bounded. However, in the mechanical sense is not bounded, because stress states are limited in stress space by a cone. Therefore, no extremum exists and the method breaks down. Furthermore, a similar conclusion is drawn by Castillo and Luceno (1982) that the functional given by Baker and Garber is unbounded and in consequence the problem is incorrectly stated. Moreover, the successful application of calculus of variations for non-homogenous slopes is seldom seen in the literature. As concluded by De Josselin De Jong (1981), although available for many years and apparently useful for solving a basic problem in soil mechanics, the calculus of variation method is not pursued by leading scientists.

3.4 CALCULUS-BASED OPTIMIZATION METHODS


Calculus-based optimization methods are one of the most popular types of methods used to search for critical slip surfaces. Some of these approaches are reviewed in the following according to the optimization method used.

3.4.1 Dynamic programming


Dynamic Programming (DP) is one of the most widely-used optimization methods, combined with either limit equilibrium technique or finite element stress field to search for critical slip surfaces. DP is a numerical algorithm for the optimization of sequential multi-stage decision problems. When DP is applied, the search space is divided into an appropriate number of stages; moreover, at each stage an appropriate number of states are created. The procedure of DP is explained in the following. Let points (i, j) and (i+1, k) denote arbitrary states j and k at any two successive stages as shown in Figure 3.7. The trajectory jk, which connects the two points, is assumed to form a part of the potential slip surface. If the function to be optimized is G, then DGi (j, k) is used to express the

52

changed value in the function G and is sometimes called the return function. If the optimal value function, Hi (j) represents the minimum value of G between the initial stage and the point (i, j), then according to Bellmans principle of optimality (-), the optimal value for the arbitrary point in the next stage can be expressed as:

H i +1 (k ) = min

j =1~ Si

[ H i ( j ) + DGi (i, k ) ] ik==11~~nS

i +1

(3.3)

with the boundary conditions:

H1 ( j ) = 0
And

j = 1 ~ S1

G m = min G = min [H n +1 ( j )]
j =1~ S n +1

where Hn+1 (j) represents the optimal value for all the states at the final stage, the minimum value has to be employed as shown in Equation 3.3. After the optimization reaches the final stage and the minimum G has been obtained, the optimal path can be traced. This is the critical slip surface as shown in Figure 3.8.

Figure 3.7 Stages and states used in DP (Baker, 1980)

The advantage of DP is that the algorithm does not require the existence and uniqueness of derivatives. However, local minima may exist in the search space and therefore, the global

53

minimum may be missed by such a procedure. A procedure to divide the search region into a number of sub-regions is proposed by Baker (1980) but no results are given. DP has been employed by Baker (1980) for several slopes with Spencers method to compute the factor of safety, and compared with results reported in the literature. The method of DP has been seen to identify more critical slip surface that yield lower factors of safety. Another attractive feature of DP is that it is easy to combine the method with the finite element method. DP has been applied by many researchers to search for the critical slip surface from a finite element stress field. This will be reviewed in Chapter 7.

3.4.2 Simplex Analysis


Generally speaking, the search algorithm for the critical slip surface by applying the simplex technique for N-dimensional space involves the following steps (Nguyen 1985). 1 Set up an initial complex structure with (N+1) vectors, satisfying that the distances between arbitrary two vectors are equal. Each vector i comprises N variables, defining a potential slip surface. 2 Evaluate the factor of safety for every vector-defined slip surface and rearrange the values in descending order, then reflect the vector with the highest functional value to the opposite side (an example of reflection in three-dimensional space is shown in Figure 3.9). 3 Avoid the vectors which define physically impossible slip surfaces. 4 Repeat the simplex reflections until convergence. Circular slip surfaces are used by Nguyen (1985) in one of his examples, which involves only three parameters to be optimised. Comparing with the pattern search method, Simplex Analysis gives similar small factors of safety, however, it involves significantly fewer factor of safety computations. Nguyen gives another example of a slip surface with four points, which involves searching in four dimensional spaces. The slip surface has fixed x coordinates. As stated by Nguyen (1985), the advantage of the simplex analysis is that the technique is simple and effective and does not require the evaluation of derivatives. However, he states that to optimise a multidimensional optimal space, the simplex algorithm may converge to a local minimum.

54

Figure 3.8 The discretization scheme (Baker, 1980)

Figure 3.9An example of Simplex reflection for three dimensional space (Nguyen, 1985)

55

Bardet and Kapuskar (1989) apply the same technique to search for the critical slip surface. They use an eleven-dimensional optimal space in a homogenous slope, which gives factor of safety similar to the pattern search method using circular slip surfaces. When the technique is applied to non-homogenous slopes, they use only the two wedges method, which needs a three-dimensional space. The reason they use only the two wedges method in non-homogenous slopes is because of lack of effectiveness of the simplex technique in a multi-dimensional space.

3.4.3 Alternating Variable Method


In the Alternating Variable Method (AVM), each parameter used to define slip surface is chosen in turn and all the others are kept fixed. The minimum of the chosen parameter is obtained by a search technique for the optimization of the function with a single variable. The optimal parameter is then held constant at this conditional minimum point and the procedure is repeated for all the parameters used to define the slip surface (Li and White, 1987). Within this kind of optimization method, a simplified version has been proposed by Celestino and Duncan (1981). The slip surface is expressed as successive straight lines connecting several points, where the coordinates of the points are the parameters to be optimized. To search for the critical slip surface, a number of trials are performed. Each trial begins by shifting each of the points in the slip surface to two new positions (Figure 3.10), forming two new slip surfaces while the other points are kept in their original position. The first and second shifts are always equal distances and in the same direction. Then factors of safety are evaluated for each potential slip surface and the optimum position is estimated using the following equations:

x i* = x i0 +

F0 F1 x + x 2 F0 2 F1 + F2 F0 F1 y + y 2 F0 2 F1 + F2

(3.4)

y i* = y i0 +

(3.5)

where xi* and yi* are the estimated optimum position; x0 and y0 are original coordinates of the point i; x and y are shift increment; F0, F1 and F2 are factors of safety of the original slip surface, the slip surfaces after first and second shifts, respectively.

56

Then, the factor of safety is evaluated for the optimal position and compared with the factor of safety for the original position. If the value is smaller than the original one, the point i is shifted to the optimum position. If not, the point is returned to the initial position. The same process is repeated for each point in turn in the initial estimated slip surface, in turn, to minimize the objective function.

Figure 3.10 Shifting points on the slip surface (Celestino and Duncan,1981)

Two examples are investigated by Celestino and Duncan (1981) using the above technique with Spencers method to calculate the factor of safety. It is concluded that the method can give identical results to those using a circular arc and can estimate unusual slip surfaces. However, they comment that the assumption is only for local approximations of factor of safety and the objective function does not globally obey the assumptions. An extended version of the AVM is developed by Li and White (1987). The shift of each point has been constrained by classifying them into three categories: unconstrained nodal points, which can shift along x direction and y direction independently; nodal points with prescribed directions, which can only move along the boundary of different layers; and points moving along the slope surface, which are used to describe the starting point and end point of the slip surface. Li and White (1987) realize that a problem frequently occurring when using a poor initial guess is that the slip surface is made with relatively large number of points. To overcome this problem, they propose to first minimize coordinate y along the slip surface before searching for the

57

minimum of coordinate x for unconstrained nodal points. For example as shown in Figure 3.11, the searching sequence should be: sA, yB, C, D, yE, sF, xB, xE. The initial estimated slip surface is suggested to be expressed using few nodes and then new nodal points are introduced at the mid-point of the straight line joining successive nodes. An optimization technique, known as rational approximation (Zhou, 1982), is used to obtain the optimal position for each parameter to be optimized. This technique is thought to be suitable for functions of factor of safety, where the function cannot be expressed explicitly in terms of the factor of safety.

Figure 3.11 Description of non-circular slip surface by nodal points (Li and White, 1987)

3.4.4 Mixed Optimization Techniques


In the literature, some researchers use more than one optimization techniques to optimise the objective function, such as Chen and Shao (1988) and Yamagami and Ueta (1988). The reason for this is to test the benefits of different optimization techniques for slope stability problems.

58

Chen and Shao (1988) In the approach of Chen and Shao, the objective function can be expressed as:

F = F (Z )

(3.6)

where Z is the vector which presents the variables of the points shown in Figure 3.12. Instead of joining points by straight lines, Chen and Shao propose that cubic splines are used to connect points. This requires fewer individual points and the decrease in degrees of freedom is advantageous for numerical convergence of the problem. Nelder and Meads simplex method (1965), the method of steepest descent and DavidsonFletcher-Powell (DFP) method and its modification (1959, 1963) are used to optimize the Equation 3.6. It is noted that both the method of steepest descent and the DFP method require evaluation of the gradient of the function and some numerical method has to be used to obtain a reasonable estimation of the gradient. The above optimization methods are tested on four problems and prove successful. Chen and Shao (1988) conclude that while various methods are applicable, the modification of the DFP method is found to be essential because of special properties of slope stability analysis. Yamagami and Ueta (1988a) A similar objective function is adopted by Yamagami and Ueta (1988a) and Morgenstern-Price method is used to calculate the factor of safety. In their study, four nonlinear programming techniques are employed to solve the unconstrained optimization problem: simplex method proposed by Nelder and Mead, Powells conjugate direction method, the Davidson-FletcherPowell (DFP) method and the Broyden-Fletcher-Golfarb-Shanno (BFGS) method. The details of the theoretical aspects of above optimization methods are referred to textbooks (e.g. Avriel (1976) and Reklaitis (1983)). Among them, the FGP and the BFGS methods are gradient-based methods. Some issues have been researched by Yamagami and Ueta (1988a): for example, the effects of the initial estimates and the number of independent variables. One example used in their article is adopted to test and validate the new procedure proposed in this thesis.

59

However, special care should be taken for different problems and the critical slip surface found by the AVM strongly depends on the initial slip surface.

Figure 3.12 Discretization patterns (Chen and Shao, 1988)

3.4.5 Comments on Calculus-based optimization


Based on the above review, some common defects of calculus-based optimization can be drawn. First, it tends to identify a local minimum instead of the global minimum and the results are highly dependent on the initial slip surface assumed. Second, it is only applicable when there are few parameters to be optimised, which limits its application for complex problems.

3.5 RANDOM SEARCH (MONTE CARLO) TECHNIQUES


Random Search techniques are another optimization technique extensively used in the literature to search for the critical slip surface and the associated minimum factor of safety (or critical acceleration). The procedure followed is that potential slip surfaces are generated randomly and

60

factors of safety (or critical accelerations) are evaluated for each surface. Assuming that the number of generated potential slip surfaces is statistically large enough to cover most possible slip surfaces, the minimum factor of safety can be identified by comparing the factor of safety of each surface. Within the Random Search framework, there two main classes: random jumping and random walking. In random jumping, a large number of potential slip surfaces are generated, taking no account of previous trials (Siegel, 1981; Chen, 1992). In random walking, iterative slip surfaces are constructively generated by modifying the previous trial to get the next attempt (Greco, 1996; Husein Malkawi et al. 2001a, b).

3.5.1 The application of Random Search in slope stability analysis


Siegel et al. (1981) Siegel et al. (1981) state that the generated slip surfaces may be nonsensical both in shape and position if there is no control on the method. They force the generated slip surfaces to satisfy the following rules: (1) the starting point stays within the logical range of positions; (2) the slip surface is kinematically acceptable; (3) the whole slip surface lies above a strong layer; (4) the end point stays within the logical range of positions. They consider two kinds of conditions: the first is for general conditions and the second is for a zone of weakness within the mass of a slope. In the first condition, the initial inclination angle of the slip surface, , can be expressed as:

= 2 ( 1 2 )R 2
where 1 and 2 are shown in Figure 3.13 and R is a random number between 0 and 1. The successive inclination angle of the slip surface can be expressed as:

(3.7)

= 2 + ( 1 2 )R (1+ R )

(3.8)

61

where 1 and 2 are shown in Figure 3.14 and R is a random number between 0 and 1.

Figure 3.13Selecting initial line segment (Siegel et al., 1981)

If a zone of weakness within the slope exists, a series of boxes is located along the zone. A point is randomly selected from each box and the points are connected by straight lines between different boxes (Figure 3.15). The generation of slip surface outside the weakness zone is similar to the general condition.

Figure 3.14 Direction limits for successive line segments of an irregular surface (Siegel et al., 1981)

62

No results are given based on the above procedure. This procedure can be classified as a random jumping method, which only becomes practical for a few degrees of freedom because of the large number of slip surfaces that need to be computed.

Figure 3.15Sliding generator using more than two boxes (Siegel et al., 1981)

Chen (1992) Another method of generating potential slip surfaces is proposed by Chen (1992). The method involves: first, defining a search area that is expected to cover all the potential slip surfaces and second, generating a slip surface by connecting several points using straight lines (Figure 3.16). The points can be defined as:

cos i Z i = Z il + ri Di sin i

(3.9)

where Zil and Zir are successive points within the upper and lower boundaries chosen by the user,
Di is the distance between points Zil and Zir, i is the inclination angle of line ZilZir and ri is a

63

random number between 0 and 1. The kinematically unacceptable slip surfaces are abandoned immediately after their generation. Chen (1992) concludes that Random Search alone can be used to determine the critical slip surface or to estimate the initial trial for calculus-based methods. If purely Random Search is used, the solution approaches the minimum very slowly. However, if composite methods are used, it provides a quick and accurate solution.

Figure 3.16 The search and confidence areas (Chen, 1992)

Greco (1996) In Grecos method (1996), the potential slip surface is approximated by successive straight lines joining some points, where each point has two degrees of freedom (x and y coordinates). The optimal function can be expressed as:

S = {x1 , y1 , x2 , y 2 , , x n , y n }

(3.10)

Then a random walking optimization method is used to search for the minimum of the above function. In the proposed method, each search stage is divided into two phases: exploration and

64

extrapolation. In the exploration phase, each point i on the slip surface is moved from (xik , yik) to (xik+1 , yik+1), where:

xik +1 = xik + ik
yik +1 = y ik + ik

(3.11)

(3.12)

where:

ik = N x R x Dxik
ik = N y R y Dyik
where Rx and Ry are two random numbers between -0.5 and 0.5; Dxik and Dyik are widths of the search steps in x and y directions for point i at stage k; Nx and Ny are two defined numbers whose value can be either -1 or 1. If the move of the point i gives a smaller factor of safety, no further trials are made for that point and the widths of the search step, Dxik and Dyik, are increased. If not, the widths of the search step, Dxik and Dyik, are reduced. In the extrapolation phase, a new slip surface is then generated with points:

xie = 2 xik +1 xik


yie = 2 yik +1 yik

(3.13)

(3.14)

If the factor of safety associated with the new slip surface is less than for the previous slip surface the point i is updated to the new point. The iterative procedure is repeated for every vertex in the slip surface until defined criteria are satisfied. Greco (1996) concludes that the quality of the solutions of the proposed method is certainly comparable with those of the calculus-based methods and is sufficiently robust for layered soils with weak, thin-inclined layers.

65

Malkawi, Hassan & Sarma (2001 a) Malkawi, Hassan & Sarma (2001 a) present a new search procedure combining both random jumping and random walking to locate the critical circular slip surface. Every potential circular slip surface can be represented by three points: one is the centre of the circle and the others are the two points on the slope surface. Two boundaries, Xmin and Xmax, need to be defined as the search area. Then two points A1 and B1 (Figure 3.17) can be generated randomly as follows:

X1 A1 = X min + R1

( X max X min )
2

(3.15)

X1 B1 = X max R2

( X max X min )
2

(3.16)

1 YA = g X1 A 1

( )
( )

(3.17)

1 YB = g X1 B 1

(3.18)

where g is the slope surface function. The centre of the circular slip surface is believed to exist between D1 and E1 in the line perpendicular to the line A1B1. The critical centre can be obtained by moving the centre of the circle from E1 to D1 and evaluating factors of safety for all the possible slip surfaces. The method of search is to move point A and B systematically until defined criteria are satisfied. For example, the position of A2 as shown in Figure 3.18 is expressed as:
2 1 XA = X1 A +A

(3.19)

where A1 is a random distance and is given by:


1 1 A = N X R X DX A

66

where Rx is a random number between -0.5 and 0.5, Nx is either -1 or 1 and DXA1 is the current width of the search. If the factor of safety at the next trial is reduced, the width of the search is increased by some factor and vice versa. The procedure is also applied to the point B until both of two widths of search satisfy some defined criteria.

Figure 3.17Generation of the first slip surface (Husein Malkawi et al., 2001 a)

In this approach, both random jumping and random walking are employed. From their experience, it is better to start with random jumping and then resort to random walking.

Figure 3.18 Moving point A to the next stage (Husein Malkawi et al., 2001 a)

67

Malkawi, Hassan & Sarma (2001 b) Malkawi, Hassan & Sarma (2001 b) propose another random walking method to search for noncircular critical slip surfaces. The whole procedure is shown in Figure 3.19. In the first step, the starting and end points are generated randomly from two boundaries, Xmin and Xmax, and 1, 2 and

3 are three inclination angles that are randomly generated. Then the initial estimate of the slip
surface can be obtained as shown in Figure 3.19. In the next step, each segment of the slip surface rotates around both of its vertices. The angle of rotation is defined as

i j +1 = i j + R ij N d
where R is a random number between -0.5 and 0.5, previous slip surface,
j j i

(3.20)

is the range of rotation angle and Nd is

either -1 or 1. If the factor of safety of the newly generated slip surface is lower than that for the
i

is reduced and vice versa. The whole search procedure stops when

defined criteria are satisfied. Malkawi, Hassan & Sarma (2001 b) conclude that the procedure gives accurate, efficient and high-quality solutions and can handle any slope geometry, soil layers and external loads.

3.5.2 Comments on Random Search technique


The difference between the various methods in the literature is how they generate the potential slip surfaces using random numbers. The simplest method is that the slip surface is expressed by successive straight lines joining several points whose coordinates can be seen as the random numbers. In such a case, each point has two degrees of freedom. As pointed out by Chen (1992), hundreds or even thousands of trial slip surfaces may be needed if the problem involves five to seven degrees of freedom, which means only for three to four points. In other words, the success of the random search technique depends on the effectiveness of the generation of potential slip surfaces. For the procedures based on random walking such as Greco (1996) and Husein Malkawi et al. (2001a, b.), they admit that the theoretical background of the proposed methods is very poor and does not ensure that the minimum found is global.

68

The critical defect of random search is that a solution approaching the optimum is found only by chance. If the same problem is re-analysed, the technique seldom gives an identical solution.

Figure 3.19 Step-by-Step search procedure for rotating three-segment example to minimize factor of safety (Husein Malkawi et al. (2001 b)

69

3.6 NEW OPTIMIZATION METHODS


With the development of modern optimization methods, more and more techniques are employed in slope stability analysis to optimise the slip surface function to obtain the critical slip surface. A few of them are reviewed in the following section but it does not include all the comments made within the reported studies. Cheng (2003) The Simulated Annealing Method (SAM) is used to by Cheng (2003) to obtain the critical slip surface. According to Cheng (2003), SAM can perform well for non-convex functions with the presence of many local minima, which is very suited for the optimization of the slip surface function. Another attractive feature of the SAM is that a good initial estimate is not required. The potential slip surface ACDEFB is shown in Figure 3.20, in which A and B are the starting and end points, respectively. Two upper and lower boundaries for A and B can be fixed by the user. As for the interior points C, D, E and F in the slip surface, they are assumed to be uniformly distributed in the horizontal distance between points A and B. For the first interior point C, C1, which has the same x-coordinate as C in the slope surface, is used as an upper boundary for point
C. The y-coordinate of the lower point is set equal to C1 AB / 4. For the second interior point D, G, which has the same x-coordinate as D in the extension line of AC, is taken as the lower

boundary of D. The upper boundary as D is similar to the setting of the lower boundary of C. The same procedure is repeated until the entire upper and the lower boundaries of the control variables are defined. The generation of trial slip surface and the search direction proceeds with the SAM and the global minimum can be located easily with high accuracy. Cheng (2003) concludes that the proposed technique is efficient for relatively complicated cases and is applicable even when a very thin soft band is present.

70

Figure 3.20 Generation of a non-circular failure surface (Cheng, 2003)

McCombie & Wilkinson (2002) and Zolfaghari, Heath & McCombie (2005) A Simple Genetic Algorithm (SGA) is used by McCombie and Wilkinson (2002) to find the critical circular slip surface. A similar technique is adopted by Zolfaghari et al. (2005) to search for the critical non-circular slip surface. The main operations involved in the SGA are described below. (1) Initialisation. A SGA starts with a population of N possible solutions to a given problem, each possibility represents a potential slip surfaced and is generated randomly. (2) Reproduction. The probability of a possible solution being selected for reproduction is directly proportional to its fitness. (3) Crossover. Crossover occurs if a randomly selected value is below a fixed probability. Two new potential slip surfaces are generated by some crossover rule based on two old potential slip surfaces. (4) Mutation. This process is used to prevent the loss of potentially useful information, which is applied to only a small proportion of the population. (5) Evaluation. The new generation is evaluated.

71

All the operations are repeated until defined criteria are satisfied. When the technique is applied for circular slip surfaces, each possibility has only three variables, which are the x and y coordinates of the centre and the radius R. McCombie and Wilkinson (2002) conclude that it performs better than either a pattern search procedure or a random search procedure. The application of SGA to non-circular slip surface needs the formulation of the slip surface using random numbers. As shown in Figure 3.21, the first x coordinate of the slip surface is created randomly. Then the angular difference fi between each successive slip surface slope are chosen randomly for all slices. Because fi is generated randomly, for starting from a sufficient population, 11 different categories of fi are defined. Each category gives different range for the possible fi. A set of N different slip surfaces can be created based on the above formulation and the SGA can be used to minimize the objective function.

Figure 3.21 Non-circular failure surface used in the Genetic Algorithm (Zolfaghari et al., 2005)

Zolfaghari, et al. (2005) conclude that the present search method can be used in order to analyse the stability of earth dams, finite and infinite natural slopes. For non-homogenous slopes, noncircular slip surfaces are essential for an accurate assessment of stability.

72

3.7 COMMENTS ON THE OPTIMIZATION TECHNIQUES USED IN SLOPE STABILITY ANALYSIS


It can be predicted that with the development of new optimization methods, more and more such techniques will be used to search for critical slip surfaces and combined with the limit equilibrium method. With careful generation of the potential slip surfaces using variables and proper use of optimization techniques, the mathematically minimum factor of safety (or critical acceleration) and the associated critical slip surface can be found. Until now, no stress acceptability criteria are used to judge the critical slip surface obtained from optimization techniques. Therefore, although the results obtained from any optimization technique may mathematically be the minimum, some results may not be acceptable from soil mechanics principles.

73

4. Acceptability Criteria and Formulation of the New Procedure

4.1 SCENARIO
It is known that the limit equilibrium technique of slope stability analysis is statically indeterminate. As shown in Section 4.2, given a possible slip surface, inter-slice force distributions exist that will produce any critical acceleration as a solution. However, this distribution (and therefore the critical acceleration) may not be acceptable. A new method of finding the critical slip surface in slope stability analysis is developed, which uses stress acceptability as a prerequisite to derive a system of non-linear equations to determine the slip surface. The formulation of the new procedure is presented in this chapter, starting from a general slice, followed by special considerations for the first slice and the last two slices.

4.2 THE REQUIREMENT OF ACCEPTABILITY CRITERIA - THE BASICS OF LIMIT EQUILIBRIUM TECHNIQUE FOR THE CRITICAL ACCELERATION ANALYSIS
A schematic diagram for a slope is shown in Figure 4.1 and the variables are listed in the following:

The slope is defined by y = g(x). The slip surface is defined by y = f(x). The assumed line of thrust is defined by y = t(x). The height of the line of thrust from the slip surface z(x) = t-f. The slope of the slip surface is given by = arctan (f (x) ). The slip surface starts from x = 0 and ends at x = B.

74

g(x) y x t(x)

dx f(x)
E z X

dx kcdW dW

E+dE X+dX T z+dz

N Figure 4.1 A schematic diagram for a slope and the forces and their acting points in one slice

For an elemental slice of width dx: The weight of the slice dw = ( g f ) dx. The normal force on the slip surface N= n dx sec . The shear force on the slip surface T= dx sec . The normal force on the inter-slice boundary E The shear force on the inter-slice boundary X

Consider an elemental slice of width dx. Vertical equilibrium of the slice gives:

n dx + f dx = dw + dX
Horizontal equilibrium gives:

(4.1)

dx n f dx = k c dw + dE
Taking moments about the mid point of the slip surface of the slice gives:

(4.2)

75

(E + dE )(z + dz + f dx / 2) E (z f dx / 2) ( X + dX ) dx / 2 Xdx / 2 + k c dw (g f ) / 2 = 0
By neglecting higher order terms, we get:

(4.3)

Edz + zdE + Ef dx Xdx + k c dw( g f ) / 2 = 0


Dividing by dx throughout and taking limits:

(4.4)

Ez + zE + Ef X + k c w( g f ) / 2 = 0
Mohr- Coulomb failure criterion along the slip surface gives:

(4.5)

tan + c =n

(4.6)

For the purpose of demonstration, the pore water pressure is neglected in the current stage, therefore, n=n. Also we write (for ease of presentation) tan' = a. Then from Equations 4.1, 4.2 and 4.6 we get:

n (1 + af )dx = dw + dX c f dx n (a f )dx = k c dw + dE c dx
From Equations 4.7 and 4.8 we get:

(4.7)

(4.8)

(a f )[dw + dX c f dx] (1 + af ) [k c dw + dE c dx] = 0


Dividing by dx throughout and taking limits, we get:

(4.9)

(a f )[w + X cf ] (1 + af ) [kcW + E c] = 0
or

(4.10)

76

(a f ) X (1 + af )E = (1 + af ) k c w (a f )w c (1 +

f 2

(4.11)

We can eliminate X from Equations 4.11 and 4.5 by differentiating Equation 4.5 first and replacing X from Equation 4.11:

E (z + f ) + E (2 z + f ) +

E z 1 / (a f ) (1 + af )E + (1 + af ) k c w (a f )w c 1 + f 2 + k c w( g f ) / 2 + k c w( g f ) / 2 = 0
Rearranging:
E z + E [2 z + f (1 + af ) / (a f )] + E [z + f ] = 1 / (a f ) (1 + af ) k c w (a f )w c 1 + f 2 k c w( g f ) / 2 k c w( g f ) / 2

)]

(4.12)

)]

(4.13)

By taking limits, note that


w= (g-f) w= (g-f)

Then:

E z + E [2 z + f (1 + af ) / (a f )] + E [z + f ] = 1 / (a f ) (1 + af ) k c w (a f ) w c 1 + f 2 k c w( g f ) / 2 k c w( g f ) / 2 ( g f )

)]

= k c [( g f )(1 + af ) / (a f ) ( g f )( g f )]

(4.14)

(1 + f )c
2

(a f )

Therefore, for a given slope and an arbitrary slip surface, assuming that the line of thrust is known and the critical acceleration is also assumed known, Equation 4.14 is a 2nd order linear differential equation in E with variable coefficients. With the known boundary conditions that E=
0 at x=0 and at x=B, the above equation can be solved numerically or otherwise. There is a

second set of boundary condition that X=0 at x=0 and at x=B, which are automatically satisfied

77

by the solution of E since z=0 at x=0 and at x=B from Equation 4.5. Alternatively, for any chosen
E function that satisfies the two boundary conditions, kc can be determined.

Example 4.1 The geometry and the soil properties of an example slope are shown in Figure 4.2. For simplicity, the slip surface is arbitrarily expressed as f (x ) = 0.1111x 2 0.5 x + 1 . The line of thrust is arbitrarily taken as a straight line, joining the toe point and the end point of the slip surface.

Figure 4.2 An example slope and its soil properties

Equation 4.14 is solved by the three-stage Lobatto IIIa formula available in the Differential Equations Toolbox in MATLAB (Kierzenka and Shampine, 2001) by assuming different kc values. The variations of E resulting from different values of kc are shown in Figure 4.3a. The variations of X can be found from Equation 4.5 and the local factor of safety along the vertical inter-slice boundaries can be evaluated as shown in Figure 4.3b. None of these results are acceptable when local factors of safety are computed (the reason for that is mainly from the assumed line of thrust). For the others one has to check the X functions and further the local factors of safety along the inter-slice boundaries for acceptability.

78

(a)

(b) Figure 4.3 (a) The variations of E resulting from different values of kc (b) The variations of local factors of safety resulting from different values of kc

79

This solution clearly shows that without any control on the inter-slice forces E or X or both within the slipping mass, we can get infinite numbers of solutions of kc since in the above formulation, we can choose any kc and any line of thrust. Some of these solutions may not be acceptable. Therefore, it is very essential that some acceptability criterion is provided to curtail the number of solutions. The available limit equilibrium solutions apply some sort of control over the inter-slice forces to obtain reasonable results. However, these still may not be perfectly acceptable. The method presented in the thesis shows a way to apply acceptability criteria with rigour.

4.3 NEW PROCEDURE AND GENERAL ASSUMPTIONS

4.3.1 New procedure


The basic feature of the new procedure is to answer this question: Is it possible to find an acceptable slip surface in a slope for a critical acceleration? If many acceptable slip surfaces are found with different critical accelerations then the surface with the minimum critical acceleration is considered as the critical slip surface. The new procedure is developed to determine the critical slip surface along with the critical acceleration. Sarma (2004) formulates this procedure in its basic state for homogenous slopes. In this thesis, the new procedure is modified and enhanced to include non-homogenous slopes in terms of strength and pore-water pressure. This procedure is within the framework of the limit equilibrium technique, satisfying acceptability criteria in terms of stresses and kinematics. Moreover, it provides information on the critical acceleration and the kinematically acceptable inter-slice boundaries for the analysis of seismic displacements using a multi-block sliding model (Sarma and Chlimintzas, 2001; Chlimintzas, 2003). In this procedure, a segment of a slip surface and the corresponding inter-slice boundary are found by solving a set of non-linear equations given later in Section 4.4. Therefore, the procedure starts with solving the first segment. The next segment is solved following the same procedure

80

until an entire slip surface from the toe end to the crest is found. Thus the slip surface comprises of a series of straight lines with associated inter-slice boundaries. In traditional methods of slices, the sides of a slice are normally taken as vertical. However, even if the mass contained within the slip surface is in a state of limiting equilibrium, the mass will not be able to move unless shear surfaces are formed within the body, except for planar or circular slip surface. The slices need not be vertical or even parallel and the inclinations of slices are so defined that a kinematical slip mechanism can develop as adopted in the inclined slices method (Sarma 1979). If the generated slip surface is not kinematically acceptable, then the desired critical acceleration has to be changed and this needs an iterative technique which is discussed in Chapter 5.

4.3.2 General assumptions


(1) The soil is a rigid-perfectly-plastic material with shear strength obeying Mohr-Coulomb failure criterion. (2) At the state of failure, the shear strength is fully mobilised on the slip surface and on the inter-slice boundaries. (3) The soil properties do not change significantly during earthquakes (if there is degradation of strength due to increased pore water pressures, then this must be incorporated in the analysis). (4) The vertical component of ground motion, affecting only the gravity, is not considered (The way to incorporate the vertical component of earthquake acceleration has been discussed in Section 1.2). (5) The slip surface comprises of a series of straight lines and the inter-slice boundaries are represented as straight lines, connecting the slip surface and the slope surface.

4.4 THE ANALYSIS OF A STANDARD SLICE


The analysis for a general standard slice is presented first, where the number of unknowns is identified in order to configure the geometry of the slice and the forces acting on the slice. In the new procedure, the slip surface and the inter-slice boundaries are not predefined. The equilibrium

81

of the slice and the acceptability criteria (defined later) determine the slip surface and the interslice boundaries of the slice. Pore water pressure is also included in the analysis.

4.4.1 Unknowns of a standard slice


A possible slip surface in a slope is shown in Figure 4.4, which is at incipient failure under the action of a critical acceleration kcg. Note that the slip surface is unknown at the beginning but the applied acceleration is assumed. The acceleration is critical since the forces on the slip surface will be assumed to be in a limiting state. The slip surface from a starting point O to point A is assumed solved already and the corresponding inter-slice boundary at point A is AD. Therefore, all the forces on the slip surface and the inter-slice boundaries up to the point A from O are known. In order to define a new point on the slip surface, an arbitrary slice is considered. Within this slice, there are many unknowns and therefore, some assumptions have to be made before the problem can be solved analytically.
Geometric unknowns

The slice i in Figure 4.4 is composed of a possible slip surface AB, a free surface CD and two internal boundaries AD and BC. The points A and D are known. The aim of the solution technique is to find points B and C to satisfy the adopted criteria. For a given slope to be analysed, the ground surface function g(x) is known. Any horizontal increment x can be assumed in the ground surface from D, which, with g(x) defines point C (The effect of increment x is discussed in Chapter 5 and Chapter 6 for homogenous and non-homogenous slopes). Therefore, the only unknown in the new slice is another point in the slip surface, B. The two geometric unknowns are the angles, i and i+1, which are the angles made by the slip surface AB with horizontal and the inter-slice boundary BC with the vertical, respectively. Assuming the two angles as known, the position of B can be defined as: If i+1 = 0, then

xi +1 = xsi +1
If i+1 0, then

(4.15a)

82

xi +1 = ( xs i +1 / tan i + xi tan i + ysi y i ) /(tan i 1 / tan i ) y i +1 = y i + ( xi +1 xi ) tan i


where x & y and xs & ys are shown in Figure 4.4.

(4.15b)

(4.16)

sx D(xsi,ysi) Xi Ei O y g(x) zi i A(xi,yi) x i li

C(xsi+1,ysi+1)
Soil Layer 1

Xi+1 Ei+1 kcWi (xgi,ygi) i+1 Wi zi+1 Ti B(xi+1,yi+1) bi Ni

Soil Layer 2

Soil Layer 3

Soil Layer 4

Figure 4.4 Forces acting on an inclined slice (The total forces N and E are shown, which are the sums of the effective force and the force due to pore water).

Force unknowns

Assuming that the geometric unknowns are known, the forces acting on the new slice are shown in Figure 4.4. The weight of the new slice Wi is obtained from the geometry and the unit weight of the soil. The horizontal force kcWi acting on the slice is known since the critical earthquake acceleration kcg is assumed. On the known boundary AD, the effective normal and shear forces,
E'i and Xi, are known from the previous slice. The forces due to pore water pressures Ui, PWi and PWi+1, acting on the slip surface and the inter-slice boundaries respectively, are known once the

boundaries are defined (the definition of pore water pressures is discussed later in Section 4.4.6). Therefore, the effective normal and shear forces on the slip surface, N'i and Ti, and the effective

83

normal and shear forces on the inter-slice boundary, E'i+1 and Xi+1, are the four unknowns in terms of forces to be solved. The soil is assumed to obey the Mohr-Coulomb failure criterion in terms of effective stresses (any other failure criterion can be seen as an extension of this model). Under the action of the critical acceleration kcg, the forces on the slip surface are in limiting equilibrium and therefore, applying the Mohr Coulomb failure criterion gives:

Ti = N i tan i + cibi sec i

(4.17)

where ' and c' are the shear strength parameters on the slip surface AB and bi is the horizontal distance between AB. As stated in Section 4.3.1, the effective normal force and the shear force on the inter-slice boundaries are also in a state of limiting equilibrium with factor of safety equal to unity, so that Mohr-Coulomb criterion gives:

X i = E i tan i + cid i X i +1 = E i+1 tan i +1 + c i+1 d i +1

(4.18)

(4.19)

where di and di+1 are the length of AD and BC respectively and' andc' are the shear strength parameters on the inter-slice boundaries. In non-homogenous slopes, these parameters are the weighted average values for the materials and are discussed in Chapter 6. Defining = tan-1kc, then from the equilibrium of the slice i and using Equations 4.17, 4.18 and 4.19, see Sarma (1979):

E i+1 sec i +1 cos( i i + i +1 i +1 ) =

Wi sec sin ( i i ) + E i sec i cos i i + i i + ci Li cos i cid i sin ( i i i ) + ci+1 d i +1 sin ( i i i +1 ) + PWi cos( i i i ) PWi +1 cos( i i i +1 ) U i sin i

(4.20)

Ei +1 = Ei+1 + PWi +1

(4.21)

84

N i sec i cos i i + i +1 i +1 =

( PW sin (
i

Wi sec cos i +1 i +1 + + E i sec i sin i +1 i +1 i + i + i +1 + i

ci Li sin i +1 i +1 i c id i cos i +1 i +1 + i + c i+1 d i +1cos i +1 +


i +1 i +1

) ) PW

sin i +1 U i cos i +1 i +1 i

(4.22)

N i = N i + U i

(4.23)

Thus, if two angles, i and i+1 are assumed known, then the geometry and force unknowns can be determined from Equations 4.20, 4.21, 4.22 and 4.23. However, two additional equations have to be set up to obtain i and i+1, which cannot be obtained from the traditional limit equilibrium method. Another assumption, termed the acceptability criterion is used to solve this problem.

4.4.2 The acceptability criteria


There are two criteria that must be satisfied if the obtained solution is to be acceptable. The first is a criterion in terms of stresses and the second is a criterion in terms of kinematics. The details about the acceptability criteria have been presented in Chapter 2. The stress acceptability criterion exists because the slice is made up of soil, which has a limited strength. The difference between the stress characteristics method (Sokolovski,1960) and the current method and is that in the former, the entire soil mass is in limiting equilibrium while in the present method, the limiting equilibrium of stresses are only along the slip surface and the inter-slice surface; the rest of the soil may exist in any state prior to failure. When the slope fails and starts to move, it has to be kinematically acceptable. Then the slip surface has to be concave upward, which means that

1 2 3 i n
where i is the angle made by the slip surface with the horizontal in the i th slice.

(4.24)

85

4.4.3 The internal plane and m in the standard slice


To define two angles i and i+1, the standard slice is divided into two segments by the line BD. The segment ABD is named the slip surface side and the segment BCD is named the inter-slice boundary side (Figure 4.5). The reciprocal of the factor of safety of a plane inclined at an angle is termed m. For the convenience of plotting, this is used throughout the analysis.

D Xi Ei di F EL X L A i
slip surface side inter-slice boundary side

C C' kcWB WB Xi+1 i+1 di+1 Ei+1

kcWi Wi

E X

N
Layer 1: weak material

A' S T kcW WS i li

Ti Ni

Layer 2: strong material

F'

Figure 4.5 The internal plane and forces acting on it

Slip surface side A plane inclined at an angle to the slip surface creates a small segment ABA' inside the slipping mass as shown in Figure 4.5. The reciprocal of the factor of safety of the plane BA', mS, is defined as:

S m =

(N

S U ) tan S + c L

(4.25)

where N and T are the normal and shear forces in terms of total stresses acting on the plane BA', U is the force due to pore water acting on the plane, L is the length of BA' and S and cS are the weighted average strength parameters on the plane.

86

From the equilibrium of the segment ABA' (normal and orthogonal to plane BA) : N = N i cos + Ti sin WS sec cos( i + ) + PWL ) sin ( i + i ) + X L cos( i + i ) (E L T = N i sin + Ti cos WS sec sin ( i + ) + + PWL ) cos( i + i ) + X L sin ( i + i ) (E L

(4.26)

(4.27)

where effective normal effective force E'L and shear force XL on the small part of side boundary AA are defined by Equations 4.38 and 4.40 later in Section 4.4.4. PWL is the force due to pore water pressure acting on the plane AA as defined in Equations 4.67. WS is the weight of the small segment. From the geometry of the small segments ABA', L and WS can be obtained as: cos ( i + i ) cos ( i + i )

L = Li

(4.28)

WS =

cos ( i + i )sin 1 1 Li +1 L sin = L2 i +1 2 2 cos ( i + i )

(4.29)

/ 2 + i + i / 2 i i
i
Li

Figure 4.6 Geometry of the segment ABA in the slip surface side

87

Inter-slice boundary side Similarly, considering a plane inclined at an angle to the inter-slice boundary BC creating a small segment BCC' inside the slice, the reciprocal of the factor of safety of the plane BC', mB, is defined as:

mB =

(E

B PW ) tan B + c D

(4.30)

where E and X are the normal and shear forces in terms of total stresses acting on the plane BC', PW is the force due to pore water acting on the plane, D is the length of BC' and B and cB are the weighted average shear strength parameters of the plane. Note that due to the sign conventions of the forces used in the analysis, mS or mB, computed on an internal plane from either side are of opposite sign. From the equilibrium of the segment BCC': E = Ei +1 cos + X i +1 sin WB sec sin ( i +1 )

(4.31)

X = Ei +1 sin + X i +1 cos + WB sec cos( i +1 )

(4.32)

where WB is the weight of the small segment. Again, from the geometry of the small segment BCC', D and WB can be obtained as:

D = d i +1

cos( i +1 + ) cos( i +1 + )

(4.33)

WB =

cos( i +1 + ) sin 1 1 d i +1 D sin = d i2+1 2 2 cos( i +1 + )

(4.34)

88

/ 2 + i +1 +

/ 2 i +1

d i +1

i +1

Figure 4.7 Geometry of segment BDD

4.4.4 Effective normal stress distribution


In traditional methods of slices within the limit equilibrium framework, the effective normal stress distributions along the slip surface and the inter-slice boundaries are not necessary to be known for the derivation of the factor of safety or the critical acceleration. However, in the new procedure, the stress distribution is essential for two reasons: first, to obtain N and T in Equations 4.26 and 4.27, it is necessary to know the contribution of the effective normal force EL and shear forces XL on the small part of slice boundary AA' (Figure 4.5), which therefore requires the information about the stress distribution on plane AD. Second, for any plane in nonhomogeneous slopes, the effective normal stress distribution is needed to define the weighted average angle of shearing resistance. Throughout this thesis, we assume that the effective normal stress at point A, i, is equal on both planes AB and AD (Figure 4.5), since the two planes are in limiting equilibrium. The effective normal stress at the starting point of slip surface in the first slice, 1 (point O in Figure 4.4), can be obtained from the Mohr circle as:

= c1 cos 1 1

(4.35)

89

If the number of slices is taken large enough, in other words, the length of slip surface in each slice is small enough, the linear effective normal stress distribution on the slip surface is a good approximation. Then, the effective normal stress at the end point of the slip surface of the slice can be expressed as:

i = 2 N i1 / Li 1 i1

(4.36)

Then the effective normal stress at point A on plane AD (Figure 4.5) is known as it is assumed equal to the stress i. For simplicity, the effective normal stress distribution is assumed as linear along plane AD. Then the normal effective stress at point D on plane AD can be computed as

i 0 = 2 E i / d i i
Then, E'L and X'L on the plane AA (Figure 4.5) can be defined using the following equations:

(4.37)

= il EL

( )l
i i
0

2d i

(4.38)

+ PW L EL = EL
tan i + cil X L = EL

(4.39)

(4.40)

However, if there is no surcharge on the slope surface, the effective normal stress at the point D (Figure 4.5), can also be obtained from Mohr circle analysis which gives:

i 0 = ci 0 cos i 0

(4.41)

where i0 and ci0 are the soil parameters just below the point D. Generally speaking, the values of 'i0 computed by Equation 4.37 and Equation 4.41 are different, which means that the effective normal stress distribution along plane AD is not linear. However, it is shown later in Equations 4.48 through 4.51 that only differential forms of effective normal force EL and shear forces XL are needed in the analysis for homogenous slopes. The analysis for homogenous slopes are dependent of the value of i but independent of the stress distribution along plane AD.

90

In the analysis for non-homogenous slopes, the actual values of effective normal force EL and shear forces XL are needed. Therefore, the equations used in non-homogenous slopes (discussed in Chapter 6) are dependent of stress distribution of plane AD. A parabolic distribution discussed in Appendix A may be an alternative assumption for the analysis in non-homogenous slopes. However, the linear effective normal stress distribution is assumed on any plane of the slice in this thesis. This greatly simplifies the procedure, especially for the calculation of weighted average friction angle and it is further discussed in Chapter 6.

4.4.5 Application of acceptability criterion


The acceptability criterion for homogenous slopes can be expressed as the following. It is known that when =0, the internal plane is either the slip surface or the inter-slice boundary with m=1. Considering an angle , positive as shown in Figure 4.5 within the slip surface side or the interslice boundary side, it is expected that m1 for an acceptable solution. It is also expected that if a negative angle is selected, i.e. a thin section into the body of the slope or into the next slice, m1. Therefore, it is expected that m is maximum and equal to unity when =0. The condition can be expressed as:

dm / d = 0 when = 0
m=1 when = 0 Applying the acceptability criterion to slip surface side Differentiating Equation 4.25, we can obtain:

(4.42)

(4.43)

S m S m

S ( N U ) tan S + c L

=0

=0

=0

=0

tan S =0

=0
S

)+ c

=0

+ =0 L T = =0

(4.44)

S m

( N U ) =0 sec 2 S =0

=0

+
=0

S c

=0

=0

91

As stated above,

S m

S = 0 and m

=0

=0

= 1 . As to the differential form of soil parameters, such

as

and
=0

S c

, they are assumed to be zero. It is true for homogenous slopes because


=0

the soil parameters are constant. In non-homogenous slopes, another methodology is adopted so the minor effect of differential form of soil parameters is ignored in current stage. Equation 4.44 becomes:

=0

tan S =0

=0

)+ c

=0

=
=0

(4.45)
=0

Differentiating Equation 4.26 and 4.27, we can obtain:

=0

WS sec cos( i + ) N i sin + Ti cos E L S cos( i + i ) W sec sin ( i + ) sin ( i + i ) + E L = PWL sin ( + ) + PW cos( + ) + X L cos( + ) + i i L i i i i X L sin ( i + i ) =0

(4.46)

=0

WS sec sin ( i + ) + N i cos Ti sin E L S sin ( i + i ) W sec cos( i + ) + cos( i + i ) + E L = + PWL cos( + ) + PW sin ( + ) + X L sin ( + ) i i L i i i i X L cos( i + i ) =0

(4.47)

The differential values and the values of EL and XL at =0 are:

E L

=
=0

iLi cos( i + i )

(4.48)

=0 = 0 EL

(4.49)

92

X L

=
=0

iLi Li tan ii + cii cos( i + i ) cos( i + i )

(4.50)

X L =0 = 0

(4.51)

wherec'ii is average effective cohesion and'ii is average effective friction angle on the plane AA. The force due to pore water pressure on the plane AA, PWL, is defined later in Equation 4.67. It is noted that the differential form of E'L and XL are independent of the stress distribution of plane AD, then:

=0

iLi 1 = Ti L2 sin ( i + i ) i +1 sec cos( i + ) 2 cos( i + i )


sin ( i + i ) +
(4.52)

cos( i + i )

u i Li

iLi Li tan ii + cii cos( i + i ) cos( i + i ) cos( i + i ) T

=0

1 = N i L2 i +1 sec sin ( i + ) + 2
(4.53)

iLi u i Li cos( i + i ) + cos( i + i ) + cos( i + i ) cos( i + i )


iLi Li tan ii + cii sin ( i + i ) cos( i + i ) cos( i + i )

The force due to pore water pressure on the plane AA, U, is discussed later in Equation 4.68, then:

93

1 N i + Ti tan i L2 i +1 sec sec i sin (i i ) 2 iLi sec i sec ii sin (i + ii i i ) cos( i + i ) cii u i Li sec i cos(i i i ) + cos( i + i ) Li sec i sin (i i i ) ci Li tan ( i + i ) cos( i + i )
(4.54)

u i Li 1 Li tan i + (u i + u i +1 )Li tan ( i + i ) tan i = 0 2 2d i cos( i + i )


where ui is the pore water pressure at point A and ui+1 is the pore water pressure at point B, which is discussed in Section 4.4.6. Applying acceptability criterion to inter-slice boundary side Differentiating Equation 4.30, we can obtain:

mB mB mB

(E PW ) tan B + cB D
=0

=0

E =0

=0

PW

tan B =0

=0
B

)+ c
=0

=0

+ =0
=0

(4.55)

(E PW ) =0 sec 2 B =0

=0

cB

D
=0

X =

=0

Using the same assumption as used in the slip surface side, Equation 4.55 becomes:

=0

PW

tan B =0

=0

)+ c

=0

=
=0

(4.56)
=0

Differentiating Equation 4.31 and 4.32, we can obtain:

94

=0

WB E sin X cos sec sin ( i +1 ) + + i +1 i +1 = B ( ) W sec cos i +1 =0

(4.57)

=0

WB E cos X sin sec cos( i +1 ) + + i +1 i +1 = B ( ) W sec sin i +1 =0

(4.58)

The force due to pore water pressure on plane BC, PW, is discussed in Equation 4.73, then:

Ei +1 + X i +1 tan i +1 ci +1 d i +1 tan ( i +1 + ) 1 2 d i +1 sec sec i +1 cos i +1 i +1 + ci+1 d i +1 tan ( i +1 + ) + 2 1 u i +1d i +1 tan ( i +1 + ) tan i +1 = 0 2
It is noted that there is no stress distribution needed in the inter-slice boundary side.

(4.59)

4.4.6 The definition of pore water pressure


It is known that the most realistic position of the critical slip surface is obtained when effective strength parameters are used in the analysis. Effective strength parameters are only meaningful when they are used in conjunction with pore water pressures. Due to the importance of pore water pressures in a stability analysis, there are various ways of specifying the pore water pressure conditions. This section gives an overview of the available options, and presents one of the options incorporated in the analysis in the current stage. Piezometric line One way of defining pore water pressure conditions is with a piezometric line. If a slip surface is assumed with known pore water pressure informations, it is possible to construct a piezometric line, which represents the pore water pressure at a point on the slip surface by the difference in a level (the piezometric head). It will be appreciated that the position of the piezometric line is

95

related to the initially assumed slip surface. However, it cannot be used to describe a pore water pressure regime of any complicity (Bromhead, 1992). Ru Coefficients The pore-water pressure ratio Ru is a coefficient that relates the pore water pressure to the overburden stress. The coefficient is defined as:

Ru =

u tHs

(4.60)

where t is the total unit weight, u is the pore water pressure at any point and Hs is the height of the soil column. One of the difficulties with the Ru concept is that the coefficient varies throughout a slope. It is necessary to establish Ru at a number of points and then by some weighted averaging method to calculate one single overall average value for the slope (or a slice). The result is that the simplicity of the method is lost. More seriously, this results in a change of the computed factor of safety: deep-seated failure modes have higher factors of safety while shallow failure modes tend to have lower factors of safety than they actually have (Bromhead, 1992). However, because it is easy to express the pore water pressure in analytical form, it is incorporated in the analysis in the current stage. Even that, we have to recognize that it is difficult for practical use. The pore water pressure at points A and B can be expressed as (Figure 4.5)

u i = Ru t hi u i +1 = Ru t hi +1

(4.61)

(4.62)

The forces due to pore water pressures Ui, PWi and PWi+1 (Figure 4.5), on the slip surface and the inter-slice boundaries respectively, are known once the boundaries are defined:
1 (u i + ui +1 )li 2

Ui =

(4.63)

96

PWi =

1 ui d i 2 1 u i +1 d i +1 2

(4.64)

PWi +1 =

(4.65)

As the pore water pressure of point A' can be expressed as:

l u A = ui 1 d i

(4.66)

In the slip surface side, the forces due to pore water pressures PWL and U on the AA' and the A'B respectively, are known once the boundaries are defined:

PWL =

1 ui + ui 2

l 1 h i

l + ui +1 L

(4.67)

U =

l 1 (u A + ui +1 ) L = 1 ui 1 2 2 di

(4.68)

Then:

PWL

=
=0

u i Li cos( i + i )

(4.69)

PWL

=0

=0

(4.70)

=
=0

ui Li 1 Li (ui + ui +1 ) Li tan ( i + i ) 2d i cos( i + i ) 2

(4.71)

=0

1 (ui + ui+1 ) Li 2

(4.72)

97

In the inter-slice boundary side, the forces due to pore water pressures PW on plane BC' is known once the boundaries are defined: 1 u i +1 D 2

PW =

(4.73)

Then:

PW

=0

1 = u i +1 d i +1 tan ( i +1 + ) 2

(4.74)

PW

=0

1 u i +1d i +1 2

(4.75)

Pore water pressure based on flow nets


The net, which encompasses the intended point, is identified first and then the pore water pressure of that point is computed by interpolating the value of the corner points. The power of this approach is that the pore-water pressures can have any irregular distribution.

Finite element computed pore water pressures


To use the finite element computed pore-water pressures, it is necessary to find the element that encompasses the intended point. Next, the local element coordinates are determined at the point. The pore water pressures are known at the element nodes or directly from the gauss point values. And the inherent finite element interpolating functions are then used to compute the pore-water pressure at the intended point.

Comments on the definition of pore water pressure


In the above methods, the forces due to pore water pressure can only be expressed analytically with constant Ru coefficient and then the differential form of the forces can be expressed analytically. It is the only option available in the current stage. For other options, the pore water

98

pressure and the induced forces can only be expressed numerically and numerical difference has to be used. They can be incorporated in the analysis without any difficulty for the future research.

4.4.7 Line of thrust and moment equilibrium


Line of thrust
As stated in Section 4.4.4, the effective normal stress distributions along the inter-slice boundaries are assumed to be linear. Then the stress at an arbitrary point on the plane AD, A (Figure 4.5) can be expressed as:

= i

( ) l
i i 0

di

(4.76)

The point of E'i acting on the plane AD, z'i can be obtained by taking moment about A:
di

l = Eiz i

(4.77)

then:

z i =

id i2 + 2 i 0 d i2
6 Ei

(4.78)

The point of Ei acting on the plane AD, zi can be expressed as:

z i = (E i z i + PWi z w ) / E i

(4.79)

where zw is the point of PWi acting on the plane AD. After a slice is solved, the acting point of force E'i or Ei can be obtained. Then after the slip surface is defined slice by slice, the line of thrust can be obtained by connecting all the acting points with a straight line.

99

Moment equilibrium
In the new procedure, moment equilibrium is not prerequisite to derive non-linear equations to determine the unknown slices. The moment equilibrium can be satisfied by specifying the acting points of all the forces within a single slice. The governing equation is obtained by taking moments of all the forces on the ith slice about the corner point A (Figure 4.5) as stated by Sarma (1979). This gives:

N i l i X i +1bi sec i cos( i + i +1 ) + E i +1 [z i +1 + bi sec i sin ( i + i +1 )] E i z i Wi x gi xi + k c y gi y i = 0

(4.80)

in which (xgi, ygi) are the coordinates of the centre of gravity of the slices. The point of application of the Ni is given by li measured along the slip surface and the point of application of Ei is given by zi measured along the inter-slice boundary from point A (Figure 4.5). After the slip surface is found, we can start from the first slice, where z1=0. By obtaining zi from Equation 4.79, li can be determined. Proceeding to the last slice, the last ln is determined from the moment equilibrium of the last slice. Based on the acceptability criteria stated in Chapter 2, zi and li should lie within the slice, preferably in the middle third of the corresponding plane. This is further examined in Chapter 5.

4.5 THE ANALYSIS OF THE FIRST SLICE


The geometry and forces acting on the first slice are shown in Figure 4.8. Because the plane AC is the free surface, there is no forces contribution on the small plane AA, which is different from the standard slice. The analysis for the first slice is given in details in Appendix B. As shown in Figure 4.8, Two angles, 1 and 2, are needed to define point B (Equations 4.15 and 4.16 are still valid with replacing i by 1). Four unknown forces, the effective normal and shear forces on the slip surface, N'1 and T1, and the effective normal and shear forces on the inter-slice boundary, E'2 and X2, are the four unknowns in terms of forces to be solved (Equations 4.20, 4.21, 4.22 and 4.23 are still valid with replacing i by 1).

100

A' 1
kWS WS

C' X E
kWB WB

kW1 N W1 2 B

X2 E2

N1

T1

Figure 4.8 Forces acting on the first slice. The total forces N and E are shown, which are the sums of the effective force and the force due to pore water.

From the Appendix B, the two equations used to solve the first slice are:
L1 1 2 L1 sec sec 1 sin (1 1 ) + tan ( 1 ) 2

N1 + T1 tan 1 c1

L1 1 u2 tan 1 = 0 2 tan ( 1 )

(4.81)

d 2 tan ( 2 + ) E 2 + X 2 tan 2 c 2

1 2 d 2 tan ( 2 + ) + d 2 sec sec 2 cos( 2 2 + ) c 2 2 1 u 2 d 2 tan ( 2 + ) tan 2 = 0 2

(4.82)

wherec'1 and'1 are the weighted shear strength on the inter-slice plane BC, c'1 and '1 are the strength parameters on the slip surface, u2 is the pore water pressure at point B.

101

4.6 THE ANALYSIS OF THE LAST TWO SLICES


The last two slices are solved simultaneously as shown in Figure 4.9. The reason for solving the last two slices together is for convergence consideration and this will be considered in detail in Chapter 5. Note that a different critical acceleration, kgend, applies to the last two slices, which is an unknown to be solved. We consider as if the whole slope is divided into n slices. The last two slices are (n-1)th slice and nth slice. The (n-1) slice is the one which does not satisfy the kinematical acceptability for the first time due to the effect of the corner point (more details are given in Chapter 5).

C D kendWn-1 Wn-1 N 1 A' A n-1 Xn En Xn-1 En-1

C' X kendWn E' E Wn N2 T2 n n B Tn-1 Nn-1

F'

F En+1 hc Tn E Nn

Ln

1 n-1 T

Figure 4.9 Forces acting on the last two slices. The total forces N and E are shown, which are the sums of the effective force and the force due to pore water.

4.6.1 The geometry and forces acting on the last slice and initial conditions
It is assumed that there is a vertical tension crack with height hc in the last slice, so if n-1, n and

n are assumed, the shape of the last two slices can be defined (Figure 4. 9). Equations 4.15 and

102

4.16 are still valid to define the position of point B with replacing i by n-1.The position of E can be defined as following:

y n +1 = ys n +1 hc x n +1 = x n + ( y n +1 y n ) / tan n

(4.83)

(4.84)

In the (n-1)th slice, the unknown forces are the effective normal and shear forces on the slip surface AB, N'n-1 and Tn-1, and the effective normal and shear forces on the inter-slice boundary

BC, E'n and Xn, which can be obtained from Equations 4.20 to 4.23 with replacing i by n-1 and
replacing by end, where end=tan-1kend. In the nth slice, we assume there is a tension crack EF. If the assumed critical acceleration is used, it ends up a resultant force, E'n+1, which is the effective normal force acting on the plane EF. The acceptable critical acceleration for a given starting point is the one that makes E'n+1 equal to zero. Therefore, there are three unknown forces in the last slice, which are the effective normal and shear forces on the slip surface BE, N'n and Tn, and the effective normal on the inter-slice boundary EF, E'n+1. The forces can be obtained from the analysis similar to section 4.4.1.2 as:

sec n cos( n n ) =Wn sec end cos end + E n sec n sin n + n Nn d n cos n + PWn sin n U n cos n c n Ln sin n c n

(4.85)

+Un Nn = Nn
tan n + c Tn = N n n Ln
+1 cos( n n ) = Wn sec end sin ( n n end ) + En sec n cos n n + n n + En Ln cos n cn d n sin ( n n n )+ cn

(4.86)

(4.87)

(4.88)

n n ) PWn +1 cos ( n n ) U n sin n PWn cos( n +1 + PWn +1 E n +1 = E n

(4.89)

103

4.6.2 The application of acceptability criteria into the last two slices
The acceptability criterion is applied to the two slip surface AB and BE and the unknown shear surface BC to get three equations. The first two equations are exactly same as Equations 4.54 and 4.59 with replacing i by n-1 and replacing by end. The reciprocal of factor of safety of any plane BE within section BEF can be expressed as:

S m =

(N

S U ) tan S + c L

(4.90)

Differentiating the equation, we can obtain:

S m S m

S L (N U ) tan S + c

=0

=0

N =0

=0

tan S =0

=0
S

)+ c
=0

=0
S c

=0 T = =0

(4.91)

S m

( N U ) =0 sec 2 S =0

=0

L
=0

=0

Using the same assumption as Section 4.4.5.1, Equation 4.91 becomes:

=0

tan S =0

=0

)+ c

=0

=
=0

(4.92)
=0

Based on the force equilibrium of the slice n, the normal force and shear force acting on internal plane BF' can be expressed as:

N = N n cos Tn sin W cos( n + ) + k end W sin ( n + )


T = N n sin + Tn cos W sin ( n + ) k end W cos( n + )
Differentiating the Equations 4.93 and 4.94, we can obtain:

(4.93)

(4.94)

104

W N cos( n + ) + W sin ( n + ) = N n sin Tn cos W sin ( n + ) + k end W cos( n + ) + k end W T sin ( n + ) W cos( n + ) = N n cos Tn sin W cos( n + ) + k end W sin ( n + ) k end
Based on the geometry of the last slice:

(4.95)

(4.96)

l = Ln

sin sin ( n + ) sin n sin ( n + )

(4.97)

L = Ln

(4.98)

W =

1 2 sin n sin sin Ln + Ln hc 2 sin ( n + ) sin ( n + )

(4.99)

Using the definition of pore water pressure in section 4.4.6, the force due to pore water pressure and the differential form of the force can be expressed as: 1 (u n + u n +1 ) L 2

U =

(4.100)

U Ln 1 = (u n + u n +1 ) 2 tan n

(4.101)

Pore water pressure un+1 is due to the hydrostatic force because the tension crack may be filled with water:

u n +1 = w hc

(4.102)

105

The extent to which the crack is full with water, is expressed by a number between zero and unity. Zero means that the crack is dry and unity means that the crack is completely full of water. However, considerable thought has to be given to apply such a significant fluid pressure in a tension crack because that the crack can hardly hold the water at any time and the volume of the water existing the crack is very small so that a slight lateral movement may result in the disappearance of the associated force, especially during earthquakes, provided that there is no supply of water to the tension crack. By combining Equations from 4.94 through 4.104, we can obtain:

1 2 Ln + sin ( n n ) N n + Tn tan n 2 Ln + sin hc sec sec n n Ln L 1 + cn (u n + u n+1 ) n = 0 tan n 2 tan n

(4.103)

Since then, we have three equations to solve the three unknowns n-1, n and n. As stated before,

E'n+1 should be zero due to the existence of tension crack. kgend can be varied to satisfy E'n+1 equal
to zero. We have another unknown, kgend, and another equation:

+1 = 0 En

(4.104)

In the last two slices, we have four unknown and four equations. Therefore, the last two slices are solvable.

106

5. The Application of Acceptability Criteria to Homogenous Slopes

5.1 SCENARIO
The formulation of the new procedure has been developed in Chapter 4. In this chapter, its application to homogenous slopes is presented and the following points are discussed in detail. After solving for the first slice (the toe end), the results are compared with stress analysis using the Mohr circle. Because of the great non-linearity of the equations to be solved, an appropriate solution procedure has to be found. As shown in Section 4.4.1.1, to define the end point of the slip surface in the new slice, B (Figure 4.4), a horizontal increment x has to be assumed. Its effect on a single slice and on a whole slope is studied in Section 5.4. Finally, a simple and rational procedure to determine the global critical slip surface is proposed and examined.

5.2 THE FIRST SLICE


The first slice of the sliding mass based on the methods of slices can be seen as a passive wedge in the sloping ground. The geometry of the passive wedge under seismic loading can also be obtained from the Mohr circle solution (Appendix C), which gives us a first opportunity to validate the new procedure. The geometry of an arbitrary passive wedge on the sloping ground is shown in Figure 5.1. 1 and

2 (Figure 5.1) can be obtained using equations 4.81 and 4.82 based on the new procedure shown
in Chapter 4, which involves no assumptions about the stress distribution on the free surface. Then the two angles, 1 and 2 (Figure 5.1), which are the inclination angles of the two failure planes relative to the ground surface, can be obtained as:

1 = + 1

(5.1)

107

2 =

(5.2)

where is the inclination angle of the slope.

1 1
H

2
B

Figure 5.1 The geometry of a passive wedge on the sloping ground

After the geometry of the slice is obtained, H, which is the vertical distance between point B and the ground surface, can be defined. If the normal and shear stress on a plane parallel to the ground surface at point B are approximated as Equations C.2 and C.3, the two inclination angles 1 and

2 can also be obtained from Mohr circle solutions as shown in Appendix C. A comparison of the
results from the new procedure and the Mohr circle solutions is shown in the following example. Example 5.1 A sloping ground surface with an inclination angle equal to 150 is considered, which is subjected to 0.1g horizontal acceleration from an earthquake. The soil under the ground is first considered as cohesionless, dry, homogenous and isotropic with = 300 and = 20kN/m3. If the increment

x is assumed to be unity, from Equations 4.81 and 4.82, the two angles 1 and 2 are found to be
17.850 and 42.150, respectively, and H is 0.25456 m. The same geometry, soil parameters and the value of H are input into Equations C.4 and C.5 to obtain the Mohr circle solutions. The resulting solutions for the two angles are exactly the same. Thus, it is interesting to see that, at least for the

108

first slice and in cohesionless soils, the geometry of the slip surface and the inter-slice boundary obtained by the new procedure obeys the Rankine state. This is an encouraging sign. It is shown in Appendix C that the Mohr circle solutions for failure planes in cohesionless soil are independent of the stress assumption. In the next step we check the results for cohesive soils. The geometry of the slope, the horizontal increment x, the applied critical acceleration and the unit weight of the soil are kept the same as above but the soil under the sloping ground is assumed to be cohesive with c = 10kPa and = 300. From Equations 4.81 and 4.82, the two angles 1 and 2 are found to be 28.490 and 31.510 respectively, and H is 0.30854 m. The same geometry, soil parameters and the value of H are input into Equations C.4 and C.5, and the corresponding two angles are found to be 27.270 and 32.730, respectively. This shows that there is a small difference between the two solutions for cohesive soils. To examine the cause of the discrepancy, different horizontal increments x are used, which result in different values of H. There are two sets of angles 1 and 2 corresponding to each value of H, one from the new procedure and the other from the Mohr circle solutions. They are plotted in Figure 5.2. It is seen from Figure 5.2 that with a decrease in the value of H, the difference between the two sets of angles becomes smaller. The difference appears to come from the stress assumption on the plane parallel to the ground surface at point B used in the Mohr circle analysis. With a reduced H, implying less effect of the stress components used, the two sets of solutions approach identical results.

109

Figure 5.2 Two sets of angles 1 and 2 corresponding to different values of H, N represents the solution from the new procedure and M represents the solution from the Mohr-circle solutions

5.3 METHODS TO SOLVE THE NON-LINEAR EQUATIONS


According to the analysis of Chapter 4, the geometry of a standard slice is obtained by solving two non-linear equations (4.54 and 4.59) simultaneously. The equations are solved using the Trust-Region Dogleg Method (Powell, 1970) available within MATLAB. The method is a modified version of Gauss-Newton method, with improved robustness when starting far from the solution. Newtons method for approximating the solution vector, p, to a set of nonlinear equations:

F ( x ) = ( f1 , f 2 ) = 0

(5.3)

is by assuming a starting point, computing the values of equations and evaluating the gradient at that point, where f1 represents Equation 4.54 and f2 represents Equation 4.59. In each iteration, the

110

change in a component fi with respect to the change in the variable xj is described by the partial derivative

fi , where x1 represents i and x2 represents i+1. xj

Then the matrix:

f1 x J (x ) = 1 f 2 x1

f1 x2 f 2 x2

(called the Jacobian matrix) is computed either by analytical derivatives, if available, or by finite difference approximation. As a sequence, Newtons method for finding the solution p to the nonlinear equations F(x) = 0 has the form:

p (k ) = p (k 1) J p (k 1)

[(

)] F ( p ( ) )
1
k 1

(5.4)

Then an iteration procedure can be set up until the values of the functions are less than the termination tolerance.

5.3.1 The determination of the Jacobian matrix


The Jacobian matrix can be obtained by analytical derivatives, if they are available, and can also be approximated using finite differences. For the slice to be solved, the analytical Jacobian matrix is only available for dry homogenous slopes with simple geometry. The Jacobian matrix can also be determined by the finite differences procedure available in MATLAB. The following example is used to check the accuracy and robustness of the finite difference approximation of the Jacobian matrix in MATLAB. Example 5.2 The geometry of an example slope and its soil properties are shown in Figure 5.3. The slip surface is solved slice by slice until an arbitrary slice, chosen here to be the nineteenth slice but it

111

could be any slice. The geometry and forces acting on the nineteenth slice are listed in Table 5.1 as initial conditions. The twentieth slice shown as dashed lines in Figure 5.2 is first solved using the analytical Jacobian matrix. i and i+1, the geometry and the forces acting on the twentieth slice in each iteration are listed in Table 5.1. Then the twentieth slice is re-solved using the finite differences approximation of Jacobian matrix provided in MATLAB and the same information at each iteration is listed in Table 5.1. It can be seen from Table 5.1 that the difference in the solutions from the analytical Jacobian matrix and the finite difference approximation of the Jacobian matrix is very small and can be neglected. The above example shows the accuracy and the robustness of the finite difference approximation of the Jacobian matrix in MATLAB. In more complicated situations, for example, in homogenous slopes with pore water pressures or in non-homogenous slopes, no analytical Jacobian matrix can be obtained. Therefore, the finite difference approximation of the Jacobian matrix has to be used. Fortunately the experience of its use for homogenous slopes gives us the confidence to do so.

Figure 5.3 A general slice within an example slope

112

Table 5.1 Comparison of results of each iteration between analytical Jacobian matrix and finite difference Jacobian matrix

Initial conditions Solution from analytical Jacobian matrix Iteration 0 1 2 3 Finite difference Jacobian matrix Iteration 0 1 2 3

0.0441

0.4758

x (slip surface) 14.665

y (slip surface) -1.132

W 81.725

N 82.832

T 52.634

E 224.52

X 155.12

f1 ()2 + f2 ()2

0.0441 0.0519 0.0516 0.0516

0.4758 0.4685 0.4685 0.4685 15.5828 15.6305 15.6305 -1.0915 -1.0819 -1.0819 82.7308 84.6683 84.6683 81.9455 85.3743 85.3743 51.9049 54.1251 54.1251 235.8645 236.6559 236.6559 162.5644 162.8701 162.8701

41.4487 0.00820386 3.65278e-010 1.10218e-024

0.0441 0.0519 0.0516 0.0516

0.4758 0.4685 0.4685 0.4685 15.5828 15.6305 15.6306 -1.0915 -1.0819 -1.0822 82.7308 84.6683 84.6813 81.9455 85.3743 85.4147 51.9049 54.1250 54.1487 235.8645 236.6559 236.6931 162.5644 162.8701 162.8927

41.4487 0.00820385 3.65069e-010 7.03857e-025

* Where f1 represents Equation 4.54 and f2 represents Equation 4.59. The sum of square values of two functions for each set of and .

113

5.3.2 Validation of stress acceptability


After an arbitrary slice is solved, for example the twentieth slice in Example 5.2, the stress acceptability can be inspected. The variation of m, both in the slip surface side and in the interslice boundary side can be computed using Equations 4.25 and 4.30 as shown in Figure 5.4. Note that the negative value of m is because of the sign convention used in the two sides. It is seen that the assumption that m is the local extremum in the two boundaries makes, providing that the soil parameters are constant in the slice, any plane within a slice satisfy the stress acceptability.

Figure 5.4 The variation of m in the twentieth slice in Example 5.2

5.3.3 The initial approximation of i and i+1


The success of solving non-linear equations using the Gauss-Newton based method depends on the close approximation of the unknowns in the initial step. Even for the Trust-Region Dogleg Method (Powell, 1970), a good initial approximation is still needed to obtain a valid solution and to accelerate convergence.

114

As shown in Section 5.2, the solutions based on Mohr-circle analysis can be a very good initial approximation of the two angles in the first slice. If the number of slices is large enough, in other words, the horizontal increment in the ground surface, x (Figure 4.4) is small enough, the variation of i and i+1 in all slices is smooth. For example, the values of i and i+1 of the first twenty slices in Example 5.2 are shown in Figure 5.5. It is shown in Figure 5.5 that i and i+1 of the previous slice are good initial approximations of the angles of the current slice. Generally speaking, this statement is true.

Figure 5.5 The variation of i and i+1 in the first twenty slices in the Example 5.2

However, one special case has been identified. If the starting point of the slip surface is below the toe point in steep slopes, then for the slice immediately after the toe point, it is difficult to solve for i and i+1 starting from the previous solution. This is explained in the following example. Example 5.3 A slope of gradient 1:1.5 (H : V) is considered and is shown in Figure 5.6 along with its soil properties. The starting point of the slip surface is set at -1m below the toe point. If the horizontal increment in the ground surface, x, is assumed to be unity, then for a critical acceleration of 0.12g the first slice of the slope gives 1 and 2 as -17.820 and 61.540, respectively. Using these

115

values as the initial approximation to solve the two non-linear equations for the second slice, the result is shown as the dashed line in Figure 5.6. The slip surface is above the slope surface, which is obviously unacceptable. In such a situation, the initial approximation of i+1 to be input into the non-linear equations is reduced to a value of half of i of the previous slice. The above example is re-solved and the result is shown as the solid line in Figure 5.6, which is an acceptable solution. Generally speaking, the algorithm to solve non-linear equations using its inherent finite difference approximation of Jacobian matrix provided in MATLAB is robust and has the ability to handle an initial approximation far from the solution. However, there are some special cases where the physically acceptable solution depends on the initial approximation of the solution. Although it is rare, the user should be aware of its possibility.

Figure 5.6 Solutions from two different initial approximations of i and i+1 in the second slice

116

5.4 ACCEPTABLE SLIP SURFACE

5.4.1 Infinite slopes


An infinite slope of gradient 1:3 with c=15kPa, =250 and =20kN/m3 is taken as an example. For any starting point of the slip surface picked from the slope surface, if kcg is less than the value of tan ( ) , where is the inclination angle of the slope, the slip surface obtained from the new procedure will not end at the slope surface but any kcg larger than tan ( ) can produce a surface slice by slice using the procedure shown above. If the vertical distance between the starting point and the end point of the slip surface is defined as h, then the relationship between

kcg and h can be shown as in Figure 5.7.

Figure 5.7 The relationship between vertical distance and desired critical acceleration in the example infinite slope

It is clear from the figure that for infinite slopes, there is no unique solution for the critical acceleration but the critical acceleration can change from tan ( ) = 0.115 to . The slip surface for the smallest critical acceleration goes to an infinite height while for a large critical acceleration, the height tends to zero. The minimum solution is the same as for a cohesionless soil

117

with the same friction angle. But for a cohesionless soil, the slip surface is parallel to the ground surface while for a cohesive soil, it goes much deeper.

5.4.2 Finite height slopes


For a finite slope, the critical acceleration must, therefore, be smaller than the one with the same height in infinite slopes for the surface which ends in the corner point. It is found that using a small value of kcg the surface does not converge towards the crest of the slope, which is contrary to kinematical acceptability. Therefore, the desired critical acceleration is not acceptable. On the other hand, using a large kcg makes the slip surface end within the slope as in infinite slopes and therefore, obviously, that kcg is too large to be the minimum one as is shown in the following example. Example 5.4 A slope of gradient 1:5 is considered and its soil properties are shown in Figure 5.8. The pore water pressure coefficient, Ru, is assumed to be constant throughout the slope and equal to 0.2. If the desired critical acceleration is assumed to be 0.15g, the associated slip surface is shown as a solid line in Figure 5.8, which is against the kinematical acceptability after the top line of the slice passes the corner. Obviously the desired critical acceleration is not acceptable. If the desired critical acceleration is assumed to be 0.19g, the slip surface will end at 40m measured horizontally from the toe point and if the desired critical acceleration is assumed to be 0.21g, the slip surface will end at a point 32m measured horizontally from the toe point. Both of those critical accelerations are too large to be the critical one. The critical acceleration therefore lies between 0.15g and 0.19g.

118

Figure 5.8 The different slip surfaces produced by different values of kc

5.4.3 Acceptable critical acceleration


A procedure is presented here to change the desired critical acceleration. We consider that the whole slope is divided into n slices. A starting point is picked from the ground surface and any kcg is assumed, then solve slice by slice until the (n-2) slice. The (n-1) slice is the one which does not satisfy the kinematical acceptability for the first time. Then, the last two slices are solved simultaneously as shown in Section 4.6. It is assumed that there is a vertical tension crack with height hc in the last slice, so if n-1, n and n are assumed, the shape of the last two slices can be defined (Figure 4.9). The acceptability criterion is applied to the two slip surfaces AB and BE and the unknown shear surface BC to get three equations as shown in Chapter 4. To explain the above procedure, Example 5.4 is considered again to validate the robustness of the procedure. The starting point is set at the toe point and a relatively small initial approximation of the desired critical acceleration, 0.12g, is assumed. The corresponding value of kgend is shown in Figure 5.9. The new value of the critical acceleration for the next iteration is computed as:

k c g = k c g + k c /

(5.5)

119

where:

k c = (kg end k c g )
and is the convergence parameter. Note that for each iteration, the value of kc is updated and a new slip surface is produced based on the new value of kc. If the value of is set at 50, the kcg and kgend in each iteration are as in Figure 5.9. It is shown in Figure 5.9 that by assuming an initial

kc equal to 0.12g gives a kgend equal to 0.496g. After 20 iterations, both curves converge at a
single value, 0.17023g, which is the acceptable critical acceleration of the corresponding starting point. Based on an extensive study the value of can be taken as 50 in homogenous slopes or 200 in non-homogenous slopes (considered in Chapter 6).

Figure 5.9 The values of kc and kgend at each iteration

The first kcg is acceptable for all but the last two slices while kgend is acceptable for the last two slices. If both are equal, then the kcg is acceptable for the whole slip surface.

120

5.4.4 Initial value of the critical acceleration


It is shown above that a good initial estimate of the critical acceleration would accelerate convergence. However, the initial critical acceleration is unknown. Therefore, two procedures are proposed in the following to approximate the value. The better way to select a starting critical acceleration would be to select a possible slip surface from experience (either for homogenous or non-homogenous slopes) and find the associated critical acceleration for that surface using an established stability analysis method, e.g. Sarma (1973) or Sarma (1979). Note that the initial slip surface may not be acceptable in terms of stress acceptability. This value of the critical acceleration can be used as the initial value in the above convergence procedure. An automatic but less efficient procedure can also be adopted. A relatively large critical acceleration is adopted as the initial value. If the acceleration makes the slip surface end within the slope, the value of acceleration is reduced systematically until a value of critical acceleration can make the end point of the slip surface pass the corner point of the slope. The convergence procedure is then triggered as described above.

5.5 THE EFFECT OF INCREMENT x


As discussed in Section 4.4.1.1, the definition of point C in the unsolved slice depends on an assumed increment x, which is a horizontal increment on the slope surface (Figure 4.4). After the increment is assumed, the point C can be defined along with the slope surface function as shown in Equations 4.54 and 4.59. In the following section, the effects of increment x on a single slice and on the whole slope are discussed separately.

5.5.1 The effect of increment x on the solution of i and i+1 for a single slice
In the following section, the effect of increment x to the solutions of i and i+1 is discussed. Example 5.2 is considered again. Nineteen slices have been solved by assuming unit increment

x and the twentieth slice is seen as a general slice (it is an arbitrary slice). The geometry of the

121

slope, the soil parameters and the initial conditions have been given above. As we know, the width of the slope is 50m. Then the different values of increments are assumed to check the dependency of i and i+1 on the increment x. The results are shown in Table 5.2
Table 5.2 The effect of increment x on the values of i and i+1 for an arbitrary slice

x (the percentage of
the width of slope) 0.5% 1.0% The twentieth slice 2.0% 2.5% 5.0% 10%

i (in degrees) 2.80 2.85 2.96 3.00 3.27 3.76

i+1 (in degrees) 27.15 27.05 26.84 26.74 26.25 25.33

An important conclusion can be drawn based on the above example: the values of i and i+1 depend on the assumed increment x. When increment x increases, the value of i increases. This feature is important, especially when a slip surface develops from one soil layer to another in non-homogenous slopes, in which case the kinematical acceptability is normally violated if a constant increment x is assumed. If the slip surface obtained by assuming constant increment x is not kinematically acceptable, i.e. the value of i is less than i-1, the increment x is increased systematically until kinematical acceptability is satisfied. The details are given in Chapter 6. Since the assumed increment x will affect i, for a given distance, by assuming different x, the difference of the points resulting from different increments x needs to be examined. Example 5.2 is again used and the first nineteen slices have been solved. An arbitrary distance in the slope surface, for example 5m, is taken after nineteen slices. By assuming x as 0.5%, 1%, 2%, 2.5%, 5%, 10% of the width of the slope as shown in table 5.2, the number of slices needed to reach the given distance is 20, 10, 5, 4, 2 and 1, respectively. The resulting B is shown in Figure 5.10 for different assumed values of x. It is seen that the assumed increment x does affect i, but for a given distance it will give similar points to point B in Figure 5.10.

122

Figure 5.10 The resulting point B by assuming different x

5.5.2 The effect of increment x on the critical acceleration of a whole slope.


The effect of the assumed increment x on the computed critical acceleration of a whole slope also needs to be checked. It is generally known in slope stability analysis that the more slices are used, the more accurate factor of safety (or critical acceleration) is obtained. However, a larger number of slices is more computationally demanding in the limit equilibrium framework. Especially in the new procedure, the determination of each slice involves solving two non-linear equations. Therefore, the use of a large number of slices is not welcome. The following example checks what is the acceptable value of the increment x that can provide enough precision but also reduce unnecessary computation.

123

Example 5.5 Three slopes, all 10m high but with gradients of 1:1.5, 1:3, 1:5 are considered as examples. The soil properties and the starting point for each example slope are listed in Table 5.3. If the horizontal distance between the toe point and the corner point is defined as B, then N=B/x. In each example slope, four different increments x, which are: 1/20, 1/40, 1/50 and 1/100 of width of B, are used. The acceptable critical accelerations are listed in Table 5.2 for each example slope. It is noticed from Table 5.3 that N larger than 40 gives solutions that are accurate to three decimal points. On the other hand, for steep slopes with relatively small B, the use of large values of N causes numerical problems in some cases. This is because of extreme asymmetry in the last two slices. Therefore, N of about 40~50 is recommended in the new procedure, which provides sufficient accuracy and reduces unnecessary computation.

124

Table 5.3 The effect of increment x on the computed critical acceleration in three example slopes N* 20 kc (g) 0.34835 40 0.34993 50 0.35044 100 Numerical problems

Case 1 (1:1.5)

c (kN/m )
2

(degrees)
30

ru 0.2

Starting point (m) -1.5

20

Case 1 (1:3)

c (kN/m )
2

(degrees)
20

ru 0.4

Starting point (m) -3 kc (g)

N* 20 0.078474 40 0.074525 50 0.074101 100 0.073713

10

Case 1 (1: 5)

c (kN/m )
2

(degrees)
10

ru 0.0

Starting point (m) -5 kc (g)

N* 20 0.060678 40 0.059927 50 0.05978 100 0.059481

5
*

Note that the whole slip surface is divided in more slices than N because it includes some slices if the starting point is below the toe point and if

the end point is after the corner point. For example, in case 1, if x is taken as 0.75m, N = 15/0.75 = 20. However, there are 2 additional slices below the toe point if the starting point is set as -1.5m and 1 additional slice after the corner point. Thus, the actual number of slices is 20+2+1=23.

125

5.5 LINE OF THRUST


After an acceptable slip surface is obtained, the line of thrust can be determined as stated in Section 4.4.7. The application points of forces Ni can be computed based on moment equilibrium (Equation 4.80) or from assumed linear effective normal stress distribution along the slip surface. Rigorous solutions should make the above two sets of application points of forces Ni equal. As in the acceptability criteria stated in Chapter 2, the line of thrust and the application points of forces

Ni have to lie within the sliding mass, preferably in the middle third of the corresponding plane. It
is good practice to inspect the validation of the assumed stress distribution along the inter-slice boundaries. Example 5.6 An example slope of gradient 1:3 is shown in Figure 5.11 with its soil properties. If the starting point is set as -2m below the toe point the acceptable critical acceleration can be obtained as 0.039502g. The corresponding slip surface and inter-slice planes for each slice are shown in Figure 5.11. We define:

z _ pos (i ) = z (i ) / d (i ) %

where z_pos is the percentage of the application point Ei in the corresponding inter-slice plane. Similarly we can define:

l _ pos (i ) = l (i ) / L(i ) %

where l_pos is the percentage of the application points Ni in the corresponding slip plane of the slice. Then z_pos and l_pos for each slice are shown in Figure 5.12. Note there are two sets of l_pos: one from moment equilibrium and the other from assumed linear stress distribution along the slip surface. Figure 5.12 shows that for some slices the line of thrust falls outside the slices. This means that the linear stress distribution may result in tension in some slices.

126

To overcome this problem, a parabolic stress distribution can be used, as shown in Appendix A. The corresponding line of thrust is shown in Figure 5.11. The corresponding z_pos and l_pos can be obtained as above, and are shown in Figure 5.13. Figure 5.13 shows that the parabolic stress distribution results in an acceptable line of thrust. As shown in Appendix A, in homogenous slopes the stress distribution does not affect Equations 4.54 and 4.59, which are solved to obtain the geometry of the new slice. It is because of that only differential form of EL and XL are needed in the analysis of homogenous slopes. However, in nonhomogenous slopes, the actual values of EL and XL are required, meaning that the assumed stress distribution will affect the equations used to obtain the geometry of the new slice. However, the stress acceptability is checked after the slice is defined as in Equations 5.54 and 4.59 as shown in Chapter 6 for non-homogenous slopes. Therefore, for simplicity, linear distribution is used for the analysis of homogenous and non-homogenous slopes (initial step) throughout this thesis. Parabolic stress distribution is only used to obtain an acceptable line of thrust.

Figure 5.11 Lines of thrust resulting from different assumed stress distribution along the inter-slice boundaries

127

Figure 5.12 Points of application of Ni and Ei+1 resulting from linear stress distribution

Figure 5.13 Points of application of Ni and Ei+1 resulting from parabolic stress distribution

128

The application point of Ni can also be obtained from the assumed linear stress distribution along the slip surface. It is plotted as a dash-doted line in Figure 5.12 and 5.13 for comparison. It is seen from Figure 5.13 that parabolic stress distribution is still not an exact solution. Mathematically robust normal stress distributions along the inter-slice plane and slip surface, which results in the satisfaction of moment equilibrium, need to be researched further.

5.6 GLOBAL CRITICAL SLIP SURFACE


For each starting point an acceptable critical acceleration can be obtained and, finally, the slip surface with minimum critical acceleration is the critical slip surface. The choice of the starting points will depend on the problem. For example, in the presence of a load bearing foundation near a slope, the starting point for the critical surface may be on the ridge and not on the slope. While starting from a point on the slope, the slip surface may not see the foundation, another starting point, perhaps away from the ridge, will see the foundation. In homogenous slopes with simple geometry, the global critical slip surface can be found by starting from the toe point and then moving below the toe point with a specified increment. The acceptable critical acceleration for each starting point is stored and successive values are compared until an increase of the lowest value is encountered. Then the lowest value is the global critical acceleration and the corresponding slip surface is the critical slip surface. Example 5.4 is again used to explain the above procedure. The corresponding acceptable critical accelerations for the starting points of: 0, -1m and -2m are: 0.17024g, 0.16973g and 0.17059g, respectively. Then the slip surface with critical acceleration of 0.16973g, starting -1m below the toe point, is the global critical slip surface for the analysed slope. In homogenous slopes with berms, the determination of global critical slip surface becomes more complicated. Each berm of the slope can be treated as a mini slope and the above procedure is suited for the determination of the local critical slip surface with each berm. The local critical slip surface of each berm has to be compared with the slip surface of the whole slope to identify which is more critical. This is demonstrated in the following example. The flowchart to determine the critical slip surface in homogenous slopes is given in Figure 5.14.

129

START

Select a starting point

Assume a critical acceleration accordding to Section 5.4.4

Determine the slip surface slice by slice by solving Equations 4.54 and 4.59 until n-2 slice
No

Solve the last two slices together to obtain kgend

kcg = kcg + kc /

kcg = kgend ?

No

Yes

Are all the starting points analysed?

Yes

Output the critical slip surface and the corresponding critical acceleration

END

Figure 5.144 Flowchart of determining the critical slip surface in homogenous slopes

130

Example 5.7 Consider a two-step slope, the first step is a 1: 3 slope 10m high and the second one is a 1:5 slope also 10m high. The distance between points B and C is 10m. The soil parameters for this homogenous slope are shown in Figure 5.15 along with the geometry of the problem. The local critical slip surface for the first step of the slope starts at -3m below point A as shown by the dashed line in Figure 5.15 and the corresponding critical acceleration is 0.42073g. The local critical slip surface for the second step of the slope starts at -7m below point C, as shown by the dash-dotted line and the corresponding critical acceleration is 0.463g. The critical slip surface for the entire slope starts at -16m below point A as shown by the solid line and the corresponding critical acceleration is 0.38g. By comparing the three critical accelerations, the global slip surface is selected as the critical slip surface for the entire slope, as shown by the solid line in the Figure 5.15.

Figure 5.155 Local critical slip surfaces and the global slip surface

131

6. The Application of Acceptability Criteria to NonHomogenous Slopes

6.1 SCENARIO
This chapter considers the application of acceptability criteria to non-homogenous slopes to determine the critical slip surface. The algorithm to obtain an acceptable standard slice in nonhomogenous slopes is presented. The slice is initially defined using the same assumption as used in homogenous slopes and then the stress acceptability is inspected after the slice is obtained. Further constraints are applied to the slice if the stresses are not acceptable. As seen in Chapter 5, the horizontal increment on the slope surface, x, has an effect on the geometry of the slice. In the new procedure, the slip surface is obtained slice by slice and therefore, if the solution of the current slice does not satisfy the kinematical acceptability criterion, the value of x is increased systematically until the criterion is satisfied. Some uncertainties are involved when the slip surface encounters soil layer boundaries. A procedure is proposed to analyse different slip surface paths and then the critical one is picked. Finally, a complete methodology for the determination of the critical slip surface, both for homogenous and non-homogenous slopes, is presented and its practical application is shown.

6.2 STRESS ACCEPTABILITY FOR A STANDARD SLICE IN NONHOMOGENOUS SLOPES


A standard slice within a slope is shown in Figure 4.5 and m, which is the reciprocal of the factor of safety for an arbitrary plane in the slice, is defined as in Chapter 4. In homogenous slopes, the assumption that m is the local extremum in the two boundaries has been validated where the soil parameters are constant within the slice. However, in non-homogenous slopes, the soil parameters may vary in a single slice and therefore, the variation of soil parameters may result in the

132

maximum value of m in the slice exceeding unity. In such a case, stress acceptability is violated and the corresponding slice is not acceptable. However, in non-homogenous slopes, the soil parameters may be constant in a single slice even the soil underneath the slope is non-homogenous. Also the maximum value of m in a single slice may be equal to or smaller than unity even with the variation of the soil parameters in that slice. Therefore, the assumption that m is the local extremum in the two boundaries is a good starting point in non-homogenous slopes to obtain the initial approximation of the geometry but this assumption needs further examination. The geometry of a standard slice is initially defined using Equations 4.54 and 4.59, providing that the weighted average soil parameters are used in the equations. After the geometry of the slice is obtained, the variation of m needs to be inspected and, therefore, the value of m for an arbitrary plane within the slice is required.

6.2.1 m in a standard slice in non-homogenous slopes


The definition of m for an arbitrary plane both in the slip surface side and in the inter-slice boundary side has been given in Chapter 4. For ease of presentation, some of the equations are repeated in the following. The reason for repeating them here is that , c, EL & XL need to be explained. As defined in Chapter 4, mS, the reciprocal of the factor of safety for an arbitrary plane BA' (Figure 4.5) in the slip surface side, is defined as:

S m =

(N

S U ) tan S + c L

(4.25)

where S, cS are the weighted average soil parameters on plane BA' and:

N = N i cos + Ti sin WS sec cos( i + ) + PW L ) sin ( i + i ) + X L cos( i + i ) (E L

(4.26)

133

T = N i sin + Ti cos WS sec sin ( i + ) + + PW L ) cos( i + i ) + X L sin ( i + i ) (E L

(4.27)

It has been shown that after the geometry of the slice is assumed, the normal forces and shear forces, N'i, Ti, E'i+1 and Xi+1, can be obtained from Equations 4.20 through 4.23 and forces due to water pressures can be obtained from the analysis shown in Section 4.4.6. The normal force EL and shear forces XL on the small part of slice boundary AA' (Figure 4.5) depends on the effective normal stress distribution assumption as described in Section 4.3.4 and the weighted average soil parameters along plane AA'. Therefore, the main difference in the computation of m for homogenous slopes and non-homogenous slopes is that the soil parameters are constant in homogenous slopes, while the weighted average soil parameters have to be used if more than one soil layer exists in a single slice. Similarly, the reciprocal of the factor of safety of the plane, mB, of an arbitrary plane BC' (Figure 4.5) in the inter-slice side is defined as:

mB =

(E

B PW ) tan B + c D

(4.30)

where B, cB are weighted average soil parameters on the plane BC' and:

E = Ei +1 cos + X i +1 sin WB sec sin ( i +1 ) X = Ei +1 sin + X i +1 cos + WB sec cos( i +1 )

(4.31)

(4.32)

Again, the weighted average soil parameters are needed to compute the value of mB in the interslice boundary side.

Computation of the weighted average soil parameters


It is shown in Equations 4.20 through 4.23 that to compute the effective normal forces and shear forces, N'i, Ti, E'i+1 and Xi+1, acting on the planes AB and BC (Figure 4.5), the weighted average soil parameters are needed on the slip surface and the inter-slice boundary. To simplify the

134

problem, the slip surface of a single slice is constrained to stay in the same soil layer by adjusting

x, which is detailed in Section 6.3.1. Therefore, no weighted average computation is needed for
the slip surface. However, the inter-slice plane and the internal planes in a single slice (defined in Chapter 4) will pass through several soil layers in non-homogenous slopes and therefore, the weighted average soil parameters have to be used. Let us assume that the ith inter-slice plane passes through several soil layers and the effective normal force and shear force acting on that plane are Ei and Xi, respectively. If the plane is divided into n segments by soil layer boundaries and if the effective normal forces and shear force on the jth segment are denoted Eij and Xij, then:

X i = Ei tan i + cid i

= Ei1 tan i1 + Ei 2 tan i 2 + ...Ei j tan i j + ... + Ei n tan i n +


1 i di 1

( ( c

+ ci

d i2

+ ... + ci d i + ... + ci
j j

d in

(6.1)

where cij and ij are soil parameters in the jth soil layer, dij is the length of the jth segment of the ith inter-slice plane andi andci are weighted average values. Therefore, the weighted average cohesionci can be obtained as:

ci =

ci1 d i1 + ci 2 d i2 + ... + ci j d i j + ... + ci n d in di

(6.2)

And the weighted average angle of shearing resistancei can be obtained as:

i = arctan

Ei1 tan i1 + Ei 2 tan i 2 + ...Ei j tan i j + ... + Ei n tan i n Ei

(6.3)

To know the effective normal force on the jth segment, Eij, it is necessary to know the effective normal stress distribution along the ith inter-slice plane. As already discussed in Chapter 4, the linear stress distribution can significantly simplify the computation and it is adopted in the current analysis. For linear stress distribution, if the effective normal stress acting at the end point of the

135

plane and the effective normal force acting on that plane are known, the actual stress distribution is defined and the value of Eij can be obtained.

The application of the weighted average soil parameters


Let us consider a general slice as shown in Figure 4.5. Assume that all the forces and the stress distribution on plane AD are known. Therefore,ci andi can be obtained using Equations 6.2 and 6.3. To obtain the geometry of the new slice ABCD, initial approximations of the two angles

i and i+1 are required as discussed in Chapter 5. As shown in Equations 4.20 through 4.23, to
obtain N'i, Ti, E'i+1 and Xi+1, it is necessary to know the weighted average soil parameters of the inter-slice boundaryci+1 andi+1 (the slip surface is constrained to stay in the same soil layer and no average computation is needed). However, the values ofci+1 andi+1 are dependent on the effective normal force on the plane BC, E'i+1 and the effective normal stress at the point B on the plane BC, i+1, which are unknown. To solve the above problem, an iterative procedure has to be adopted. Initially, the values ofci+1 andi+1 are assumed to be equal to the values ofci andi. Then the values of N'i, Ti, E'i+1 and

Xi+1 can be computed from Equations 4.20 through 4.23. The effective normal stress at the point B on the plane BC, i+1, can be computed using Equation 4.36 and therefore, the new values of

ci+1 and i+1 can be updated from Equations 6.2 and 6.3. The above steps are repeated until the
difference of two successive values of ci+1 and i+1 are smaller than a certain criterion (2% is taken in this thesis). Practical examples show that the values of N'i, Ti, E'i+1 and Xi+1 are not very sensitive to the soil parameters on the inter-slice plane, therefore, the above procedure normally terminates within a few iterations. To obtain the weighted average soil parameters on the internal planes, the effective normal stress distribution along that plane is needed as shown above. For example, to obtain S and cS in equation 4.25, the effective normal stress at point B on the plane BA (Figure 4.5) and the effective normal stress at point A on the plane BA are required. If the effective normal stress at point B on the plane BC is known, the effective normal stress at point B on the plane BA can be obtained from the Mohr circle but for simplicity, these two values are assumed to be equal in this

136

thesis. The same assumption is applied to the effective normal stress at point A on the plane BA. Practice shows that the computed weighted average soil parameters on the plane BA are not sensitive to this minor difference.

6.2.2 Validation of stress acceptability and further constraints in non-homogenous slopes


Validation of stress acceptability in non-homogenous slopes
As discussed in Chapter 4, for a standard slice, stress acceptability requires that the factor of safety for any plane within the slice must be greater than or equal to unity. The assumption that

m is the local extremum in the two boundaries is valid in homogenous slopes since the soil
parameters are constant within the slice. However, in non-homogenous slopes, if the slice is initially defined by Equations 4.54 and 4.59, the variation of m has to be inspected. As shown in Section 5.3.2, m keeps decreasing when the internal plane moves from plane BA to

BD if and c are constant in homogenous slopes. If an arbitrary plane is considered in nonhomogenous slopes, for example, an arbitrary plane in slip surface side as shown in Equation 4.25: if S and cS are equal to or larger than i and ci, which are the soil parameters on the slip surface, the value of mS on that plane will be equal to or smaller than the corresponding plane in the same position in homogenous slopes. On the other hand, if S and cS are smaller than i and

ci, the value of mS on that plane will be greater than on the corresponding plane in the same
position in homogenous slopes, in some cases even larger than unity. If the value of mS exceeds unity, it does not satisfy stress acceptability. Therefore, the slice which is defined using the assumption of local extremum in the two boundaries cannot guarantee the stress acceptability in non-homogenous slopes, as shown in the following examples. Example 6.1 and 6.2 Let us consider two simple example slopes as shown in Figures 6.1 and 6.2. All the soil parameters and initial conditions before the unsolved slice (shown as the dashed line) are listed in the Table 6.1 and 6.2. In Example 6.1, the slip surface of the unknown slice will go into a weaker soil layer while in Example 6.2, the slip surface of the unknown slice will develop in a stronger soil layer. Equations 4.54 and 4.59 are used to obtain an initial approximation of the unknown

137

slice. The computed geometry of the slice and the corresponding forces acting on that slice are listed in Table 6.1 and 6.2. Note that the difference of the two solved slices comes from the different weighted average soil parameters used in Equations 4.54 and 4.59. After the slice is obtained using Equations 4.54 and 4.59, let us check the variation of m both in the slip surface side and in the inter-slice boundary side. The variations of m as the internal plane varies from the slip surface to the inter-slice boundary for Examples 6.1 and 6.2 are shown in Figures 6.3 and 6.4, respectively. Note that the negative value of m is due to the sign convention. It is shown in Figures 6.3 and 6.4 that m is a local extremum when the internal plane is either the slip surface or the inter-slice boundary. However, Figure 6.4 shows that m exceeds unity even if

m has a local extremum when is equal to zero.


In practice, most of the slices defined by the assumption of local extremum in the two boundaries satisfy the stress acceptability in non-homogenous slopes, as shown in the Example 6.1 and Figure 6.3. It is a good starting point to obtain the initial approximation of the unknown slice. However, it is necessary to inspect the variation of m of the internal planes after the slice is obtained. If the stress acceptability is violated as shown in Example 6.2 and Figure 6.4, further constraints have to be applied to that slice.

Figure 6.1 A general slice in two soil layers (a stronger soil layer overlaying a weaker soil layer)

138

Table 6.1 (a) Soil properties (b) Initial conditions and details of solved slice in Example 6.2 (a)

(kN/m2)
First soil layer Second soil layer 20 20 (b)

c (kPa) 16 8

(in degrees)
25 15

Initial conditions x
2.5989

y
-0.5

xs
3.8284

ys
1.2761

-0.1783

0.6188

W
20.656

N
27.839

T
22.647

E
66.867

X
65.743

The slice after solving Equations 4.54 and 4.59 x*


3.4323
*

y
-0.7350

xs**
4.8284

ys
1.6095

-0.2748

0.5371

W
36.483

N
50.166

T
20.369

E
94.028

X
81.911

**

x and y are coordinate of the position of the point B (Figure 4.5) xs and ys are coordinate of the position of the point C (Figure 4.5)

Figure 6.2 A general slice in two soil layers (a weaker soil layer overlaying a stronger soil layer)

139

Table 6.2 (a) Soil properties (b) Initial conditions and details of solved slice in Example 6.3 (a) (kN/m2) First soil layer Second soil layer 19 19 (b) Initial conditions x
-0.55831

c (kPa) 10 20

(in degrees) 25 30

y
-0.8

xs
0.3488

ys
0.1163

-0.4246

0.7804

W
5.6154

N
13.167

T
10.071

E
30.725

X
27.221

The slice after solving equations 4.54 and 4.59 x


-0.0094323

y
-1.0455

xs
1.3488

ys
0.4496

-0.2744

0.7264

W
16.796

N
35.109

T
32.296

E
59.553

X
53.819

Figure 6.3 The variation of m in the slice in Example 6.1

140

Figure 6.4 The variation of m in the slice in Example 6.2 (derivative may not be zero due to neglection of differential form of soil parameters in Equation 4.44)

Further constraints used in non-homogenous slopes


The variation of m is not the main interest of the new procedure. As long as no value of m of any plane within the slice is larger than unity, the stress acceptability is satisfied. Therefore, the maximum value of m is the main interest of the new procedure. After the slice is initially defined by Equations 4.54 and 4.59, the maximum values of m on both sides are evaluated. If the maximum value of m is larger than unity, in other words, stress acceptability is not satisfied in either side or both sides, a set of new equations has to replace Equation 4.54 or 4.59. These are defined as:

max mS = 1
or:

( )

(6.4)

max mB = 1

( )

(6.5)

141

The algorithm for solving a standard slice in non-homogenous slopes is shown in Figure 6.5. In each step, two non-linear equations need to be solved. The Trust-Region Dogleg Method (Powell, 1970), which is used for solving non-linear equations for homogenous slopes, is also used to solve non-linear equations for non-homogenous slopes.

6.2.3 Method to find the maximum value of m


As discussed above, the success of finding the maximum value of m is essential for the application of stress acceptability to non-homogenous slopes. The details are discussed in the following section. The maximum value of m can be found by equally dividing both the slip surface side and the inter-slice boundary side into a number of small segments, denoted by . The value of m is evaluated in each plane and the maximum value in the corresponding side is picked. Obviously, by using a small value of it is possible to miss the true maximum, which may cause the true

i and i+1 to be missed also. On the other hand, using a large value of is computational
demanding. For the purpose of illustration, Example 6.2 is considered again. As stated above, the new slice is first solved using Equations 4.54 and 4.59. By adopting as 20, 50 and 100, the maximum values of m are listed in Table 6.3 after the slice is defined. It is seen that the values of mS in each case are larger than unity. According to the algorithm for a standard slice in non-homogenous slopes as shown in Figure 6.5, Equations 6.4 and 4.59 are adopted to resolve the slice. The values of i and i+1 obtained from Equations 4.54 and 4.59 are used as initial approximations. i and i+1, the weighted average soil parameters on the inter-slice boundary and the maximum values of m in each iteration are listed in Table 6.3 for different values of . The total number of calculations required to evaluate m is also listed in Table 6.3 for different values of . It is noticed that there is only a very small difference between the results from equal to 50 and those from equal to 100. Based on extensive case studies, the value of can adopted as 50-60, which provides enough precision and reduces the computation demand.

142

START

Input x

dmS / d = 0 dmB / d = 0

solve i and i+1

Yes

max mS > 1
No

( )

max mS = 1 dmB / d = 0

( )

solve i and i+1

max mB > 1
No

( )

Yes

max mB > 1
Yes

( )

No

dmS / d = 0
max mB = 1

( )

solve i and i+1

( ) max (m ) = 1
max mS = 1

No

max mS > 1
Yes

( )

solve i and i+1

max mS = 1 max mB = 1

( ) ( )

output i and i+1

solve i and i+1

END

Figure 6.5 The algorithm of solving a single slice within non-homogenous slopes

143

Table 6.3 Geometry of the slice, weighted average soil parameters and maximum value of m in each iteration when is taken as: (a) 20, (b) 50 and (c) 100 (a)

=20
Iteration 0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 x* 0.0791 0.0416 0.0695 0.0594 0.0487 0.0379 0.0267 0.0153 0.0237 0.0166 0.0219 0.0175 0.0129 0.0084 0.0048 y* -0.9795 -1.0170 -0.9891 -0.9984 -1.0072 -1.0156 -1.0236 -1.0312 -1.0255 -1.0302 -1.0266 -1.0296 -1.0325 -1.0353 -1.0375

i
-0.2744 -0.3470 -0.2926 -0.3107 -0.3289 -0.3470 -0.3651 -0.3831 -0.3696 -0.3809 -0.3724 -0.3795 -0.3865 -0.3936 -0.3991

i+1
0.7264 0.7280 0.7268 0.7276 0.7286 0.7299 0.7314 0.7331 0.7319 0.7329 0.7322 0.7328 0.7335 0.7343 0.7349

i+1
0.4531 0.4614 0.4550 0.4571 0.4592 0.4613 0.4635 0.4658 0.4641 0.4656 0.4645 0.4654 0.4663 0.4672 0.4679

ci+1
11.2559 11.4795 11.3146 11.3699 11.4220 11.4714 11.5178 11.5614 11.5285 11.5557 11.5353 11.5522 11.5687 11.5847 11.5970

mS 1.0812 1.0231 1.0618 1.0438 1.0326 1.0233 1.0146 1.0066 1.0126 1.0076 1.0113 1.0082 1.0052 1.0023 1.0000

mB 1.000 1.000 1.000 1.000 1.000 1.000 1.000 1.000 1.000 1.000 1.000 1.000 1.000 1.000 1.000

Total number of evaluation of m: 560

144

(b)

=50
Iteration 0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 x (slip surface) 0.0791 0.0416 0.0695 0.0594 0.0487 0.0379 0.0267 0.0153 0.0237 0.0166 0.0219 0.0175 0.0129 0.0084 0.0038 -0.0008 -0.0055 -0.0094 y (slip surface) -0.9795 -1.0170 -0.9891 -0.9984 -1.0072 -1.0156 -1.0236 -1.0312 -1.0255 -1.0302 -1.0266 -1.0296 -1.0325 -1.0353 -1.0381 -1.0408 -1.0434 -1.0455

i
-0.2744 -0.3470 -0.2926 -0.3107 -0.3289 -0.3470 -0.3651 -0.3831 -0.3696 -0.3809 -0.3724 -0.3795 -0.3865 -0.3936 -0.4006 -0.4077 -0.4147 -0.4206

i+1
0.7264 0.7280 0.7268 0.7276 0.7286 0.7299 0.7314 0.7331 0.7319 0.7329 0.7322 0.7328 0.7335 0.7343 0.7351 0.7359 0.7367 0.7375

i+1
0.4531 0.4614 0.4550 0.4571 0.4592 0.4613 0.4635 0.4658 0.4641 0.4656 0.4645 0.4654 0.4663 0.4672 0.4681 0.4690 0.4699 0.4707

ci+1
11.2559 11.4795 11.3146 11.3699 11.4220 11.4714 11.5178 11.5614 11.5285 11.5557 11.5353 11.5522 11.5687 11.5847 11.6003 11.6154 11.6302 11.6422

mS 1.0812 1.0233 1.0632 1.0476 1.0342 1.0234 1.0145 1.0079 1.0128 1.0087 1.0118 1.0092 1.0067 1.0047 1.0034 1.0022 1.0010 1.0000

mB 1.000 1.000 1.000 1.000 1.000 1.000 1.000 1.000 1.000 1.000 1.000 1.000 1.000 1.000 1.000 1.000 1.000 1.000

Total number of evaluation of m: 1800

145

(c)

=100
Iteration 0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 x (slip surface) 0.0791 0.0392 0.0689 0.0580 0.0466 0.0349 0.0230 0.0317 0.0243 0.0299 0.0252 0.0204 0.0157 0.0108 0.0060 0.0010 -0.0039 -0.0089 -0.0140 -0.0102 y (slip surface) -0.9795 -1.0194 -0.9898 -0.9996 -1.0089 -1.0178 -1.0262 -1.0198 -1.0251 -1.0212 -1.0245 -1.0276 -1.0308 -1.0338 -1.0368 -1.0397 -1.0425 -1.0453 -1.0479 -1.0459

i
-0.2744 -0.3519 -0.2938 -0.3132 -0.3325 -0.3518 -0.3711 -0.3567 -0.3687 -0.3597 -0.3672 -0.3747 -0.3823 -0.3898 -0.3973 -0.4049 -0.4124 -0.4199 -0.4274 -0.4218

i+1
0.7264 0.7281 0.7268 0.7277 0.7288 0.7302 0.7319 0.7308 0.7318 0.7310 0.7317 0.7324 0.7331 0.7339 0.7347 0.7356 0.7364 0.7374 0.7383 0.7376

i+1
0.4531 0.4620 0.4552 0.4574 0.4596 0.4619 0.4643 0.4625 0.4640 0.4629 0.4638 0.4648 0.4657 0.4667 0.4677 0.4686 0.4696 0.4706 0.4716 0.4708

ci+1
11.2559 11.4935 11.3185 11.3772 11.4323 11.4842 11.5328 11.4961 11.5266 11.5038 11.5227 11.5410 11.5589 11.5762 11.5931 11.6095 11.6254 11.6407 11.6557 11.6444

mS 1.0812 1.0207 1.0621 1.0457 1.0320 1.0209 1.0122 1.0186 1.0131 1.0171 1.0137 1.0109 1.0082 1.0061 1.0041 1.0027 1.0014 1.0007 1.0000 1.0005

mB 1.000 1.000 1.000 1.000 1.000 1.000 1.000 1.000 1.000 1.000 1.000 1.000 1.000 1.000 1.000 1.000 1.000 1.000 1.000 1.000

Total number of evaluation of m: 4000


*

Note that the values of x and y are the coordinates of the end point of the slip surface of the unknown slice

146

6.3 THE INCREMENT OF x IN NON-HOMOGENOUS SLOPES


As shown in Chapter 4, the increment x along with the two inclination angles i and i+1 defines the geometry of a standard slice. The value of x is usually assumed. The effects of x on a single slice and on the whole slope in homogenous slopes are discussed in Section 5.4. In homogenous slopes, the increment x is constant except for the last slice, as shown in Section 4.6. However, in non-homogenous slopes, this is generally not the case. There are two reasons for a variable x to be used in non-homogenous slopes: first, as discussed in Section 6.2.1.2, the slip surface of a single slice is restricted to stay within the same soil layer to simplify the computation of the weighted average soil parameters; second, for satisfaction of kinematical acceptability, the value of x may need to be increased systematically to make i larger than i-1.

6.3.1 Variable x used to make the slip surface of a single slice stay within the same soil layer
The application of this technique is illustrated in the following example. Example 6.3 The geometrical properties of the slope and the position of the soil layer boundary are shown in Figure 6.6, and the soil properties are listed in Table 6.4. Assume that the end point of the slip surface of the previous slice is close to the soil layer boundary (Figure 6.6). The geometrical properties of the previous slice are listed in Table 6.5 as initial conditions. If a constant x is used, the slip surface of the new slice will enter into the second soil layers as shown by the dashed line in Figure 6.6, which makes the problem more complicated (the weighted average soil parameters need to be computed for the slip surface). To avoid this, a smaller x is used to make the slip surface of the new slice stay within the first soil layer and the end point of the slip surface touch the soil layer boundary (solid line in Figure 6.6). The slip surface of following slices will start from the beginning of the second soil layer. The procedure is presented below. The results of the example are illustrated in Table 6.5. A trial step, which uses the two inclination angles of the previous slice and constant x, is tested first. If the end point of the slip surface of the new slice stays within the same soil layer as the start point of the surface, the constant x is

147

adopted in the algorithm shown in Figure 6.5 for solving the new slice. If not, by using i and i+1 and the soil layer boundary, a smaller x can be defined (step 1-1). By placing the re-defined x into the algorithm shown in Figure 6.5, the geometry of the slice can be updated and new i and

i+1 can be obtained (step 1-2). Again, by using the new i and i+1 and the soil layer boundary, a
new x can be defined (step 2-1). An iterative procedure can be set up until the difference between two successive values of x is smaller than a preset criterion.

Figure 6.6 Variable x to make the slip surface of a single slice stay within the same soil layer

Table 6.4 Soil properties in Example 6.3

(kN/m2)
First soil layer Second soil layer 20 20

c (kPa) 10 20

(in degrees)
10 30

148

Table 6.5 Geometry of the new slice assuming different x

Initial condition Step test 1-1 1-2 2-1 2-2 3-1 -0.3982 -0.3982 -0.3721 -0.3721 -0.3752 -0.3752 -0.3982

0.58801

x
1

x (slip surface) 1.3312

y (slip surface) -0.6031

Difference between x

Criterion for x

Satisfied?

0.5880 0.5880 0.5615 0.5615 0.5621 0.5621

1 0.6914 0.6914 0.6756 0.6756 0.6716

2.0081 1.7992 1.8468 1.8357 1.8339 1.8310

-0.8879 -0.8000 -0.8043 -0.8000 -0.8011 -0.8000

0.3086

0.005

No

0.0158

0.005

No

0.004

0.005

Yes

149

6.3.2 Variable x to satisfy kinematical acceptability


As discussed in Chapter 2, to satisfy kinematical acceptability, the inclination angle of the slip surface of the current slice, i, has to be larger than or equal to the inclination angle of the slip surface of the previous slice, i-1. In non-homogenous slopes, especially when the slip surface develops from one soil layer to another, the kinematical acceptability may not be satisfied using a constant x (as shown in Figure 6.1). If so, the value of x is systematically increased until kinematical acceptability is satisfied, since the value of i is dependent on x, as shown in Section 5.4.

6.4 ACCEPTED SLIP SURFACE IN NON-HOMOGENOUS SLOPES


For a given slope to be analysed in non-homogenous slopes, by picking a starting point on the slope surface, assuming a critical acceleration kcg and specifying a slip surface path (discussed later in Section 6.5), we can determine a slip surface slice by slice from toe end to crest end. For each slice to be solved, the increment x is adopted as the procedure shown in Section 6.3 and then the algorithm to solve that slice is shown in Figure 6.5. As discussed in Chapter 5, any desired critical acceleration kcg can produce a slip surface, but this may not be acceptable on the basis of kinematics. However, in non-homogenous slopes, for a given starting point and a specified slip surface path, there is only one acceptable slip surface and corresponding acceptable critical acceleration. By changing the assumed critical acceleration the generated slip surface can be made to end close to a critical point (in general, the critical point is the corner point of the crest but it can be any designated point). Then the last two slices can be solved together as shown in Chapter 4. The geometry of the last two slices along with kgend can be solved to make the last effective normal force equal to zero based on the forces and geometrical information from the previous slices. The

kgend may be different from the starting kcg. Therefore, an iteration algorithm is used to solve this
problem until the starting kcg and kgend become equal (the procedure is the same as the one used for homogenous slopes). Then the slip surface is the acceptable slip surface for the chosen starting point and the specified slip surface path.

150

6.5 SLIP SURFACE PATH IN NON-HOMOGENOUS SLOPES


When the slip surface encounters soil layer boundaries, then there are some uncertainties as to the path that the slip surface will follow. To analyse this phenomena, an idea of slip surface path is proposed in this thesis. The concept of slip surface path is illustrated in the following example. Example 6.4 It is shown in Figure 6.7 that if the starting point is set at point A and a critical acceleration is assumed, the slip surface encounters the first soil layer at point B (or B' or, B'', if different kcg values are used but these points are very close to point B) using a constant increment x. After that, a new slice needs to be solved. The slip surface of the new slice may go back into the same soil layer, developing along the soil layer boundary with the strength associated with the layer above or enter into the new soil layer. If x is kept constant, the new slip surface point will follow

BC. However, experience shows that, depending on the soil properties, the slip surface may
follow the soil layer boundary. Under such situations, i, which is the inclination angle of the first soil layer boundary, is known. Then i+1 and x, instead of i and i+1, are the two unknowns. In such a case, if a varying x is solved, the new slip surface B'D will be found by using the algorithm shown in Figure 6.5. Whenever the slip surface encounters a soil layer boundary, there are two possibilities: one from a constant x and the other from a varying x. The rest of the slices should be solved following both possibilities. A robust procedure is that every possibility is considered and the one with minimum acceptable kcg selected.

Table 6.6 Soil properties in Example 6.4

(kN/m2)
First soil layer Second soil layer Third soil layer 20 20 20

c (kPa) 5 5 5

(in degrees)
20 10 30

151

Figure 6.7 Slip surface path. Note that different surfaces encounter the first soil layer boundary at three different points near B but the differences are too small to be seen

Therefore, as shown in Figure 6.7, three slip surface paths are designed. The first one is defined as when the slip surface encounters the first soil layer boundary; it will develop along the boundary and then carry on. The second one is defined as that when the slip surface crosses the first soil layer boundary but develops along the second soil layer boundary. The third one is defined as that when the slip surface crosses all the soil layer boundaries. The acceptable slip surfaces should be found for each slip surface path using the iterative procedure described before in Section 6.4. In this example, three acceptable slip surface paths have been defined and the second with kcg equal to 0.095 is the accepted slip surface path with minimum kcg for the corresponding starting point A in non-homogenous slopes.

152

6.6 METHODOLOGY TO DETERMINE THE CRITICAL SLIP SURFACE


The methodology to determine a critical slip surface, both in homogenous and non-homogenous slopes, is summarised in the following section. (1) Select a possible slip surface from experience and find the associated critical acceleration for that surface using an established stability analysis method, e.g. Sarma (1973) or Sarma (1979). Note that the initially assumed slip surface may not be acceptable in terms of stress and kinematical acceptability. (2) A starting point is picked from the slope surface and the critical acceleration found above is used as an initial approximation of the critical acceleration. (3) A slip surface path is designed as shown in Section 6.5. (4) A slip surface is obtained slice by slice. The procedure to select x is shown in Section 6.3 and then each slice is solved using the algorithm shown in Figure 6.5. (5) The acceptable slip surface and the associated critical acceleration are found using the iterative procedure discussed in Section 6.4 for each slip surface path. (6) Comparing the acceptable critical accelerations obtained from different slip surface paths the slip surface path with the minimum acceptable critical acceleration is selected. (7) Repeat steps (2) to (6) for different starting points. Note that from the second loop, the critical acceleration obtained from the previous loop can be used as the initial approximation of the critical acceleration. By repeating steps (2) to (6), every starting point on the slope surface can be analysed and then the acceptable slip surface and the associated critical acceleration can be determined for each point. As stated in Chapter 5, the critical slip surface can be found by defining an area where the starting point may exist. The area is equally divided into a number of segments and each point is selected as a starting point. At each starting point, a minimum critical acceleration with acceptable slip surface is found for either homogenous or non-homogenous slopes. The slip surface with the global minimum critical acceleration is the critical slip surface. Another possible option to obtain the critical slip surface is that the starting point can be generated randomly within the possible area for starting points. Providing that the number of random points is large enough, the minimum value of the critical accelerations is close to the

153

statistical global minimum. In this thesis, the option of randomly generated starting points is not considered. It is possible topic of further research.

6.7 PRACTICAL APPLICATION


In the following section, some examples are taken from the literature to validate the new procedure. Example 6.6 The following two example slopes are selected from Zolfaghari et al. (2005). The geometries of the slopes are shown in Figure 6.8 a and b. The soil properties are the same in the two examples and are listed in Table 6.7. The slip surfaces found by Zolfaghari et al. using a genetic algorithm are shown as dashed lines in Figures 6.8 a and b and their computed factors of safety are shown in the same figures. Note that the coordinates of the critical slip surfaces by the authors are digitized from the figures of their paper and then the critical accelerations of those surfaces are recomputed here using Sarma method (1973) for the purpose of comparison.

Table 6.7 Soil properties in Example 6.6

(kN/m2)
First soil layer Second soil layer Third soil layer Fourth soil layer 19 19 19 19

c (kPa) 15 17 5 35

(in degrees)
20 21 10 28

The slip surfaces found by the new procedure are shown in the same figures as solid lines and after the slip surfaces are found, Sarmas program (2000) [the program uses Sarma method (1973)] is used to compute the factors of safety for the purpose of comparison. The current procedure finds smaller factors of safety with different critical slip surfaces. Note that a tension crack is

154

used for the new procedure even though the use of the tension crack is not absolutely necessary as far as the method is concerned.

Figure 6.8 (a)

(b) Figure 6.8 (a) and (b) Comparison of slip surfaces of four-layer slope given by the new procedure and that obtained by Zolfaghari et al. (2005) using a genetic algorithm approach

155

These examples show that when applying optimization methods to slope stability analysis, the relationship between the function to express the slip surface and the optimization method needs be handled carefully. The more free parameters used to express the function, the greater the likelihood that they express the slip surface in complex conditions. However, too many parameters can cause numerical problems when an optimization technique is used. On the other hand, if the used function cannot express the complexity in non-homogenous slopes and if the optimization method lacks robustness, it will fail to identify the critical slip surface. However, in the present study, by changing the increment x, any number of slices can be used to analyse the proposed slope and therefore, it has great advantages when applied to non-homogenous slopes with complex geometry. Example 6.7 This example is taken from Yamagami and Ueta (1988b), where a slope in layered soil is analysed using different calculus-based methods of minimization. A Monte Carlo technique is also used by Greco (1996) to find the minimum factor of safety for the same problem. The geometrical features of the analysed slope and slip surfaces found by the above methods are shown in Figure 6.9. The soil properties of different soil layers are listed in Table 6.8.
Table 6.8 Soil properties in Example 6.7

(kN/m2)
First soil layer Second soil layer Third soil layer Fourth soil layer 20.38 17.64 20.38 17.64

c (kPa) 49 0 7.84 0

(in degrees)
29 30 20 30

The coordinates of the critical slip surface from Yamagami and Ueta (1988b) are digitized from the figure reported in the their paper. The coordinates of the critical slip surface of Greco (1996) are provided by the author. The associated factors of safety are given by the authors. For the purpose of comparison, the critical accelerations are computed using Sarma method (1973). After the critical acceleration is obtained, the effective normal forces and shear forces acting on the vertical planes are available and then the local factors of safety on those planes can be evaluated as:

156

mv =

(N v U v ) tan v
Tv

+ cv Lv

(6.6)

where Nv and Tv are the total normal and shear forces acting on the vertical plane and Uv is the force due to water pressure on that plane. Thev andcv are weighted average soil properties on the vertical plane and Lv is the length of the plane. The local factors of safety are shown in Figures 6.10 b and c. It is seen from the figures that the results are not acceptable because factors of safety in some vertical planes are less than unity and therefore, the stress states inside those slices exceeds the limited strength of the soil. The critical slip surface and the associated critical acceleration for that slope can also be found by the new procedure and are shown as the solid line in Figure 6.9. The factor of safety for that critical slip surface can be computed by any available stability methods, e.g. Sarma (1973), and the factors of safety on the vertical planes can be evaluated using Equation 6.6 as shown in the Figure 6.10a. The factors of safety (or critical acceleration) are similar using the different methods but the slip surfaces are significantly different (this example will be further examined in Chapter 7 by using the finite element method). Figure 6.10a shows that only the current procedure gives inter-slice factors of safety that satisfy stress acceptability. This example shows that the methods based on optimization techniques may obtain smaller factors of safety, but their results may not be acceptable.

157

Figure 6.9 Comparison of slip surfaces of a four-layer slope estimated using different methods

158

Figure 6.10 (a)

Figure 6.10 (b)

(c) Figure 6.10 Internal factors of safety on vertical planes within surfaces obtained: (a) by the new procedure; (b) by Yamagami & Ueta (1988b); and (c) by Greco (1996)

159

7. COMPARISON OF THE NEW PROCEDURE WITH FINITE ELEMENT METHOD FOR DETERMINING THE CRITICAL SLIP SURFACE

7.1 SCENARIO
The development of finite element method and its application to geotechnical problems has provided geotechnical engineers with a powerful analysis tool. In this chapter, finite element analysis is used as an independent tool to validate the new procedure in both homogenous and non-homogenous slopes. The finite element method was introduced to slope analysis by Clough, Richard & Woodward (1967) to analyse embankment stresses and deformations. Since then finite element method has been the most popular method for analysing stresses and deformations of slopes (Duncan, 1996). The application of the finite element method to slope stability analysis in the literature is briefly reviewed in Section 7.2. In Section 7.3, the procedure to determine the critical acceleration and the failure mechanism of slopes by finite element analysis is presented, including the description of the model, the loading sequence and the definition of the failure mechanism. Finally, a comparison of the new procedure with finite element analysis is carried out, in terms of the computed critical acceleration, the critical slip surface and the stress distribution along the slip surface both in homogenous and non-homogenous slopes.

7.2 INTERPRETATION OF RESULTS FROM FINITE ELEMENT SLOPE STABILITY ANALYSIS


Finite element analysis of slopes is a complex problem. It requires the definitions of initial conditions, boundary conditions, stress-strain properties of the soil underneath the slope and the

160

loading sequences. Finite element analysis gives the stresses and deformations in a confined space and shows the region of high strain, which can then be interpreted to deduce a slip surface. The choice of the model to simulate the stress-strain behaviour of fill or foundation materials of the slope and the elaborate modelling of formation of a slope either by excavation or embankment is beyond the scope this thesis. The interest of this thesis is to interpret the results from finite element analysis of slopes in terms of stability.

7.2.1 Interpretation of the results from finite element analysis in terms of factor of safety
The finite element analysis of stresses and deformations of slopes does not give direct information about the factor of safety, which is a main concern when geotechnical engineers design new embankments or evaluate existing natural or man-made slopes. Within the finite element framework, there are two main sets of procedures existing in the literature to obtain such information: firstly, search for the critical slip surface in the stress field obtained from stresses and deformations analysis of slopes; or secondly, obtain the factor of safety directly from a separate finite element analysis using the strength reduction technique.

Search for the slip surface in the stress field


The first type of procedure is to search for the critical slip surface using stress fields obtained from a finite element analysis. The advantage of this procedure is that it can follow any finite element analysis of slopes. It can take into account any type of slope history (natural, excavation or embankment), any loading sequence, and accommodate appropriate stress-strain models for fill and foundation materials. a) Definition of the factor of safety in the stress field In this procedure, the factor of safety can be defined as the ratio between the overall available strength and the shear stress obtained by the stress analysis along a pre-defined slip surface. The factor of safety, F, along any pre-defined slip surface can be expressed as:

161

F=

d
(7.1)

d
i

where i is the shear stress at point i along the slip surface, f is the available shear strength at point i along that surface, and is the length of the slip surface. If a slip surface can be assumed, the normal and shear stress acting on any point of the slip surface, i and i, can be obtained as:

i =

x + y
2

x y
2

cos 2 xy sin 2

(7.2)

i =

x y
2

sin 2 + xy cos 2

(7.3)

where x, y and xy are the stress components at point i and is the angle made by the tangent of the slip surface with the horizontal at the ith point. The available shear strength at the ith point can be expressed as:

f = ci + i tan i
where ci and i are soil parameters at point i in terms of effective stress. b) The problems associated with the definition of the factor of safety in the stress field

(7.4)

There are two problems associated with the definition of the factor of safety in the stress field obtained by finite element method. The first is how to exactly evaluate stress components at any point along any plane in the stress field and the second is how to determine the critical slip surface. The stress components are available at gauss points after finite element analysis has been performed. The stress components at any point can be interpolated from the known stress components at gauss points around that point. It is found that this direct calculation of stress results in discontinuous stresses and, lower accuracy at the boundary of the elements. To overcome this deficiency, the global stress smoothing method (Hinton and Compell, 1974), the

162

superconvergent patch recovery method (Zienkiewicz and Zhu, 1992) and other similar methods have been proposed in the literature. However, in this thesis the stress components of any point are found from interpolation of the surrounding gauss points. The critical slip surface is not clearly defined after the stress and deformation analysis of slopes has been carried out. The engineers can draw the critical slip surface based on their experience from certain contours of strains. There is no routine procedure to draw the critical slip surface and most engineers rely on the search for the critical slip surface in the stress field. c) Literature review on the search for the critical slip surface in the stress field The first attempt to search for the critical slip surface in the stress field was made by Kulhawy, Seed and Duncan (1969). Several circles are drawn in the stress fields based on their experience and the factors of safety for each circle are evaluated using Equation 7.1. By comparing the factors of safety, the slip circle with minimum factor of safety is identified as the critical slip surface. In more recent studies, the critical slip surface is found by applying optimization techniques in the stress field to find the minimum factor of safety and the associated critical slip surface, which are similar to optimization methods adopted in the limit equilibrium method as discussed in Chapter 3. Yamagami and Ueta (1988b) adopt the Dynamic Programming Method (DPM), similar to the one used by Baker (1980) in searching for the critical slip surface within the limit equilibrium framework, to search for critical slip surfaces in the stress field obtained from finite element analysis. The application of DPM to slope stability analysis requires that the slip surface is assumed to be a chain of linear segments connecting two state points located in two successive stages (the details of DPM have been given in Chapter 3). The difference in applying DPM using the two techniques is that the normal and shear stresses come from either the limit equilibrium method or the finite element method. In their work, the global stress smoothing method (Hinton and Compell, 1974) is applied to obtain a continuous stress field at the corner points of each element and then they interpolate the stress components at each point within the element based on the stress components at the corner points of that element. Yamagami and Ueta (1988b) compare the results obtained from their approach with the results from the Bishop and Morgenstern-Price methods based on critical circular surfaces. They find large discrepancies for the critical slip surface as well as for the factor of safety and they believe that the differences come from the

163

initial stress in the foundation ground adopted in the finite element analysis, which the limit equilibrium method can not consider. Zhou, Williams and Xiong (1995) apply the improved dynamic programming method (IDPM) to search for the critical slip surface within the stress field. This is a modified version of the method used by Baker (1980) and Ymamgami and Ueta (1988b). The difference between IDPM and DPM is that in IDPM, the slip surface may cross or develop along the stage lines (defined in Figure 3.7) while in DPM, the slip surface can only cross the stage lines. The IDPM provides more varieties for expressing the shape of the slip surface. In their work, the stage line is assumed to be either vertical or perpendicular to the slope surface at certain distances but they state that the direction of the stage lines has little effect on the deduced critical slip surface. Results from this procedure show good agreements with those obtained from the limit equilibrium method in homogeneous slopes and in a test embankment. Kim and Lee (1997) apply the Broyden-Fletcher-Goldfarb-Shanno (BFGS) (Avriel, 1976) optimization method to search for the critical slip surface. The global stress smoothing method is used to obtain a continuous stress field and then the factor of safety is calculated for a defined slip surface by a stress integration scheme, which transforms a slip line from a two-dimensional global coordinate system to a one-dimensional local coordinate system. The optimization strategy starts with circular surfaces, and then the critical circular surface is used as the initial approximation of the slip surface in searching for the noncircular critical slip surface. The noncircular surface is initially expressed with a few points, which reduces numerical problems, and then points are added in the middle of previous points to make the slip surface smooth. Kim and Lee (1997) conclude that the proposed procedure shows good agreement with the limit equilibrium solutions for homogenous slopes but that the predicted failure height of an embankment by their procedure is underestimated when compared with the actual embankment failure. More recently, a DPM approach, similar to the one used by Yamagami and Ueta (1988b) but with further kinematical restrictions applied to the trial slip surface, is used by Pham and Fredlund (2003) to search for the critical slip surface. They find that the factors of safety consistently increase with an increase of Poissons ratio. The factors of safety found by the limit equilibrium method lie between the factors of safety found by finite element analysis with higher (0.48) and lower Poissons ratio (0.33). The critical slip surface tends to be more circular in shape when the

164

Poissons ratio increases from 0.33 to 0.48. According to Pham and Fredlund (2003), when Poissons ratio approaches 0.5, corresponding to zero volume change (or rigid body motions), the computed factors of safety based on finite element analysis are similar to those obtained from the limit equilibrium method. d) Comments on searching for the slip surface in the stress field The advantage of this procedure is that it can accommodate more complex stress-strain behaviour of soil and simulate more complex loading sequences, such as excavation and embankment construction. However, it is based on operative conditions instead of limit conditions. This method may find the mathematically minimum factor of safety on any shaped surface but the shape may not be representative of a true failure surface. This is mainly because the stresses are not at failure and therefore no real slip surface exists.

Determination of factor of safety using the strength reduction technique


The second type of procedure to determine the factor of safety within finite element framework is the so called strength reduction technique, which is first used by Zienkiewicz, Humpheson and Lewis (1975) and later named by Matsui and San (1992). In this procedure, finite element analysis (normally with Mohr-Coulomb failure criterion) is used to determine the factor of safety directly. The idea of the method is to reduce soil properties, which simultaneously reduces the cohesion and angle of shearing resistance as it does in limit equilibrium. The factor of safety is the value that makes the slope fail under gravity. The procedure is done by trial and error. Zienkiewicz, Humpheson and Lewis (1975) make the first attempt using this type of procedure. In their finite element analysis, an elastic perfectly plastic Mohr-Coulomb model is used and the associated and non-associated (angle of dilation equal to zero) flow rule is adopted. Based on the their work, in homogenous slopes, factors of safety obtained from finite element analysis with the associated flow rule at collapse are very close to those from standard slip circle analysis in the limit equilibrium method (although which method is used to compute the factor of safety for the standard slip circle analysis is not clear). Then a more complex three-layer slope is analysed, in which both the associated and non-associated flow rules are adopted, to compare with the results obtained by different methods within the limit equilibrium framework. According to Zienkiewicz, et al. (1975), the difference in the factors of safety by adopting different flow rules is negligible.

165

However, the difference between their results and results from limit equilibrium is remarkable: 33% lower factors of safety than the highest values by limit equilibrium. They also make the conclusion that provided the displacement constraints are not excessive, the limit loads given by the associated and non-associated flow rules are similar. Matsui and San (1992) name this type of procedure strength reduction technique. The hyperbolic stress-strain relationship of soil (Duncan and Chang, 1970) is adopted in their finite element analysis. In their work, total shear strain contour is used to define the critical slip surface, which is not straightforward and is highly dependent on the experience of the user. Also as noted by Matsui and San (1992), well defined failure shear strain zone do not exist in excavated slopes. Griffiths and Lane (1999) use the strength reduction technique in finite element analysis, adopting elastic perfectly plastic soils with Mohr-Coulomb failure criterion. Non-associated flow rule (zero angle of dilation) is assumed in their analysis because they believe that slope stability analysis is relatively unconfined, therefore, the choice of dilation angle is less important. Gravity is applied to the finite element model by a single increment and a non-convergent solution with an iteration ceiling of 1000, is used to define the failure of the slope. They state that the factor of safety is independent of the number of increments in applying the gravity load. This statement is rather surprising. The factor of safety is found by trial and error and for each trial factor of safety, a limit load problem is solved. However, in limit load analysis based on the finite element method, the number of increments plays a significant role in achieving the correct answer. This has been extensively validated by Potts and Zdravkovic (1999). As proposed by Zheng, Liu and Li (2005), the Poissons ratio should be reduced with the reduction of the strength. The reduced Poissons ratio does not significantly change the computed factor of safety, but it reduces the number of iteration to achieve a convergent solution. The factor of safety is obtained by interpreting results from finite element analyses. In terms of the computed factor of safety, almost all researchers report similar results to those obtained by the limit equilibrium method. However, in terms of the critical slip surface, few studies have yet compared the critical slip surfaces obtained using the limit equilibrium and finite element methods.

166

7.2.2 Interpretation of the results from finite element analysis in terms of critical acceleration
To determine the critical acceleration under pseudo-static conditions using finite element analysis is straightforward. The stresses induced by gravity or other loading sequences can be regarded as the initial stress conditions. After that, a horizontal acceleration is gradually applied to the model until the slope fails. Loukidis, Bandini and Salgado (2003) use this procedure to compare the critical acceleration obtained using limit analysis and finite element methods. They determine the lower and upper bounds of the critical acceleration based on finite-element-based limit analysis. A commercial finite element package, ABAQUS, is used for their finite element analysis. The size of the elements is found to have a substantial effect on the value of the collapse load and a very large number of elements, for example 5625, is required to analyse a homogeneous slope. In their analysis, the slope is formed by excavation on the level ground and the horizontal earth pressure coefficient, K0, is found to have no effect on the computed critical acceleration. The critical accelerations obtained by finite element analysis stay in the region determined by the upper and lower bounds from limit analysis. They also observe that if the adopted angle of dilation becomes smaller for a given angle of shearing resistance, the computed critical acceleration also becomes smaller. Loukidis, et. al (2003) use a very complex pattern, finite element discretion, to obtain their limit analysis solutions and thereby lose the simplicity of the traditional limit analysis method, thus losing its competition to finite element methods.

7.3 FINITE ELEMENT MODEL USED TO DETERMINATE THE CRITICAL ACCELERATION AND THE CRITICAL SLIP SURFACE
All the finite element analyses in this thesis are performed using Imperial College Finite Element Program (ICFEP) (Potts and Zdravkovic, 1999). An accelerated modified Newton-Raphson scheme, with sub-stepping stress point algorithm, is employed to solve the nonlinear finite element equations. Eight nodded isoparametric elements are used, with reduced (22) integration.

167

7.3.1 Description of the model


The analysed slope is assumed to be in plane strain conditions and drained analysis is performed (no water pressures are currently considered because the concept of constant pore water pressure ratio is difficult to consider in finite element analysis). The height of the slope is assumed to be constant at 10m. The slopes are inclined at three different gradients, 1:5, 1:3 and 1: 1.5. The model is extended to 40m underneath the toe, 150m on the left side and 150m on the right side of the toe (Figure 7.1). The boundary conditions are given as horizontal fixity on the left and right boundaries and full fixity at the base. Initial stresses are defined as the stresses before the application of horizontal acceleration, accounting for the effect of loading history on the analysed slope. In this thesis, only gravity is considered. The initial stresses can be established by applying gravity to the whole model or by excavating the slope from a level ground. The initial stresses cannot be taken into account by the limit equilibrium method and therefore, their effect on the computed critical acceleration by finite element methods is presented in Section 7.3.4. A horizontal acceleration is gradually applied to the whole model until the slope fails. The acceleration which makes the slope fail is the critical acceleration. Obviously, the definition of failure is the critical point for accurate evaluation of the critical acceleration.

Figure 7.1 Model used for homogenous slopes

168

7.3.2 Definition of failure


After the initial stress field is obtained, a horizontal acceleration is gradually applied to the model. A relatively large increment, 0.001g, is used first and then the increment is reduced to 0.0001g when the slope is close to failure. For a given analysed slope, several hundred increments are used to obtain a relatively accurate solution. In each increment, the ceiling on the iteration number is set as 800. If within a very small increment, such as 0.0001g, the number of iterations to achieve a convergent solution exceeds the ceiling, the slope is regarded to have failed. In case the non-convergent solution is caused by numerical problems, the horizontal displacement of the corner point (point A in Figure 7.1), h, is monitored. A graph of the applied horizontal acceleration versus h can be plotted. If a very small increment of applied horizontal acceleration induces a relatively large increment in h, failure is identified. The above two sets of criteria are used to define the failure, thus avoiding the possibility that the non-convergent solution is caused by numerical difficulties. The failure mechanism from finite element method can be shown as the vectors of incremental displacement from the last increment before failure. The direction of nodal movements can be indicated as the orientation of the vectors and the magnitude of the nodal movements can be indicated by the length of the vectors. The failure mechanism is more relevant to the orientation and relative magnitude of the movement than the absolute value of increment displacement. When the failure mechanism is not clear from the graph of vectors of incremental displacement, a very small increment, 0.00001g, is used to clarify the failure mechanism.

7.3.3 Soil model


The elastic-perfectly plastic model with a Mohr-Coulomb failure criterion is adopted in the finite element analysis. The material behaviour is assumed to be linear before failure and Youngs modulus E' and Poissons ratio ' are needed. These two parameters cannot be taken into account in limit equilibrium analysis and therefore, their effect on the computed critical acceleration based on the finite element method is presented in Section 7.3.5.

169

The Mohr-Coulomb failure criterion is the failure criterion adopted in the new procedure and therefore, for the purpose of comparison, the Mohr-Coulomb failure criterion is used in the present finite element analysis. The Mohr-Coulomb failure criterion is presented in Chapter 4 as:

= c + tan

(7.5)

where c' and ' refer to the effective cohesion and angle of shearing resistance of the soil. This equation can be rewritten using principle stresses and adopted as the yield function in the finite element model, thus:

2c cos ( 1 3 +3 )sin F ({ }, {k }) = 1
It is more convenient to rewrite this equation in terms of stress invariants p', J and as:

(7.6)

2 sin + 3 1 1 2 = p 1 + sin J 2 3 2 1 3 sin 3


Therefore,

(7.7)

c F ({ }, {k }) = J 3 tan + p g ( ) = 0
where:

(7.8)

g ( ) =

sin sin sin cos + 3

170

As the Mohr-Coulomb model is assumed to be perfectly plastic, no hardening/softening law is required. The state parameter {k } is assumed constant and independent of plastic strain or plastic work. To complete the plastic part of the model a plastic potential function, P{( ), (m )}, is required. An associated flow rule, with P{( ), (m )} = F {( ), (m )}, could be adopted. If the plasticity flow rule is associated, the stress and velocity characteristics coincide, which is the basic assumption in limit analysis. The dilation angle, , affects the volume change of the soil during yielding. It is well known that the actual volume change exhibited by a soil during yielding is quite variable. However, this type of variable volumetric modelling is beyond this study. Because the associated flow rules predict far greater dilation than that observed in reality, this drawback can be partly corrected by adopting a non-associated flow rule with constant , The plastic potential function is assumed to take a similar form to that of the yield surface (Equation 7.6), but with ' replaced by . This gives:

c g ( c ) + p g pp ( ) = 0 P({ }, {m}) = J tan + pc g ( ) pc pp c


where:

(7.9)

g pp ( ) =

sin sin sin cos + 3

By prescribing the angle of dilation, , the predicted plastic volumetric strains can be controlled. As shown in Section 7.2, some researchers state that the slope stability analysis is a relatively unconfined problem and therefore, the choice of angle of dilation is less important and the value is assumed to be equal to or to zero (Zienkiewicz, Humpheson and Lewis, 1975; Griffiths and Lane, 1999). In the following analysis, the angle of dilation, , is assumed to be equal to ', 0.5' and zero in each analysed slope to examine the effect of the angle of dilation.

171

When gravity is applied to the model, the total unit weight, , is assigned to the model, which is equal to that used in the new procedure.

7.3.4 The effect of the initial stress condition on the computed critical acceleration
The application of gravity and the horizontal acceleration as boundary conditions requires the calculation of equivalent nodal forces. The nodal forces equivalent to the body force are calculated element-wise, using the body force contribution to the right-hand side vector as:

{RE } = [N ] {FG }dvol


T vol

(7.10)

where [N] is the shape function of the element, {FG} is the global body force and vol is the volume of the element. The body force vector {FG} is determined with respect to the global axes by using:

FxG cos = sin FyG

(7.11)

where is the angle made by the applied body force with the horizontal axis. The integral evaluates the area of each element, multiplies by the total unit weight of the soil and distributes the forces to all the nodes. Note that {FG} can be applied by increments. In the present study, if the initial stress field is established by application of gravity, it is applied in one hundred increments to achieve a precise solution. An alternative way to establish the initial stress field is to specify vertical and horizontal stresses, where different K0 values can be taken into account. A slope is then formed by excavation. Excavation is modelled by sequentially removing groups of elements. Removal of elements is simulated by firstly calculating the equivalent nodal forces arising from the stresses acting within these elements as:

{RE } = [B]T { }dVol [N ]T dVol


Vol Vol

(7.12)

172

where {} is the stress vector in the element and secondly, the nodal forces are applied to the elements remaining in the mesh in the opposite sense as new boundary conditions. In the current analysis, the mesh to be excavated is split horizontally into several layers and one layer is excavated from the mesh each time, and the equivalent nodal forces are calculated based on Equation 7.12. The equivalent nodal forces for each soil layer are applied to the remaining mesh in ten increments. The effect of the initial conditions on the computed critical acceleration by finite element methods can be shown in the following example. Example 7.1 A slope of gradient 1:3 is considered. The soil properties are assumed to be dry, homogenous and isotropic with c=10kPa, =200 and =20kN/m2. In the first case, gravity is applied to the model in 100 increments and the vertical and horizontal stress distributions after application of gravity are shown in Figure 7.2. The typical mesh for level ground is shown in Figure 7.3. The dashed line refers to the mesh to be excavated. If K0 is assumed to be unity, then the vertical and horizontal stress distributions after excavation are as in Figure 7.4.

173

Figure 7.2 (a)

Figure 7.2 (b) Figure 7.2 The horizontal (a) and vertical stress (b) distributions after the application of gravity

Figure 7.3 Mesh for the excavation from level ground (1:3 slope)

174

Figure 7.4 (a)

(b) Figure 7.4 The horizontal (a) and vertical (b) stress distributions after excavation from level ground (K0 = 1.0)

The horizontal stresses obtained from the application of gravity are quite different from those obtained from excavation from level ground. A question, therefore, arises: do the initial stresses affect the computed critical acceleration and the associated critical slip surface? To answer this question, four different values of K0: 0.658, 1.0, 1.2 and 1.5, are assumed to check their dependency, where 0.658 corresponds to the formula of Jaky (1944):

K 0 = 1 sin

(7.13)

The applied horizontal acceleration versus the horizontal displacement of point A (Figure 7.1) are plotted in Figure 7.5.

175

Figure 7.5 The effect of initial stress on the computed critical acceleration

It is also shown in Figure 7.5 that different initial stress conditions affect the shape of acceleration-displacement curve and the magnitude of the displacement. The analysis gives the highest horizontal displacement when the slope fails when initial stress conditions are formed by applying gravity loads. The magnitude of horizontal displacement increases as K0 increases. After the applied horizontal acceleration equal to 0.237g, applying a very small additional acceleration, will cause a relatively large horizontal displacement. All the cases tend to give an identical critical acceleration of 0.2378g. The failure mechanisms from the gravity turns-on option and the excavations with K0 equal to 0.658 and 1.5 are shown in Figure 7.6. Note that the thick line is the critical slip surface obtained from the new procedure. It can be seen from the figures that the critical slip surfaces are essentially same for different assumed initial stresses and agree well with the critical slip surface obtained from the new procedure. Extensive case studies have been carried out to validate the independence of critical acceleration and failure mechanism on the initial stress field and the above conclusion is confirmed. Since the initial stresses have little influence on the critical

176

acceleration and associated failure mechanism, the initial stress field established by excavation from level ground is used in all the following analyses and K0 from Equation 7.13 is adopted. The reason for this is that the initial stress field established by excavation gives a more clear failure mechanism when the slope fails.

7.3.5 The effect of stiffness on the computed critical acceleration


It is well known that the stiffness has a profound effect on the deformations in stress-deformation analysis. While in limit load problems, it becomes less important. As reviewed in Section 7.2, almost all researchers have reported that the factor of safety is independent of Youngs modulus. Only a few researchers, e.g. Pham and Fredlund (2003), state that the factor of safety is dependent on Poissons ratio. It is noted that the factor of safety obtained by Pham and Fredlund (2003) is based on operative, instead of limit, conditions. The effect of the stiffness on the computed critical acceleration is examined in the following example.

177

Figure 7.6 (a)

Figure 7.6 (b)

Figure 7.6 (c) Figure 7.6 Comparison of critical slip surfaces between the new procedure and finite element method under different initial stresses: (a) gravity turned on (b) K0= 0.658 and (c) K0 = 1.5

178

Example 7.2 An example slope with gradient 1:5 is considered with c=10kPa, =100 and =20kN/m3. Three different Youngs modulus values are assumed in the analysis. In the first two cases, Youngs modulus values are assumed to be constant and equal to 26, 000 kPa and 60, 000 kPa, respectively. In the third case, Youngs modulus values vary with depth as:

E = 6000 + 6000 z kPa

For each Youngs modulus value, the Poissons ratio is assumed as constant with two different values: 0.25 and 0.48. The reason for choosing 0.48 is because Pham and Fredlund (2003) state that the factor of safety is affected when a high Poisson ratio is assumed and a value of 0.48 is adopted in their analysis. A plot of the applied horizontal acceleration versus horizontal displacement of a corner point (point A in Figure 7.1) is shown in Figure 7.7. Although the stiffness affects the shape of acceleration-displacement curve and the magnitude of the displacement, all the cases tend to give an identical critical acceleration of 0.10278g. It is shown in Figure 7.7 that higher horizontal displacements occur at failure when a lower value of Youngs modulus is assumed. For the same Youngs modulus, higher horizontal displacements are achieved at failure when a higher value of Poissons ratio is assumed, which is consistent with the definition of Youngs modulus and Poisson ratio.

7.3.6 The effect of the angle of dilation on the computed critical acceleration
Normality, meaning that the plastic strain rate vector is assumed to be normal to the yield function, is a basic assumption of limit analysis. In limit equilibrium methods, normality does not appear as a condition. In the finite element method, it is not necessarily the case. By assuming different angles of dilation, the effect of angle of dilation on the computed critical acceleration can be inspected. Note that with the increase in the difference between the angle of shearing resistance and the angle of dilation, non-normality becomes higher.

179

Figure 7.7 The effect of stiffness on the computed critical acceleration

The effect of the angle of dilation on the computed critical acceleration is examined in the following example.

Example 7.3 An example slope with gradient of 1:3 is used in the following. The soil underneath the slope is assumed to be homogenous and isotropic with c=5kPa, =300 and =20kN/m3. A plot of the horizontal displacement of point A (Figure 7.1) versus the applied horizontal acceleration is shown in Figure 7.8 for different angles of dilation. It is shown in Figure 7.8 that the slope behaves as linear before point P and non-linear after. The slope with the associated flow rule behaves more stiffly. It appears that to achieve the same critical acceleration as the associated flow rule, the slope with the non-associated flow rule has to undergo larger deformation. However, due to the increase of non-normality, numerical problems arise and a non-convergent solution occurs earlier than in the slope with the associated flow rule. Especially for high frictional soil slopes with high degrees of non-normality, the critical acceleration obtained from

180

the non-convergent solution of finite element analysis needs to be justified. The effect of angle of dilation on the critical slip surface is studied in Section 7.4.2.

Figure 7.8 Effect of angle of dilation on the computed critical acceleration

7.4 COMPARISON BETWEEN THE NEW PROCEDURE AND FINITE ELEMENT METHOD IN HOMOGENOUS SLOPES
It is well known that the critical acceleration is dependent on the slope inclination angle, c'/h, the angle of shearing resistance ' and the pore water pressure coefficient Ru in homogeneous slopes. Three different gradients of slopes, 1: 5, 1:3 and 1:1.5 are chosen and different soil parameters are assumed for each slope angle. Note that only dry slopes are considered.

7.4.1 Comparison of the computed critical acceleration


The eighteen example homogenous slopes chosen for finite element analysis are listed in Table 7.1. For each example, three different angles of dilation are assumed, which are equal to , 0.5

181

and 0, to inspect the effect of dilation angle to the computed critical acceleration. The results from finite element analysis are plotted as symbols in Figure 7.9. The critical accelerations obtained from the new procedure expressed as a function of c'/h and angle of shearing resistance ' for different slope angles are shown in Figure 7.9.

Figure 7.9 (a) Comparison of critical acceleration between the new procedure and finite element method for slope angles: (a) 1:5 (b) 1:3 and (c) 1:1.5

182

Figure 7.9 (b)

Figure 7.9 (c) Figure 7.9 Comparison of critical acceleration between the new procedure and finite element method for slope angles: (a) 1:5 (b) 1:3 and (c) 1:1.5

183

Table 7.1 Homogenous slopes analysed by the finite element method Case number 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 1:1.5 1:3 1:5 Slope angle (V:H) c 0 0 5 5 5 10 10 10 0 0 5 5 10 10 10 5 10 10

20 30 10 20 30 10 20 30 20 30 20 30 10 20 30 30 20 30

Note that =20kN/m3 for all the cases and the slope height is assumed to be constant as 10m.

Some conclusions can be drawn from this analysis. 1) When angle of dilation is equal to angle of shearing resistance, meaning that the associated flow rule is used, the critical accelerations obtained from the new procedure are essentially the same as those obtained from finite element analysis. 2) When is less than , meaning that the non-associated flow rule is used, the critical accelerations obtained from finite elements are smaller than those obtained from the new procedure.

184

3) The difference in critical accelerations from the associated flow rule and the non-associated flow rule comes from the difference between and (non-normality) but is independent of the value of c'/h. 4) When non-normality is high, the critical accelerations based on the non-convergent solution from finite element analysis may come from numerical problems and the values require justification.

7.4.2 Comparison of the critical slip surface


The finite element method is used as an independent tool to validate the new procedure proposed in this thesis, especially for comparison of the critical slip surface. In the finite element method, the critical slip surface can be inferred from the vectors of incremental displacement of the last increment before failure. The critical slip surface obtained from the new procedure is plotted on the same figures to compare the solutions. As shown in Section 7.4.1, the critical accelerations obtained from different procedures are essentially the same if the associated flow rule is adopted in finite element analysis. All the examples shown in Table 7.1 were used to validate the critical slip surface obtained from the new procedure. However, only slopes 3 and 12 are shown in the following. Slopes 3 and 12 with three different angles of dilation are shown in Figure 7.10 and Figure 7.11, respectively. The critical slip surfaces obtained from the new procedure are plotted as thick lines in the same figures. Some observations can be made. 1) If the associated flow rule is adopted in finite element analysis, a very good agreement is achieved in the critical slip surfaces obtained from different procedures, in terms of the starting point, the shape of the critical slip surface and the ending point. 2) The angle of dilation affects the orientation of the incremental displacement, showing that the angle of dilation equals the angle between the velocity of the movement and the critical slip surface. 3) From the results of finite element analysis, if a small value of and a equal to 0 are used, in other words, the degree of non-normality is small, there is no noticeable effect of dilation angle on the shape of the critical slip surface. However, if the degree of non-normality is high,

185

the obtained critical slip surface with the non-associated flow rule is shallower to the surface than when the associated flow rule is used. 4) When there is high non-normality, it may not be possible to identify the failure mechanism, where numerical problems arise before the slope actually fails.

186

Figure 7.10 (a)

Figure 7.10 (b)

Figure 7.10 (c) Figure 7.10 Comparison of critical slip surface between finite element analysis and the new procedure for slope 3: (a) = (b) = 0.5 (c) = 0

187

Figure 7.11 (a)

Figure 7.11 (b)

Figure 7.11 (c) Figure 7.11 Comparison of critical slip surface between finite element analysis and the new procedure for slope 12: (a) = (b) = 0.5 (c) = 0

188

7.4.3 Comparison of normal and shear stress distributions along the critical slip surface
As shown in Chapter 4, the normal stress at the end point of the slip surface of each slice along the slip surface is essential for the analysis of homogenous slopes in the new procedure. In the new procedure, the normal stress is obtained by assuming linear stress distribution along the slip surface as shown in Section 4.3.4. To verify the assumption of stress distribution, the normal stress at the same positions along the slip surface are calculated based on stress field obtained from finite element analysis. If the slip surface is drawn in the finite element stress field, the normal and shear stress at any point perpendicular to the slip surface can be determined by using Equations 7.2 and 7.3. However, x, y and xy in Equation 7.2 at the specified point are unknown. They can be obtained by first, identifying four gauss points surrounding that point and second, interpolating x, y and

xy at surrounding gauss points to compute the corresponding stresses at that point. The angle of
in Equation 7.2 is the inclination angle of the slip surface of the current slice. All the examples shown in Table 7.1 were compared but only slope 3 and 13 are shown (Figures 7.12 and 7.13). It is shown in the figures that, except for a few points, a very good agreement between the normal stress distributions is achieved. Therefore, the assumption of linear stress distribution along the slip surface used in the new procedure for homogenous slopes is justified. In the new procedure, the shear strength is fully mobilised along the slip surface. The shear stress distribution along the slip surface can also be obtained from the finite element stress field. Comparisons between the new procedure and finite elements are shown in Figure 7.12 and 7.13. It is seen that in slope 3, the shear strength is also fully mobilised along the slip surface in finite element analysis but in slope 13, the shear strength is not quite fully mobilised. Note that since the shear stresses are computed along the slip surface obtained from the new procedure, they may not be the maximum shear stress at that point.

189

(a)

(b) Figure 7.12 Comparison of normal stress (a) and shear stress (b) distributions for slope 3

190

(a)

(b)
Figure 7.13 Comparison of normal stress (a) and shear stress (b) distributions for slope 13

191

7.5 COMPARISON BETWEEN THE NEW PROCEDURE AND FINITE ELEMENT METHOD IN NON-HOMOGENOUS SLOPES
To analyse non-homogenous slopes in the new procedure, stress acceptability is satisfied by a different procedure and a concept of slip surface path is proposed to consider the effect of layered soil slopes. To validate the above assumptions some example slopes are chosen, which are analysed by the new procedure and the finite element method. Then the results obtained from different methods are compared. As shown in Section 7.4, the computed critical acceleration in finite element analysis is independent of stiffness of the soil and initial stress conditions. Therefore, Youngs modulus value is assumed to be 26, 000 kPa and the Poissons ratio is assumed to be 0.25 throughout the analysis. The initial stress is established by excavating level ground and K0 is assumed as unity. The angle of dilation is found to have an effect on the computed critical acceleration and the critical slip surface, providing that the non-normality is not small. However, limit equilibrium cannot take the non-associated flow rule into account and therefore, for the purpose of comparison, non-associated flow rule is not considered in finite element analysis for nonhomogenous slopes. Example 7.4 The profile of a simple two soil layer slope is shown in Figure 7.14 and soil properties are listed in Table 7.2, where a relatively strong soil layer overlays a weaker layer. The finite element mesh of the slope after excavation is shown in Figure 7.15.

Table 7.2 Soil properties in Example 7.4 (kN/m2) First soil layer Second soil layer 20 20 c (kPa) 10 10 (in degrees) 25 15

192

Figure 7.14 Slope profile used in Example 7.4

The slope is first analysed by the new procedure, two possible slip surface paths are designed if the slip surface encounters the soil layer boundary. The global critical slip surface is identified as -13 m below the toe point by comparing a large number of starting points and the global critical acceleration is obtained as 0.23g.

Figure 7.15 Mesh of the slope after excavation

193

The example slope is then analysed by the finite element methods. A horizontal acceleration is gradually applied to the slope until failure. The critical acceleration obtained from finite element analysis is 0.22g. The difference in critical accelerations obtained from different methods is below 5%. The vector of incremental displacement of the last increment before failure is shown in Figure 7.16 to show the failure mechanism. The critical slip surface obtained from the new procedure is plotted in the same figure to compare the solutions. It is seen in Figure 7.16 that the critical slip surface obtained from the new procedure agrees with the slip surface from the finite element analysis.

Figure 7.16 Comparison of slip surfaces in Example 7.4

The effective normal and shear stress distribution along the slip surface can be obtained in finite element analysis from the accumulated stress field at the last increment before failure. The detail of obtaining such stresses using finite elements is shown in Section 7.4.3. The results are shown in Figure 7.17. The stress distributions obtained by the new procedure are shown in the same figures. It is noticed that a very good agreement of normal stresses on the slip surface exists between the new procedure and finite element analysis. The new procedure slightly underestimates shear stresses on the slip surface for the given example, compared to the results from finite element analysis.

194

(a)

(b)
Figure 7.17 Comparison of the normal stresses (a) and shear stresses (b) on the slip surface in Example 7.4

195

Example 7.5 A three soil layer slope is shown in Figure 7.18 to validate the concept of the slip surface path; its soil properties are listed in Table 7.3. It is shown in Table 7.3 that a relatively weak soil layer exists between two relatively strong layers. The slope is first analysed by the new procedure and the global critical slip surface is identified to start at -22m below the toe point and the critical slip surface develops along the second soil layer boundary. The global critical acceleration is 0.26g. Note that the critical slip surface develops along the second soil layer boundary and the length of the critical slip surface staying along the boundary is determined by the algorithm shown in Figure 6.5, providing that variable x is regarded as an unknown.

Figure 7.18 Slope profile used in Example 7.5 Table 7.3 Soil properties in Example 7.5 (kN/m2) First soil layer Second soil layer Third soil layer 20 20 20 c (kPa) 10 10 10 (in degrees) 25 15 35

196

The critical acceleration obtained from finite element analysis is 0.254g. The difference of critical accelerations obtained between the two methods is below 3%. The vector of incremental displacement of the last increment before failure is shown in Figure 7.19. The critical slip surface obtained from the new procedure is plotted in the same figures to compare the results. Again, in terms of critical slip surface, good agreement is found between the different methods. Especially the length of the critical slip surface staying along the boundary agrees well with the failure mechanism from finite element analysis, which validates the assumption of slip surface path.

Figure 7.19 Comparison of slip surfaces in example 7.5

The normal and shear stresses on the slip surface obtained by different methods is shown in Figure 7.20. It is noticed that when the slip surface develops along the second soil layer boundary, the new procedure slightly underestimates the normal stress at the end point, compared to the stress obtained from the finite element analysis. Except for that, a good agreement is found.

197

(a)

(b)
Figure 7.20 Comparison of the normal stresses (a) and shear stresses (b) on the slip surface in Example 7.5

198

Example 7.6 Example 6.7 is reanalysed by finite element analysis. The slope profile is shown in Figure 6.10 and soil properties are shown in Table 6.8. The meshes before and after the excavation for that slope are shown in Figure 7.21. The excavated soils are split into 12 layers and one layer is excavated each time. The equivalent nodal forces for each soil layer are applied to the remaining mesh by ten increments. The critical acceleration obtained from finite element analysis is 0.1502g. The difference between the new procedure and finite element analysis is around 5%. The vector of incremental displacement of the last increment before failure is shown in Figure 7.22. The critical slip surface obtained from the new procedure is plotted in the same figures to compare the solutions.

199

Figure 7.21 (a)

Figure 7.21 (b) Figure 7.21 (a) The mesh before, and (b) after excavation

200

Figure 7.22 Comparison of slip surfaces in Example 7.6

The normal and shear stresses on the slip surface obtained by different methods is shown in Figure 7.23. There are some differences in the stresses obtained by the two methods but they are acceptable, bearing in mind they are two totally different methods.

201

(a)

(b)
Figure 7.23 Comparison of the normal stresses (a) and shear stresses (b) on the slip surface in Example 7.6

202

8. Conclusions and Recommendations

8.1 CONCLUSIONS
The procedure developed in this thesis is within the framework of the limit equilibrium method for slope stability analysis. The solution satisfies the stress and kinematical acceptability criteria, which guarantee that the stress conditions along the slip surface and within the slip mass do not conflict with the limited strength of soil and that the mass can slide along the slip surface if the applied acceleration is larger than the critical. The new procedure is first validated by Mohr-circle solutions for the first slice. Some example slopes taken from the literature are used to demonstrate the new procedure. Factors of safety are found that are comparable to those obtained from optimization methods but the critical slip surfaces are usually different. Within existing slope stability analysis methods using the limit equilibrium technique, the acceptability criteria are checked after the solution is obtained, and quite often these are found to be unacceptable for some parts of the slip surface. It is usually left to the experience of the users to declare whether the solutions are acceptable. The critical slip surfaces obtained from the new procedure are always acceptable. This new procedure is the first that rigorously satisfies the acceptability criteria. The solution provides information on the critical acceleration and the kinematically acceptable internal boundaries for analysis of seismic displacements of slopes using a multi-block sliding model. The new procedure is further validated through the finite element analysis both for homogenous and non-homogenous slopes. There is good agreement between the results obtained using the new procedure and finite element analysis in terms of computed critical acceleration, the critical slip surface and the stress distribution along the slip surface. In the finite element analysis, the angle of dilation is found to have an effect on the computed critical acceleration and the critical slip surface, providing that the non-normality is not small.

203

8.2 RECOMMENDATIONS FOR FUTURE STUDY


1. Propose a satisfactory effective normal stress distribution on the slip surface and the inter-slice boundary. The requirement for the distribution is that it can produce an acceptable line of thrust and rigorously satisfy moment equilibrium. 2 Include a more variable definition of water pressure and the effect of cyclic pore water pressure into the method. 3 Expand the new procedure to accommodate more complicated geometrical conditions and more variable soil conditions. 4 Validate the new procedure with pore water pressure using finite element analysis. 5 Combine the new procedure and the multi-block dynamic model proposed by Chlimintzas (2002) to compute the seismic displacement. Validate the computed seismic displacement via the composite procedures with those obtained by dynamic finite element analysis. 6 Develop user-friendly software for application of the new procedure. 7 extend the acceptability criteria to determine the critical slip surface in three dimensional slope analysis.

204

References

Ambraseys, N. & Srbulov, M. 1995, "Earthquake induced displacements of slopes", Soil Dynamics and Earthquake Engineering, vol. 14, no. 1, pp. 59-71. Avriel, M. 1976, Nonlinear programming: analysis and methods, Englewood Cliffs ; London : Prentice-Hall. Baker, R. & Garber, M. 1978, "Theoretical analysis of the stability of slopes", Geotechnique, vol. 28, no. 4, pp. 395-411. Baker, R. 1980, "Determination of the critical slip surface in slope stability computations", International Journal for Numerical and Analytical Methods in Geomechanics, vol. 4, no. 4, pp. 333-359. Baker, R. 2005, "Variational slope stability analysis of materials with nonlinear failure criterion", Electronic Journal of Geotechnical Engineering, vol. 10A, p. 19. Bardet, J. P. & Kapuskar, M. M. 1989, "Simplex analysis of slope stability", Computers and Geotechnics, vol. 8, no. 4, pp. 329-348. Bellman, R. 1957, Dynamic Programming, Princeton N.J.: Princeton University Press. Bishop, A. W. 1955, "The use of the slip circle in the stability analysis of slopes", Geotechnique, vol. 5, no. 1, pp. 7-17. Bishop, A. W. & Morgenstern, N. R. 1960, "Stability coefficients for earth slopes", Geotechnique, vol. 10, no. 4, pp. 129-50. Boutrup, E. & Lovell, C. W. 1980, "Searching techniques in slope stability analysis", Engineering Geology, vol. 16, no. 1-2, pp. 51-61. Brent, R. P. 1973, Algorithms for Minimization without Derivatives, Prentice-Hall, Englewood Cliffs, New Jersey. Bromhead, E. N. 1992, The stability of slopes, 2nd edn, London: Blackie Academic. Castillo, E. & Luceno, A. 1982, "A critical analysis of some variational methods in slope stability analysis", Journal for Numerical and Analytical Methods in Geomechanics, vol. 6, no. 2, pp. 195209. Celestino, T. B. & Duncan, J. M. 1982 "Simplified search for noncircular slip surfaces", 10th International Conference on Soil Mechanics and Foundation Engineering, Stockholm, Sweden, pp. 391-394. Chen, W. F., Giger, M. W., & Fang, H. Y. 1969, "On the limit analysis of stability of slopes", Soils and Foundations, vol. 9, no. 4, pp. 23-32.

205

Chen, W. F. & Giger, M. W. 1971, "Limit analysis of stability of slopes", Journal of Soil Mechanics and Foundation Division, ASCE, vol. 97, no. 1, pp. 19-26. Chen, W. F. 1975, Limit analysis and soil plasticity, Amsterdam; Oxford: Elsevier. Chen, Z. Y. & Shao, C. M. 1988, "Evaluation of minimum factor of safety in slope stability analysis", Canadian Geotechnical Journal, vol. 25, no. 4, pp. 735-748. Chen, Z. Y. 1992, "Random trials used in determining global minimum factor of safety of slopes", Canadian Geotechnical Journal, vol. 29, no. 2, pp. 225-233. Cheng, Y. M. 2003, "Location of critical failure surface and some further studies on slope stability analysis", Computers and Geotechnics, vol. 30, no. 3, pp. 255-267. Chlimintzas, G. O. 2003, Seismic displacements of slopes using multi-block, Ph.D. Dissertation, Imperial College London, Civil Engineering Department, Soil Mechanics & Engineering Seismology Section. Clough, R. W. & Woodward, R. J. 1967, "Analaysis of embankment stresses and deformations", Journal of Soil Mechanics and Foundation Division, ASCE, vol. 93, no. 4, pp. 529-550. Davison, W. C. 1959, Variable metric method for minimization, Argonne National Laboratory, ANL-5990. de Josselin de Jong, G. 1980, "Application of the calculus of variations to the vertical cut off in cohesive frictionless soil", Geotechnique, vol. 30, no. 1, pp. 1-16. de Josselin de Jong, G. 1981, "A variational fallacy", Geotechnique, vol. 31, no. 4, pp. 289-190. Donald, I. B. & Chen, Z. 1997, "Slope stability analysis by the upper bound approach: fundamentals and methods", Canadian Geotechnical Journal, vol. 34, no. 6, pp. 853-862. Duncan, J. M. & Chang, C. Y. 1970, "Nonlinear analysis of stress and strain in soils", Journal of Soil Mechanics and Foundation Division, ASCE, vol. 96, no. 5, pp. 1629-1653. Duncan, J. M. 1996, "State of the art: Limit equilibrium and finite-element analysis of slopes", Journal of Geotechnical Engineering, vol. 122, no. 7, pp. 577-596. Duncan, J. M. & Wright, S. G. 2005, Soil strength and slope stability Hoboken, N.J.: Wiley. Espinoza, R. D., Bourdeau, P. L., & Muhunthan, B. 1994, "Unified formulation for analysis of slopes with general slip surface", Journal of Geotechnical Engineering, vol. 120, no. 7, pp. 11851204. Fellenius, W. 1936, "Calculation of the stability of earth dams", Transactions, 2nd Congress on Large Dams, Washington, pp. 445-462. Fletcher, R. & Powell, M. J. D. 1963, "A rapid convergent method for minimization", Computer Journal, vol. 6, pp. 163-168. Fredlund, D. G. & Krahn, J. 1977, "Comparison of slope stability methods of analysis", Canadian Geotechnical Journal, vol. 14, no. 3, pp. 429-439.

206

Friedli, D. & Giger, M. 1978, "Discussion on 'Calculus of variations applied to stability of slopes' by Revilla and Castillo (1978)", Geotechnique, vol. 28, no. 2, pp. 222-223. Greco, V. R. 1996, "Efficient Monte Carlo technique for locating critical slip surface", Journal of Geotechnical Engineering , vol. 122, no. 7, pp. 517-525. Griffiths, D. V. & Lane, P. A. 1999, "Slope stability analysis by finite elements", Geotechnique, vol. 49, no. 3, pp. 387-403. Hinton, E. & Campbell, J. S. 1974, "Local and global smoothing of discontinuous finite element functions using a least squares method", International Journal for Numerical Methods in Engineering, vol. 8, no. 3, pp. 461-480. Huang, S. L. & Yamasaki, K. 1993, "Slope failure analysis using local minimum factor-of-safety approach", Journal of Geotechnical Engineering, vol. 119, no. 12, pp. 1974-1987. Izbicki, R. J. 1981, "Limit plasticity approach to slope stability problems", Jounal of Soil Mechanics and Foundation Division, ASCE, vol. 107, no. 2, pp. 228-233. Janbu, N. 1954 "Application of composite slip surface for stability analysis", Proceedings of European conference on stability of earth slopes, Stockholm, pp. 43-49. Janbu, N., Bjerrum, L., & Kjaernsli, B. 1956, "Soil mechanics applied to some engineering problems", Norwegian Geotechnical Institute Publication, vol. 16, pp. 1-93. Janbu, N. 1973 "Slope stability computations", Embankment Dam Engineering, Casagrande Memorial Volume, Wiley-Intersci, Div of John Wiley & Sons, Inc, New York, pp. 47-86. Karal, K. 1977, "Energy method for soil stability analyses", Journal of Soil Mechanics and Foundation Division, ASCE, vol. 103, no. 5, pp. 431-445. Keefer, D. K. 1984, "Landslides caused by earthquakes", Geological Society of America Bulletin, vol. 95, no. 4, pp. 406-421. Khazai, B. & Sitar, N. 2004, "Evaluation of factors controlling earthquake-induced landslides caused by Chi-Chi earthquake and comparison with the Northridge and Loma Prieta events", Engineering Geology, vol. 71, no. 1-2, pp. 79-95. Kierzenka, J. & Shampine, L. F. 2001, "A BVP Solver based on Residual Control and the MATLAB PSE", ACM TOMS, Vol. 27, No. 3, pp. 299-316. Kim, J. Y. & Lee, S. R. 1997, "Improved search strategy for the critical slip surface using finite element stress fields", Computers and Geotechnics, vol. 21, no. 4, pp. 295-313. Kramer, S. L. 1996, Geotechnical earthquake engineering, Prentice Hall. Kulhawy, F. H., Seed, H. B., & Duncan, J. M. 1969, Finite element analyses of stresses and movements in embankments during construction: a report of an investigation, College of Engineering, University of California.

207

Li, K. S. & White, W. 1987, "Rapid evaluation of the critical slip surface in slope stability problems", International Journal for Numerical and Analytical Methods in Geomechanics, vol. 11, no. 5, pp. 449-473. Lighthall, P. 1979, Dimensionless Charts for critical acceleration and static stability of earth slopes, MSc Dissertation, Imperial College London, Civil Engineering Department, Soil Mechanics & Engineering Seismology Section. Loukidis, D., Bandini, P., & Salgado, R. 2003, "Stability of seismically loaded slopes using limit analysis", Geotechnique, vol. 53, no. 5, pp. 463-479. Makdisi, F. I. & Seed, H. B. 1978, "Simplified procedure for estimating dam and embankment earthquake-induced deformations", Journal of Soil Mechanics and Foundation Division, ASCE, vol. 104, no. 7, pp. 849-867. Malkawi, A. I. H., Hassan, W. F., & Sarma, S. K. 2001, "Global search method for locating general slip surface using Monte Carlo techniques", Journal of Geotechnical and Geoenvironmental Engineering, vol. 127, no. 8, pp. 688-698. Malkawi, A. I. H., Hassan, W. F., & Sarma, S. K. 2001, "An efficient search method for finding the critical circular slip surface using the Monte Carlo technique", Canadian Geotechnical Journal, vol. 38, no. 5, pp. 1081-1089. MATLAB. Manual of optimization toolbox. [3.0]. 2004. The MathWorks, Inc., Natick, MA. Matsui, T. & San, K. C. 1992, "Finite element slope stability analysis by shear strength reduction technique", Soils and Foundations, vol. 32, no. 1, pp. 59-70. McCombie, P. & Wilkinson, P. 2002, "The use of the simple genetic algorithm in finding the critical factor of safety in slope stability analysis", Computers and Geotechnics, vol. 29, no. 8, pp. 699-714. Michalowski, R. L. 1995, "Slope stability analysis: a kinematical approach", Geotechnique, vol. 45, no. 2, pp. 283-293. Morgenstern, N. R. & Price, V. E. 1965, "Analysis of stability of general slip surfaces", Geotechnique, vol. 15, no. 1, pp. 79-93. Narayan, C. G. P., Bhatka, V. P., & Ramamurthy, T. 1982, "Nonlocal variational method in stability analysis", American Society of Civil Engineers, Journal of the Geotechnical Engineering Division, vol. 108, no. 11, pp. 1443-1459. Nelder, J. A. & Mead, R. 1965, "A simplex method for function minimization", Computer Journal, vol. 7, no. 308, p. 313. Newmark, N. M. 1965, "Effects of earthquakes on dams and embankments", Geotechnique, vol. 15, no. 2, pp. 139-160. Nguyen, V. U. 1985, "Determination of critical slope failure surfaces", Journal of Geotechnical Engineering, vol. 111, no. 2, pp. 238-250.

208

Parise, M. & Jibson, R. W. 2000, "Seismic landslide susceptibility rating of geologic units based on analysis of characteristics of landslides triggered by the 17 January, 1994 Northridge, California earthquake", Engineering Geology, vol. 58, no. 3-4, pp. 251-270. Pham, H. T. V. & Fredlund, D. G. 2003, "The application of dynamic programming to slope stability analysis", Canadian Geotechnical Journal, vol. 40, no. 4, pp. 830-847. Potts, D. M. & Zdravkovic, L. 1999, Finite element analysis in geotechnical engineering: Theory, Thomas Telford, London, UK. Potts, D. M. 2003, "Numerical analysis: A virtual dream or practical reality?", Geotechnique, vol. 53, no. 6, pp. 535-573. Powell, M. J. D. 1970, "A Fortran subroutine for solving systems of nonlinear algebraic equations," in Numerical Methods for Nonlinear Algebraic Equations, P. Rabinowitz, ed., London : Gordon and Breach, pp. 87-114. Prater, E. G. 1979, "Yield accelerations for seismic stability of slopes", American Society of Civil Engineers, Journal of the Geotechnical Engineering Division, vol. 105, no. 5, pp. 682-687. Reklaitis, G. V., Ravindran, A., & Ragsdell, K. M. 1983, Engineering optimization: methods and applications New York ; Chichester : Wiley. Revilla, J. & Castillo, E. 1977, "Calculus of variations applied to stability of slopes", Geotechnique, vol. 27, no. 1, pp. 1-11. Sarma, S. K. 1973, "Stability analysis of embankments and slopes", Geotechnique, vol. 23, no. 3, pp. 423-433. Sarma, S. K. & Bhave, M. V. 1974, "Critical acceleration versus static factor of safety in stability analysis of earth dams and embankments", Geotechnique, vol. 24, no. 4, pp. 661-665. Sarma, S. K., 1975, "Seismic stability of earth dams and embankments", Geotechnique, vol. 25, no. 4, pp. 743-761 Sarma, S. K. 1979, "Stability analysis of embankments and slopes", American Society of Civil Engineers, Journal of the Geotechnical Engineering Division, vol. 105, no. 12, pp. 1511-1524. Sarma, S. K. 1981, "Seismic displacement analysis of earth dams", Jounal of Soil Mechanics and Foundation Division, ASCE, vol. 107, no. 12, pp. 1735-1739. Sarma, S. K. 2000, Stability analysis of embankments and slopes by Sarma (1973) method user's mannual for the computer programe EQS, Internal report, Imperial College London, Civil Engineering Department, Soil Mechanics & Engineering Seismology Section. Sarma, S. K. & Chlimintzas, G. O. 2001 "Co-seismic & post-seismic displacements of slopes", XV ICSMGE TC4 Satellite Conference on "Lessons Learned from Recent Strong Earthquakes", Istanbul, pp. 183-188.

209

Sarma, S. K. 2004 "Critical slip surface in slope stability analysis", R. J. Jardine, D. M. Potts, & K. G. Higgin, eds., Advances in geotechnical engineering: The Skempton Conference, Thomas Telford, pp. 955-966. Sarma, S. K. & Tan, D. 2006, "Determination of critical slip surface in slope analysis", Geotechnique, vol. 56, no. 8, pp. 539-550. Siegel, R. A., Kovacs, W. D., & Lovell, C. W. 1981, "Random surface generation in stability analysis", Jounal of Soil Mechanics and Foundation Division, ASCE, vol. 107, no. 07, pp. 9961002. Sokolovski, V. V. 1960, Statics of soil media, London: Butterworths Scientific. Spencer, E. 1967, "Method of analysis of stability of embankments assuming parallel inter-slice forces", Geotechnique, vol. 17, no. 1, pp. 11-26. Tan, D. & Sarma, S. K. 2006 "Static and seismic slope stability analysis and finite element validation", First European Conference on Earthquake Engineering and Seismology, Geneva, paper ID: 217. Taylor, D. W. 1937, "Stability of earth slopes", Boston Society of Civil Engineers -- Journal, vol. 24, no. 3, pp. 197-246. U.S.Army Corps of Engineers 1967, Stability of slopes and foundations, Engineering manual, Vicksburg, Miss. Whitman, R.V., and Bailey, W.A.. Wang, W. & Xu, B. 1984 "Brief introduction of landslides in loess in China", Proceedings of fourth international symposium on landslide, Toronto, pp. 197-207. Whitman, R. W. & Bailey, W. A. 1967, "Use of computers for slope stability analysis", American Society of Civil Engineers, Journal of the Geotechnical Engineering Division, vol. 93, no. 4, pp. 475-498. Yamagami, T. & Ueta, Y. 1988 "Search for noncircular slip surfaces by the Morgenstern-Price method", in Numerical methods in geomechanics : proceedings of the sixth International Conference on Numerical Methods in Geomechanics, G. Swoboda, ed., A.A. Balkema, Innsbruck, pp. 1335-1340. Yamagami, T. & Ueta, Y. 1988 "Search for critical slip lines in finite element stress fields by dynamic programming", in Numerical methods in geomechanics : proceedings of the sixth International Conference on Numerical Methods in Geomechanics, G. Swoboda, ed., A.A. Balkema, Innsbruck, pp. 1347-1352. Zheng, H., Liu, D. F., & Li, C. G. 2005, "Slope stability analysis based on elasto-plastic finite element method", International Journal for Numerical Methods in Engineering, vol. 64, no. 14, pp. 1871-1888. Zheng, H., Liu, D. F., & Li, C. G. 2005, "Slope stability analysis based on elasto-plastic finite element method", International Journal for Numerical Methods in Engineering, vol. 64, no. 14, pp. 1871-1888.

210

Zhou, H. 1982, New calculation method in optimal design (in Chinese) Xinshidai Chubanshe. Zhu, D. Y., Lee, C. F., & Jiang, H. D. 2003, "Generalized framework of limit equilibrium methods for slope stability analysis", Geotechnique, vol. 53, no. 4, pp. 377-395. Zienkiewicz, O. C., Humpheson, C., & Lewis, R. W. 1975, "Associated and non-associated visco-plasticity and plasticity in soil mechanics", Geotechnique, vol. 25, no. 4, pp. 671-689. Zienkiewicz, O. C. & Zhu, J. Z. 1992, "Superconvergent patch recovery and a posteriori error estimates. Part 1: the recovery technique", International Journal for Numerical Methods in Engineering, vol. 33, no. 7, pp. 1331-1364. Zolfaghari, A. R., Heath, A. C., & McCombie, P. F. 2005, "Simple genetic algorithm search for critical non-circular failure surface in slope stability analysis", Computers and Geotechnics, vol. 32, no. 3, pp. 139-152. Zou, J. Z., Williams, D. J., & Xiong, W. L. 1995, "Search for critical slip surfaces based on finite element method", Canadian Geotechnical Journal, vol. 32, no. 2, pp. 233-246.

211

Appendix A. Parabolic Effective Normal Stress Distribution

As stated in Section 4.4.4, the linear effective normal stress distribution along the inter-slice boundary cannot satisfy the force equilibrium along that slice. And as confirmed in Example 5.6, linear stress distribution along the inter-slice boundaries can not produce a satisfactory line of thrust in some cases. Therefore, a modified normal stress distribution is presented below to obtain a satisfactory line of thrust. The effective normal stress distribution along the slip surface is still assumed as linear and therefore, Equations 4.36 is valid. We still assume that the effective normal stress at point A, i, is equal on planes AB and AD (figure 4.5). If the effective normal stress distribution on the plane AD is assumed to be parabolic, then the effective normal stress at an arbitrary point A' (Figure 4.5), ', on the plane AD can be expressed as:

= i + Ai l + Bi l 2
where l is the length of AA', measured from the point A along the plane AD. Then, i expressed:

(A.1)

is

i 0 = i + Ai d i + Bi d i 2

(A.2)

i 0 can also be obtained from Mohr circle if the plane AD is assumed as a failure plain at point D,
therefore:

i 0 = ci 0 cos i 0
where ci0 and i0 are the soil parameters at the point D. And E'i can be estimated:

(A.3)

212

Ei =

di

dl = id i +

1 1 2 3 Ai d i + Bi d i 2 3

(A.4)

By combination of Equations B.1 through B.4, the parameters can be defined:

Ai =

2 3E i i 0 d i 2 id i di
2

[(

(A.5)

Bi =

3 i 0 d i 2 Ei + id i di
3

[(

(A.6)

If the parabolic effective normal stress distribution on inter-slice planes is assumed, the effective normal force E'L on the small part of slice boundary AA' (Figure 4.5) can be expressed as:
l l

= EL

dl = i + Ai l + Bi l 2 dl
0

= il +

1 1 2 3 Ai l + Bi l 2 3

(A.7)

and

+ PW L EL = EL
Then, XL can be defined using the following equations:
X L = (E L PW L ) tan i + cil When = 0, the values and the differential forms of EL and XL are:

(A.8)

(A.9)

E L

=
=0

iLi cos( i + i )

(A.10)

=0 = 0 EL

(A.11)

213

X L

=
=0

iLi Li tan ii + cii cos( i + i ) cos( i + i )

(A.12)

X L =0 = 0

(A.13)

By comparing Equations B.10 through B.13 with Equations 4.38 through 4,40 resulted from linear effective normal stress distribution assumption, they are identical. It means that the different stress distribution does not affect the analysis for homogenous slopes. However, in the analysis of non-homogenous slopes, the actual values of EL and XL for an arbitrary plane are needed. The stress distribution does affect the analysis for non-homogenous slope. However, as shown in Chapter 6, further constraints are applied to non-homogenous slope if the stress acceptability is violated and therefore, the linear effective normal stress distribution is used to obtain initial guess of the new slice. From the parabolic normal stress distribution, the application point of E'i acting on the plane AD can be obtained by taking moment about A.
di

l = Eiz i

(A.14)

{ l + A l
di 0 i i

+ Bi l 3 dl = Eiz i

(A.15)

A B 1 id i2 i d i3 i d i4 3 4 zi = 2 Ei
The application point of total force Ei acting on the plane AD, zi is expressed:

(B.16)

z i = (E i z i + PWi z w ) / E i

(4.17)

where zw is the application point of PWi acting on the plane AD. Then the application points of applications of Ni forces can be obtained by the governing equation 4.82.

214

Appendix B. Analysis for the First Slice

B.1 GEOMETRY AND FORCES ACTING ON THE FIRST SLICE AND INITIAL CONDITIONS
The geometry and forces acting on the first slice are shown in Figure 4.8. Two angles, 1 and 2, are needed to define point B (Equations 4.15 and 4.16 are still valid with replacing i by 1). Four unknown forces, the effective normal and shear forces on the slip surface, N'1 and T1, and the effective normal and shear forces on the inter-slice boundary, E'2 and X2, are the four unknowns in terms of forces to be solved (Equations 4.20, 4.21, 4.22 and 4.23 are still valid with replacing i by 1). An arbitrary point on the slope surface can be picked as a starting point and the length of the first inter-slice boundary, d1 is zero, therefore, geometry initial condition can be defined. If there is no surcharge in the slope surface (any surcharge can be included in the analysis), E'1 and X1 are zero. If two angles, 1 and 2, are assumed, which can be obtained by applying acceptability criterion in the first slice, the geometry and force unknowns can be determined,.

B.2 THE APPLICATION OF ACCEPTABILITY CRITERION IN THE FIRST SLICE


As shown in Figure 4.8, the definition of slip surface side and inter-slice boundary side for a standard slice in Section 4.4.3 is not valid for the first slice. Another point D, which is the centre point of BC, is used to divide the wedge into two parts. The segment ABD is named the slip surface side and the segment BCD is named the inter-slice boundary side for the first slice (Figure 4.8). A plane inclined at an angle to the slip surface creates a small segment ABA' inside the slip mass as shown in Figure 4.8. The reciprocal of the factor of safety of the plane BA', mS, is defined as:

215

S m =

(N

S U ) tan S + c L

(B.1)

where N and T are the normal and shear forces in terms of total stresses acting on the plane BA', U is the force due to pore water acting on the plane, L is the length of BA' and S and cS are the weighted average shear strength parameters on the plane. From equilibrium of the segment ABA':

N = N1 cos + T1 sin WS sec cos( + 1 ) T = N1 sin + T1 cos WS sec sin ( + 1 )

(B.2)

(B.3)

Note the difference between Equations B.2 & B.3 and Equations 4.84 & 4.85. The first slice starts from the known stress conditions; therefore there is no assumption about stress distribution involved. From the geometry of the small segments ABA', l and L, which are the lengths of plane AA' and A'B respectively, and WS is the weight of the small segment ABA' can be obtained as

l = L1

sin sin ( + 1 ) sin ( 1 ) sin ( + 1 )

(B.4)

L = L1

(B.5)

W =

1 2 sin ( 1 )sin L1 2 sin ( + 1 )

(B.6)

Differentiating Equation 4.83, we can obtain

216

S m S m

S (N U ) tan S + c L

=0

=0

N U L S tan S + c =0 =0 = 0 =0 =0 =0 S S T c S ( N U ) =0 sec 2 S + m L =0 = =0 = 0 =0 =0

(B.7)

=0

Using same assumption as 4.4.5.1, Equation B.7 becomes:

=0

tan S =0

=0

)+ c

=0

=
=0

(B.8)
=0

Differentiating Equations B.2 and B.3, we can obtain:

N WS = N1 sin + T1 cos sec cos( + 1 ) WS sec sin ( + 1 )

(B.9)

T WS = N1 cos T1 sin sec sin ( + 1 ) + WS sec cos( + 1 )

(B.10)

Using the definition of pore water pressure in section 4.4.6, the forces due to pore water pressure and the difference of the force can be expressed as: 1 u 2 L 2

U =

(B.11)

=0

L1 1 = u2 2 tan ( 1 )

(B.12)

where u2 is the pore water pressure at the point B. By combination of Equations B.1 through B.12, we can obtain the first equation to solve the geometry of the first slice:

217

N1 + T1 tan 1 c1

L1 1 2 L1 sec sec 1 sin (1 1 ) + tan ( 1 ) 2

L1 1 u2 tan 1 = 0 2 tan ( 1 )

(B.13)

In the inter-slice boundary side of the first slice, the analysis is the same as the analysis for the standard slice. Therefore, by replacing i by 1, the second equation to solve the geometry of the first slice can be obtained as:

d 2 tan ( 2 + ) E 2 + X 2 tan 2 c 2

1 2 d 2 tan ( 2 + ) + d 2 sec sec 2 cos 2 2 + c 2 2 1 u 2 d 2 tan ( 2 + ) tan 2 = 0 2

(B.14)

218

Appendix C. Mohr Circle Analysis For Passive Wedge

Let us consider a rhombic element of soil, with sides vertical and top and bottom inclined at angle

to the horizontal, at depth H in a semi-infinite mass. The vertical stress on the top plane is
inclined at angle to horizontal. The vertical stress at depth H on a plane inclined at angle to the horizontal can be assumed as:

Z = H cos

(C.1)

and is represented by the distance OB on the Mohr-Circle diagram (Figure C.1). Then the normal and shear stresses at point B due to seismically load are:

B = H cos sec cos( + )


B = H cos sec sin ( + )

(C.2)

(C.3)

It is noted that the approximation of stresses as above are only valid for semi-infinite mass. The two angles 1 and 2, which are the directions of the two sets of failure planes relative to the ground surface, can be obtained based on Mohr-circle as following: Figure A.1 shows the Mohrs circle diagram for the passive wedge, where c is the cohesion and is the angle of shearing resistance; Op and Rp are the centre and the radius of the Mohr circle; Pp is the pole; xB and yB are the horizontal and vertical coordinates of point B; xP and yP are the horizontal and vertical coordinates of point Pp. From the diagram:

1 =

P 4 2

(C.4)

219

2 =

+P 4 2

(C.5)

BOP =

OB sin ( + ) sin 2 P

(C.6)

AOP = O P sec tan

(C.7)

AOP = O OP sin
Equations C.6, C.7 and C.8 can be rewritten as:

(C.8)

RP =

q sec sin ( + ) sin 2 P

(C.9)

RP = [H sec cos( + ) + c cot + RP cos 2 P ]sin

(C.10)

RP =

[H sec cos( + ) + c cot ]sin


1 sin cos 2 P

(C.11)

Substitute equation C.9 into equation C.11 and arrange the terms:

sin 2 P H sec sin ( + ) = H sec cos( + ) tan + c sec tan cos 2 P


If we define mp as:

(C.12)

mp =

H sec sin ( + ) H sec cos( + ) tan + c

(C.11)

which is the degree of shear strength mobilisation along the inclined ground surface, then:

220

P =

m P sec 1 1 sin 2 2 2 1 + m P tan

tan 1 (m tan ) P

(C.13)

By substituting the value of P into Equations C.4 and C.5, the two angles 1 and 2 can be obtained, which are the directions of the two sets of failure planes relative to the ground surface.

A
1

PP

P
O' O c

B A' OP

ccot

Figure C.1 Mohr circle analysis for the first slice

221

You might also like