CIVL2006 Notes

Embed Size (px)

DESCRIPTION

soil

Citation preview

  • 1

    SOIL MECHANICS: PART B

    B1. SHEAR STRENGTH AND FAILURE CRITERIA

    Strength is an essential soil parameter used in the design of geotechnical structures. It is the

    ultimate shear stress that the soil can sustain before it fails. Obviously we do not design for

    geotechnical structures to be on the brink of failure, but it must be proven that the structure is

    safe. We start this section with a reminder of Mohr circles applied to soils, as this will be

    useful to demonstrate the application of strength to design in section B3.

    B1.1 Mohr circle applied to soils

    Mohr circle is the representation of the state of stress (, ) on any plane at a point in a two-dimensional field (assuming plane strain conditions: no movement in the direction

    perpendicular to the plane). A point in a stress field is conventionally represented by a small

    element to help us visualize (small element A in Figure 1).

    A

    Figure 1 Small soil element (A) behind a retaining wall.

    On each face of the element are acting a normal stress () and a stress stress (). In soil mechanics the sign convention is: for normal stresses, compression is positive, and for shear

    stresses, anti-clockwise is positive.

    Lets consider an element inclined at degrees from the horizontal, and call the stresses

    acting on the bottom plane (, ) (Figure 1). The stresses acting on the perpendicular plane

    are (, ), with = (otherwise the element would spin) and is positive if it is

    on the element anti-clockwise. The 2-D plane contains the principal stresses and which can be resolved as follows (Figure 2).

    a

    b

    c

    ab=1

    unit l

    ength

    Figure 2 Principal stresses in the ground.

  • 2

    cos + sin = 1cos (1)

    sin cos = 3 sin (2)

    We obtain:

    22sin 31

    (3)

    2cos22

    3131

    (4)

    Equations 3 and 4 describe a circle of centre [

    2

    31 ; 0] and radius equal to

    2

    31

    (see Figure 3).

    2

    31

    2

    31

    s

    Figure 3 Mohr circle of soil stresses.

    2

    31 s is the mean stress and 2

    31 t is the maximum shear stress.

    Note that the angle was measured from the major principal plane, i.e. the plane on which

    acts. Lets now assume that the principal stress plane is at an angle from the horizontal, with axes (x,y) (x and y are the horizontal and vertical axes on the page; z is the horizontal

    axis perpendicular to the page; Figure 4):

  • 3

    yx

    xy

    y

    x

    1

    Horizontal plane on

    which y acts.

    Figure 4 Stresses acting on small soil element.

    The origin of planes, or pole P, gives a real physical orientation to any stress state. We can

    determine the origin of planes on Mohr circle by drawing the physical plane (which is

    horizontal) from point (c) which represents the stress pair which acts on the plane ------

    (Figure 5). That plane intersects the opposite side of Mohr circle at point P, the origin of

    planes. From trigonometry, the planes from point P to any other point on the circle (i.e. 1) will be at half the angle subtended at the centre. For example, the physical plane on which the

    major principal stress 1 acts which passes through P is at to the horizontal and 2 at the centre.

    yyx

    a

    c

    b o

    d

    P

    xxy

    Figure 5 Determination of origin of planes or pole P.

    Figure 6 below shows how to use Mohr circle and the origin of planes to determine the stress

    pair (, ) on a plane inclined at from the horizontal in the same soil element as in Figure 4. The pole P is found in the same way as in Figure 5. In the Mohr circle (Fig. 6) the

    inclined plane crosses the circle in i. the coordinates in i are:

    2cos22

    3131

    and

    22sin 31

  • 4

    yyx

    i

    c

    d

    P

    Reference plane

    x

    Figure 6 Determination of stresses acting on plane at from the horizontal.

    B1.2 Loading tests

    Soil loading tests are mainly carried out in order to determine the strength, stiffness and

    compressibility characteristics of the soil. These tests also inform us on the full stress-strain

    response of the soil under different conditions of loading and drainage.

    The compressibility of soil is determined by carrying out one-dimensional compression tests

    in an oedometer apparatus. The change in void ratio with stress during loading and unloading

    is recorded so we can plot the compression and swelling curves of the soil (refer to Dr

    Kwoks lecture notes and laboratory class). By timing the evolution of volume change with time at every stage of the loading we can also measure the consolidation characteristics of the

    soil, which inform us on how long a fine-grained soil would take to reach full settlement for

    example.

    The strength of the soil is determined by different ways, and there is not one strength but

    several that we can measure in the laboratory. The first type may be the unconfined

    compressive strength (UCS) obtained by loading vertically an unconfined (typically

    cylindrical) specimen of soil and measuring the maximum axial load it can sustain (the UCS

    is that maximum load divided by the area of the specimen). This is only possible for soil

    specimens with enough cohesion to hold together without confinement. Granular soils such

    as sands are typically subjected to shear box tests, where they are sheared along a given

    horizontal plane and under a constant normal load. In granular soils like sands, they are

    usually drained. The ultimate shear stress that the soil reaches is unique for a given normal

    stress, and independent of the density of the soil. We determine the angle of friction ' from

    the ratio of shear stress to normal stress (which is equal to tan'). The main disadvantages of the shear box test are (1) that there is no drainage control (2) each test only gives one point in

    - plane, so we cannot plot the Mohr circle for that test; see later.

    Triaxial tests are common in commercial laboratories and research laboratories. They are

    performed on cylindrical specimens confined by a surrounding cell pressure. There are two

    stresses acting on the specimen, the axial stress (a) and the radial stress (r), and because of axisymmetry they provide a three-dimensional picture of the stress state (Figure 7). The axial

    and radial stresses are principal stresses (no shear stress on boundaries) so that we know the

    Mohr circle at every stage of the test. Corresponding strains are the axial strain (a) and the

  • 5

    radial strain (r). Isotropic compression tests (a=r) and shear tests can be performed in the triaxial apparatus, and when they are more advanced both cell pressure and axial pressures

    can be controlled. Logging of enough data allows determining the evolution of stiffness with

    strain level a vital information in design as different geotechnical structures may be required to withstand different levels of strain to be serviceable.

    a

    ru

    Figure 7 Schematic diagram of triaxial specimen.

    In research laboratories more advanced tests are performed e.g. hollow cylinder or true

    triaxial tests (allow three independent stresses to be applied; good to determine the anisotropy

    of the soil), cyclic tests, dynamic tests (for offshore wave loading or earthquake loading).

    These are very rarely commissioned for site investigations, which usually consist of the

    simplest (and cheapest) tests and are carried out in commercial laboratories.

    B1.3 Strength and failure criteria

    The strength of a soil is the ultimate state of stress that it can sustain before it fails. We can

    talk of tensile strength, compressive strength, shear strength, and although they seem to be

    different they all refer to the same fundamental measurement of the soil strength. We can

    understand this if we try to determine these strengths using Mohr circles. This is illustrated in

    Figure 8. In Figure 8(a), the element is subjected to axial tension with no confining pressure.

    The tensile strength is the maximum value of shear stress (f) that the element can withstand. These types of tests are very difficult to carry out on soils. Here the Mohr circle is drawn

    using the values of horizontal stress (zero) and axial stress at failure (f), which is negative

    because of being tension. The strength f is equal to the radius of the Mohr circle. In Figure 8(b), a soil element is subjected to axial compression with no confining pressure. The Mohr

    circle is determined from the lateral stress (zero) and the axial stress at failure (f), which is

    positive. The strength f is again equal to the radius of the circle. Finally, lets consider that a block of soil has failed and is sliding along a failure plane (Figure 8(c)). The effective

    stresses on the failure plane ('f, f) determine the limiting size of the Mohr circle at failure. The strength of the material is the maximum shear stress, and is equal to the radius of the

    Mohr circle. All three strengths are thus equal to the largest Mohr circle of stresses than can

    be sustained.

  • 6

    f f f

    0 00 0

    f

    f

    f

    f

    f f f

    ff

    (a) (c)(b)

    Figure 8 (a) Tensile strength (b) compressive strength (c) mobilized strength on shear plane.

    But this does not mean that for a given soil there is a unique value of strength. We must

    consider other factors that may affect the value and nature of the strength measured, for

    example is the pore water pressure, or the stress-volume state of the soil, important? Lets consider one of the classic laboratory tests to measure strength in soil: the shear box test. If

    we carry out a test under a normal stress , we may measure a shearing resistance , but if

    we perform the test under a normal stress , we will find a shear strength . The shear strength measured depends on the externally applied normal pressure. However

    whether the soil is fully saturated or dry (for sand) has no impact on the measured strength:

    the water has no shearing resistance per se. This means that if the shear strength is , then

    =. If the soil is only partially saturated (there is air as well as water in the voids), then the water tension can act as glue between particles and give an apparent strength to the soil. This is called suction. The behaviour of partially saturated soils however is difficult and

    beyond the scope of undergraduate courses, so in this course we shall deal with soils that are

    either dry or fully saturated.

    i. Friction

    Lets come back to the shear box experiment, and consider three tests performed on saturated

    sand under three different normal effective pressures 1

  • 7

    data curve towards the origin. Fitting a straight line to curved data is often done for

    convenience. Equation (5) is usually referred to as Mohr-Coulomb equation. It will be seen

    later that this equation can be generalized to include peak states and dilatancy.

    c

    1 32

    1

    2

    3

    Figure 9 Dependence of shear strength on normal stress and definition of cohesion and angle

    of friction.

    ii. Critical state

    If we carry out a shear test in a shear box or in a triaxial apparatus, we will always reach

    failure, when the stresses and volume of the specimen become constant under continued

    shearing. This is called critical state. Failure occurs when the effective stress Mohr circle reaches the Mohr-Coulomb failure envelope (Figure 10; here c=0).

    , a

    rpole

    r a

    u

    Figure 10 Total and effective Mohr circles at failure during triaxial testing.

    Lets consider standard triaxial tests where the cell pressure (r) is kept constant while the

    axial stress (a) increases. If the specimen was consolidated isotropically prior to shearing,

    initially, a=r. Lets first consider a drained test. The pore water pressure (u) in the soil

    remains constant during the test (Figure 11). The Mohr circle increases in size as a is increasing, until failure is reached.

  • 8

    ,

    Figure 11 Total and effective Mohr circles during drained triaxial compression.

    Lets now consider an undrained test. The pore water pressure (u) is not constant (Figure 12).

    The Mohr circle increases in size as a is increasing, until failure is reached.

    ,

    Figure 12 Total and effective Mohr circles during undrained triaxial compression.

    Although there are no shear stresses on the boundaries of the specimen, the test is a shearing

    test in which the soil is taken to failure, and a failure plane develops. The stresses on any

    plane can be derived from construction from the pole. The failure plane is represented on the

    Mohr circle by the plane passing through the pole and the failure point at the intersection

    between the circle and the failure line. The maximum shear stress is at 45 from the

    horizontal, but the failure plane is inclined at an angle of 45+/2, being the angle of friction of the soil (see exercise).

    Most soils have little or no strength at zero effective confining stress. The true failure

    envelope must pass through the origin and therefore may be curved (e.g. Figure 13). Equation

    (5) is a straight line just to fit the data. In order to determine the shape of the failure envelope

    one needs to carry out tests at different confining pressures r.

  • 9

    c

    Figure 13 Real failure envelope.

    Mohr circles are fine to see the conditions at failure, but it is often useful to see the course of

    the test with stress paths in terms of invariants. The triaxial invariant stresses are:

    3

    2 rap

    (mean stress) (6)

    3

    '2'' raupp

    (mean effective stress) (7)

    raraqq ''' (deviatoric stress) (8)

    [Note that q and q' are interchangeable.]

    The triaxial strains used are:

    rav 2 (volumetric strain) (9)

    ras 3

    2

    (shear strain) (10)

    Lets now compare shearing data from drained tests performed in the shear box apparatus and in the triaxial apparatus (Figure 14). The data show similar trends, the behaviour depending

    on the initial state of the soil. For example, a dense specimen will show peak strength (point

    P on the figure) before softening to reach the stable state of critical state (point C on the

    figure). A loose specimen will harden all the way to critical state, its peak strength

    corresponding to its critical state strength. On clay specimens, dense-type behaviour is found

    for over-consolidated specimens, while loose-type behaviour is found for normally or lightly

    consolidated specimens.

  • 10

    n

    u=0

    h

    v

    r

    a

    u=0

    a

    v

    1 10 100

    /n

    h (mm)

    v

    h (mm)

    dil.

    comp.

    p

    ctanc

    p

    c

    r

    c

    1 10 100

    q/p

    a

    (%)

    v

    a

    (%)

    dil.

    comp.

    c

    r

    p

    M r

    p

    c

    c

    (a) (b)

    Figure 14 Stress-strain and volumetric responses during shearing (a) shear box test (b)

    triaxial test.

    The peak state (P), defined as max. '/n' or q'/p', is reached at small strains (1%). It

    corresponds to the maximum rate of dilation (max. v/a or v/h).

    Critical state (C), where the soil deforms in a turbulent flow, is reached at

    medium strains (10%).

    In the shear box apparatus, the strength is determined in terms of peak ('p)

    and critical state ('c or 'cs) angle of friction. In the triaxial apparatus, critical state is defined by the gradient of the critical state line M (=q'/p' at failure).

    For dilative tests, at minimum volume, q'/p'=M.

    Figure 15 illustrates typical stress paths for shear box test and standard drained

    triaxial test. Specimens denser (dry) than critical state will dilate before they reach critical state while specimens looser (wet) than critical state will contract before they reach critical state. For a given soil and a given initial stress the specimens reach

    the same critical state point. If we carry out a series of tests at different normal stress

    (shear box) or different confining pressures (triaxial test) we will define the location

    of the critical state points. They define a line, the critical state line (CSL). There is a

    unique CSL for a given soil, defined in space e-- (shear box test) or v-q'-p' (triaxial test).

  • 11

    n

    n

    e

    q

    p

    v

    p

    cM

    3

    1

    p

    c

    Critical state line

    (CSL)

    Normal

    consolidation line

    CSL

    Swellingline

    c

    p

    = n tanc

    c

    start

    (a) (b)

    Figure 15 Location of the critical state line (a) shear box test (b) triaxial test.

    A typical value for the critical state angle of friction of clays is 22-28, and for sands

    30-35 if they have rounded particles and around 40 if they are more angular (hence

    more frictional). The angle of friction 'c depends on the mineralogy and grain shapes. It can be linked to the triaxial stress ratio q/p at critical state, M, by:

    Mc = 6sin'c / (3-'c) for compressive shearing (11)

    Note that equation (11) becomes Me=6sin'e / (3+'e) for extension, so that if the angle

    of friction is constant ('e='c='cs), the M-value in compression, Mc, is different to the M-value in extension, Me. There is no intercept on the q-axis for critical states i.e. the

    ultimate strength is entirely frictional. If we plot the CSL in v-lnp', we find a straight

    line, parallel to the Normal Consolidation or Compression Line NCL (Figure 16):

    For NCL: 'ln pNv (12)

    For CSL: 'ln pv (13)

  • 12

    ln p

    v

    CSL

    iso-NCL

    p=1kPa

    Figure 16 Schematic diagram of NCL and CSL (note slope of swelling line, )

  • 13

    iii. Peak states

    We have noticed however that when a specimen is initially dense, the stress-strain

    response displays peak strength before softening to the critical state. The peak

    strength is a result from the dilation of the dense soil. We have seen earlier that in

    clays, peak envelopes are fitted to the soil with an intercept c and a peak angle of

    friction p. In sands, it is preferred to have the peak envelope passing through the

    origin (no cohesion). The peak angle of friction p is then equal to the sum of the

    critical state angle of friction (cs) and the dilation angle (p) (see Figure 17). We thus have to be careful which definition of peak angle of friction we mean.

    CSL

    p?

    c

    cs

    peak

    failure

    p

    Figure 17 Definition(s) of peak strength and peak angle of friction.

    Unlike the critical state envelope, the peak envelope is influenced by the initial stress

    and volume (state or overconsolidation ratio (OCR or R) refer to Part A by Dr Kwok).

    iv. Undrained shear strength

    The undrained shear strength is the strength of the soil when it is loaded very rapidly

    so there is no time for drainage. Unlike in drained loading, where the stress path can

    vary (e.g. constant p test, or constant r test), the undrained stress path is unique and depends on the development of pore water pressure. Thus while we can determine

    different drained strengths for the soil depending on the stress path chosen, the

    undrained shear strength is unique. It is equal to: su = cu= qmax/2 (i.e. radius of Mohr

    circle at peak or critical state as appropriate). su is often used in design for short term

    conditions, which can be most critical for some structures e.g. foundations. The use of Su can be convenient, but predicting the direction of the undrained stress path and

    hence the change of pore pressure is difficult - if p at failure cannot be estimated then the shear strength q cannot be either.

    The undrained shear strength depends on the soil stress state, thus it is important when

    trying to measure su in the laboratory to start the test from the correct in-situ stress. As

    a soil sample is extracted from the ground with little disturbance, its mean effective

    stress, which is unknown (p'0), remains constant. If the specimen is installed in a

    triaxial apparatus and subjected to undrained isotropic compression, the rigidity of the

    water will mean that any increase in total stress will be entirely taken by the pore

    water (u=p), so that the effective stress remains unchanged (p'=0) we are

  • 14

    assuming the soil is isotropic. Thus whatever the isotropic total stress imposed to the

    specimen, the pore water response during isotropic undrained compression will be

    such that the mean effective stress remains constant, equal to p'0. If the axial total

    stress is then increased, the specimen will be sheared and will eventually fail. The

    value of qmax obtained in the undrained test is equal to 2su. It is not necessary to

    measure the pore water pressure generated during the shear test only qmax is needed, equal to the size of the Mohr circle tangent to the failure line. The same Mohr circle

    will be found whatever the confining pressure on the specimen (Figure 18).

    ,

    cseffective

    stresses alternative

    total stresses

    ar ar ar a

    r

    uu

    Su

    Figure 18 Mohr circles during undrained shearing and determination of Su.

    From the critical state theory, su is only a function of the specific volume, v, which

    means it is a function of the water content only (Figure 19).

    v

    lnp'

    CSL: v= lnp'

    pcs

    A C B

    Figure 19 Volumetric stress paths during undrained shearing.

    For an undrained test, the volume is constant, thus the stress path in a v-lnp' plot is

    horizontal. If A and B are two specimens with different stresses but same volume,

    they will reach the critical state line (CSL) at the same point C. The mean effective

    stress at failure can be determined form the equation of the CSL, and the undrained

    shear strength can be derived:

    vp cs exp' (14)

    cscs pMq ' (15)

  • 15

    vMqs csu exp

    22 (16)

    For a saturated soil, v=Gs.w, thus the undrained shear strength is only dependent on

    the water content. Because of localisation of strains in tested specimens (we never

    really have homogeneous deformations throughout the soil specimen), the strength

    predicted during site investigation is often underestimated hence it is not correct in

    practice.

    The undrained shear strength is also indirectly measured when performing a fall cone

    liquid limit test. For the majority of clays, the liquid limit corresponds to a strength of

    about 2kPa, while the plastic limit corresponds corresponds to about 200kPa.

    Assuming a semi-logarithmic relationship between Su and the water content (Figure

    20), we can derive an expression for Su in function of the liquidity index LI

    (PLLL

    PLwLI

    ):

    )210200 LIus (17)

    w

    logSu

    200kPa

    2kPa

    PL LL

    Figure 20 Dependence of undrained shear strength on water content.

    v. Residual strength

    The alignment of platy shaped particles in high plasticity soils at large displacements

    allows the strength to reduce below the Critical State to the Residual Strength (Figure

    21). Because of localisation of strains we must plot displacements.

    q

    a (%)

    peak

    critical state

    1-5 10-30 shear displacement (mm)

    high plasticity soil

    low plasticity soil

    q

    a (%)

    peak

    critical state

    1-5 10-30 shear displacement (mm)

    high plasticity soil

    low plasticity soil

    Figure 21 Stress-strain response during shearing to large deformation definition of

    residual strength.

  • 16

    Summary

    Critical state is a state of constant q and constant v or u with continued shear strain: there is turbulent shearing of the soil particles.

    Critical state must not be confused with the residual strength of a plastic clay mobilised at large displacements when platy clay particles align: in this case

    we have laminar shearing.

    The critical state angle of shearing resistance, 'cs, which corresponds to the angle of repose of an infinite slope, is related to the CSL gradient, M, by:

    sin 'cs = 3M/(6+M) for compression.

    (Note that MeMc if 'cs= constant - Mohr-Coulomb)

    Drained and undrained tests from normally consolidated or overconsolidated initial states all define one unique critical state line

    at large strains, typically 10-30% s.

    Peak states are reached at small-intermediate strains (typically 0.5-5%). They result solely from work done in dilation.The peak corresponds to the

    maximum rate of dilation and hence depends on the current combination

    of stress & volume (overconsolidation ratio).

    'cs is dependent on the nature of the soil particles (shape, size, mineralogy).

    A residual state (r) can be reached at very large strains for plastic clays only: the clay deforms in a laminar flow, with deformations concentrated

    on a plane (e.g. landslides).

  • 17

    B2. APPLICATIONS OF SOIL MECHANICS IN ENGINEERING PRACTICE

    Typical problems in geotechnical engineering are the design of foundations, of slopes

    and of retaining structures. The geotechnical engineer must find solutions (i) that are

    safe from collapse (limit state) and (ii) ensure that the ground movements caused by

    construction do not damage the appearance or reduce the useful life of a structure

    (service state). In this course, we will focus on (i), in which we will find the ultimate

    load (or condition) when geotechnical structures fail (the ultimate condition). As long

    as the actual loading condition is smaller than the ultimate loading condition, the

    geotechnical structures are considered to be safe.

    There are various ways to design geotechnical structures. The two extremes are:

    Complex method: model the soil behaviour as accurately as possible and incorporate the stress-strain model into a numerical method such as the finite

    element method to design geotechnical structures. The use of numerical

    methods allows us to solve problems with complex boundary/ initial

    conditions observed in field. However, this can become a time consuming

    effort and requires expertise knowledge of the conditions and of constitutive

    models.

    Simplified method: identify the important soil parameters that are relevant to the design objective, and use relatively simple mathematics to solve a

    boundary value problem. The boundary/ initial conditions may also need to be

    simplified to obtain the solution. The simplified method leads to formulae

    available in the design codes. The solution can be obtained easily, but

    engineering judgement is required to assess the potential errors associated with

    the simplification.

    The solutions must satisfy:

    Equilibrium (the forces must balance i.e. the sum of forces in any direction or sum of moments about any point is equal to zero).

    Compatibility (the displacements must balance i.e. no holes appear and no soil is eaten up).

    Material properties (stresses and strains in the soil are correctly related by strength and stiffness).

    Usually it is not possible to find simplified solutions which satisfy all three criteria,

    due to (1) complex material behaviour (2) complex geometry of problem (boundary

    conditions) (3) self weight effects (gravity). The drawback of using simplified

    analyses is that we cannot find easily exact solutions, so we calculate upper and lower

    bound solutions as close to each other as possible. For example, Figure 29 shows the

    load-settlement curve for a foundation; a simplified solution assuming that the soil has

    failed (upper bound) will be on the unsafe side (the load or pressure determined is

    higher than the load or pressure that the soil can bear) use a factor of safety; a simplified solution assuming that the soil is in equilibrium (lower bound) will be on

    the safe side. The exact solution is in between the two.

  • 18

    qqf

    safe lower bound unsafe upper bound (a)

    (b)

    Figure 22 Illustration of upper and lower bound.

    The upper bound solution is found by calculating the load or stress that causes the soil

    to fail. By trying different mechanisms of failure different solutions are found. The

    lower the upper bound solution, the closer it is to the exact solution (Figure 22b). The

    lower bound solution is found by calculating the load or stress under which the soil is

    in equilibrium. By trying different scenarii of stress discontinuities different solutions

    are found. The higher the lower bound solution, the closer it is to the exact solution.

    When upper bound and lower bound solutions are equal, the solution is thus exact.

    B2.1 Limit equilibrium method

    The upper and lower bound plasticity methods give us a range of acceptable values

    for design; it can be the maximum load that a foundation can carry for example, or the

    maximum height than a cut can sustain without failure. Simplified solutions that do

    not even obey two of the criteria cited are often used in geotechnical engineering.

    They are known as limit equilibrium analyses, and were introduced by Coulomb in

    1776 for analyzing rigid retaining walls. They typically assume a failure mechanism

    but with the failing soil mass in equilibrium. A simple analysis of the equilibrium of

    the forces at the boundaries of the system is carried out, neglecting internal stresses.

  • 19

    Solutions are not unique need to find the worse case scenario i.e. the largest active load on a retaining wall.

    B2.2 Plasticity solutions

    The mathematical theory of plasticity can be used to find the ultimate load or

    dimension for a problem. There are two approaches:

    Upper bound approach we will find the kinematically admissible state, i.e. only mechanisms of failure are considered and compatibility of stress is

    ignored. The failure mechanism that gives the smallest ultimate solution

    should be closer to the correct solution.

    Lower bound approach we will find the statically admissible state, i.e. failure mechanisms are ignored and compatibility of stresses is ensured. The

    stress state that gives the largest ultimate solution should be closer to the

    correct solution.

    i. Stress discontinuity line and stress fan for clay

    Stresses near geotechnical structures are not all equal, for example the stress directly

    beneath a foundation is different from the stress at the edge of the foundation. There

    are thus discontinuities in the soil stress field. We can find the rotation and magnitude

    of stresses either side of a discontinuity by assuming that the stress equilibrium is

    maintained within each block and across each discontinuity line. Lets consider Figure 23, with two failing soil elements adjacent to a discontinuity line. Here, the stresses in

    the elements satisfy: (i) na = nb and na = nb and (ii) ta tb

    Figure 23 Stress equilibrium across the discontinuity line

    The stress equilibrium on the discontinuity is maintained by having the stress states

    acting on the discontinuity line the same on both sides of the discontinuity line.

    For a general stress state condition (1(A)

  • 20

    Figure 24 Rotation of major principal stress direction by

    The soil has an undrained shear strength cu, which is the same in A and B (Mohr

    circles have the same radius cu. The circles intersect at the common stress state on the

    discontinuity line (Figure 24). This stress state is acting on the horizontal planes on

    elements A and B: we can draw the physical planes where it is acting and find the

    poles for circles A and B, P(A) and P(B). The change of total stress across the

    discontinuity line is related to the rotation of the direction of the major principal

    stress. [angle between planes where A and B act: = ( ) () = =

    90 - ]

    s = 2cu cos = 2cusin= 1(B) 1(A) (18)

    Example (Figure 25): na = nb = 0. Two Mohr circles join at the tip. The major principal stress of one circle is equal to the minor principal stress of the

    other circle. 1(B) = 1(A) + 2 cu. The rotation of the major principal stress

    direction = /2 = 90.

    Figure 25 90 degrees rotation of major principal stress direction

    Example (Figure 26): Two Mohr circles join with 2cu separation. 1(B) =

    1(A) + 2cu Rotation of the major principal stress direction = 45.

  • 21

    (a)

    (b)

    Figure 26 45 degrees rotation of major principal stress direction

    Stress fan: Using the stress discontinuity concept, it is possible to have a change of stress from one region to another across a fan of discontinuities as

    shown in Figure 27. The fan angle f is equal to the rotation of the direction

    of the major principal stress across the fan. s = 2cu sin~ 2cu when is very small. Integration gives:

    s = 2cu = 2cu f (19)

    Figure 27 Stress fan

  • 22

    ii. Application No. 1: open trench in clay (vertical cut)

    Lets consider a vertical cut in clay (Figure 28).

    Upper bound solution: if the soil finds a mechanism to fail, it will. Lets now assume that the collapse mechanism is the simple planar failure illustrated in the figure.

    45

    HW

    Figure 28 Schematic diagram of a vertical cut with planar failure mechanism.

    We can calculate the work done by the soil during the failure. Because the soil is

    assumed to have failed, the equilibrium criterion is not satisfied, so our upper bound

    analysis will only satisfy compatibility (the block slides along the slip/failure plane)

    and material properties (undrained shear strength cu). We can draw a displacement

    diagram for the different forces involved: compatibility imposes that the diagram

    closes (Figure 29).

    v v

    Figure 29 Displacement diagram for vertical cut.

    We use as unit displacement v the displacement of the weight W per unit length. By

    trigonometry in the right angle triangle, we can calculate that W=H2/2, with the unit weight of the soil. The shear force per unit length on the slip plane is: T=

    (failure)x area, with (failure) = cu and area=H/cos(45)= H2; we obtain T= cu H2 (note that it is in opposite direction to the weight). From the displacement diagram it

    has moved by 2 v. The sum of the external work (by the weight) and the internal work (on slip plane) for the failing block of soil is therefore:

    H2/2 x v - cuH2 x 2 v = 0 (20)

    We thus derive the height at which this collapse mechanism would occur:

    H = 4cu/ (21)

    Lower bound solution: lets now take another approach: lets assume that the soil is in equilibrium (lower bound; Figure 30). There are no displacements so the two

    criteria satisfied in this solution are the equilibrium and the material properties.

  • 23

    Hv

    h= 0 Figure 30 Schematic diagram of a vertical cut in equilibrium.

    Figure 31 Schematic diagram of a vertical cut in equilibrium.

    We consider the equilibrium of a small element just next to the cut. The Mohr circle

    of total stresses for this element is drawn in Figure 31. The horizontal total stress is

    equal to zero. The vertical stress is equal to the weight per unit area: v = H. Because this a total stress (undrained) analysis, the radius of the Mohr circle is equal to cu, and

    we have:

    H = 2cu (22)

    The maximum height for the cut to be in equilibrium is therefore:

    H=2cu/ (23)

    We have determined a lower bound solution (H=2Su/) and an upper bound solution

    (H=4Su/) for the vertical cut problem. By refining our choice of failure mechanism we could approach a better approximation.

    iii. Application No.2: shallow strip foundation

    Lets finally consider the case of a strip shallow foundation of width L subjected to a load F. The short-term solution is undrained.

    Upper bound solution: one of the simplest mechanisms of failure we can imagine is

    a two-wedge mechanism (Figure 32). Here we take the two wedges as isosceles

    triangles (with two sides of same length L; the third side is of length L/cos(45) =

    L2). We represent the shallow foundation as a surface foundation, as it is easier to model, but we have to take account of the fact that it is embedded below the surface.

  • 24

    To do so we consider the weight of the soil at the depth of the foundation (called S in

    the figure, for surcharge). For this simple example though we will assume no effect of the soil weight. The wedge directly underneath the foundation, which moves

    downward, is called A and the corresponding wedge which moves upward is called B.

    The soil is immobile (referred to as O).

    F

    BA

    O

    S

    Figure 32 Schematic diagram of a foundation with wedge failure mechanism.

    Lets draw the displacement diagram for the different forces acting internally (movements along shear planes between O and A, and between O and B; relative

    movement between wedges A and B; Figure 33) and externally (weight of the

    structure and foundation, F, and surcharge, S). Note that here because there is no self-

    weight effect the term in S will be zero. Compatibility dictates that F and S move by

    the same amount in opposite directions, that the movements of F and A relative to O

    are the same, and similarly B and S move in the same amount relative to O.

    O(fixed)

    F

    S B

    A

    v

    2v

    Figure 33 Displacement diagram for foundation.

    We can now equate the work done by the external forces and the internal dissipated

    energy:

    F x v - 0 = cu x L2 x v2 + cu x L2 x v2+ cu x L x 2v (24)

    The maximum load that the foundation can carry before failure is:

    F = 6cuL (25)

    Lower bound solution: lets now assume that the foundation is in equilibrium. There is a discontinuity in the stresses directly beneath the foundation and those beside it

    (Figure 34).

  • 25

    F

    A BI II

    Figure 34 Schematic diagram of a foundation with vertical discontinuities.

    Lets call qF the bearing capacity of the foundation (qF = F/L). Lets consider two small elements underneath (A) and beside the foundation (B), at a depth z. The

    vertical and horizontal total stresses acting on element A are A,v = z + qF and A,h

    respectively. The stresses acting on element B are B,v = z and B,h = A,h. Lets draw the Mohr circles for these elements (Figure 35):

    Figure 35 Mohr circles for elements A and B.

    We assume that the foundation does not create any shear stresses on the horizontal

    and vertical planes. In element A, the major principal stress is the vertical stress A,v,

    and the minor principal stress is A,h. By continuity, in element B the major principal

    stress is B,h and the minor principal stress is B,v. By using the geometry of the circles and the undrained shear strength property of the soil (which defines the size of

    the total stress Mohr circles at failure), we obtain:

    A,v B,v = 4cu = (z + qF) z (26)

    qF = 4cu (27)

    By choosing more complex discontinuity patterns (e.g. fan) we would obtain a better,

    higher, lower bound value. See examples for more upper and lower bound solutions.

    B3. CONCLUSIONS

    Even though this part of the course used the ultimate state as a design criterion, it is

    really the service state that should be considered (ground deformations will affect the

    structures). We are not designing for failure but for structures to be serviceable. So

  • 26

    both strength and stiffness are important! You will come across serviceability when

    learning about retaining earth systems, foundations.

    REFERENCES

    Coulomb, C.A. (1776). Essai sur une application des rgles des maximi et minimi

    quelques problmes de statique relatifs larchitecture. Mmoire acadmique royal prsentant divers savants. Vol. 7. Paris.

    GEO manual No.2, published by the Geotechnical Engineering Office, Civil

    Engineering and Development Department, Hong Kong SAR Government

    BIBLIOGRAPHY

    Powrie, W. (1997). Soil mechanics: concepts and applications. Spon Press.

    Atkinson, J.H. (2007). An introduction to the mechanics of soils and foundations

    through critical state soil mechanics. Spon Press.