You are on page 1of 48

CHAPTER 3

Slope instability due to infiltration

3.1. The Infiltration Process and wetting front development.

When rainfall infiltrates into a soil slope, it will clearly increase the moisture content of the soil
above the phreatic surface ( and hence reduce the suctions present there ), but as the water flows
downward, it may also result in a rise in the position of the phreatic surface. Such a rise could be
the cause of some slope instability, but as has been proven, for example by Boosinsuk and Yung
(1992), failure may be induced by direct rainfall, rather than by rising groundwater.

The process of infiltration into a soil has been studied by a number of authors, for example Philip
(1957a), Lumb (1962a). They show that the processes at work include diffusion of water due to a
moisture content differential, but also a pressure effect ( arising from gravity potential and
pressure differential ), which is sensitive to the ( saturated ) coefficient of permeability of the soil.

Lumb (1975) suggests that for prolonged infiltration, the diffusion element becomes insignificant
compared to the pressure effect, and so may be ignored. It is stated that for unsaturated soils, "the
rate of infiltration will be greater than the permeability, but will soon decrease". It is assumed that
Lumb is referring to the saturated permeability, and it should also be noted that this was for a flat
surface, not a slope. The ability for infiltration to exceed the saturated permeability is presumed to
be a reflection of the water storage capacity of the soil.

Sun, Wong & Ho (1998) discuss this issue and show more clearly how the influence of the
diffusion element decreases with time. The rate of infiltration through a soil therefore tends to the
fully saturated permeability of the soil. From Sun et al's assessment of Lumb (1962a), the
infiltration process is controlled by time, the fully saturated permeability, the porosity of the soil
and the change in the degree of saturation.

Both Lumb and Sun et al appeared to be interested primarily in the development and advance of a
wetting front, across which there is a marked change in degree of saturation. The wetting front as
proposed by Lumb (1962a) is illustrated in Figure 3.1.

Chapter 3 40
Under infiltration conditions, the soil becomes fully saturated over a very small depth from the
surface, then the degree of saturation drops to a ‘wet value’ of 80 to 90%. This degree of
saturation is maintained until the ‘wetting front’ is reached, at which, the degree of saturation
drops sharply to the original value.

Bodman and Colman (1943), quoted both by Bear (1972), and Youngs (1957), present a similar
form ( see Figure 3.2 ), with a saturated zone extending to approximately 1.5cm below ground
surface. Below this is the transmission zone, in which moisture content is approximately constant,
with a degree of saturation of about 80%. This is followed by the wetting zone, where moisture
content drops, then the wetting front, which “represents the visible limit of moisture penetration
into the soil”.

Youngs (1957) undertook a series of vertical infiltration tests, using either slate dust ( particle size
0.04mm to 0.125mm ), or grade 15 Ballotini ( glass beads, less than 0.1mm in diameter ). He
found that with these materials, there was a uniform degree of saturation within the ‘wet zone’,
with no reduction in saturation over the transmission zone. He suppositioned that this was due to
little air becoming entrapped during the wetting process.

This implies that the exact form of the wetting front profile is partly a function of the grading of
the soil. Referring back to section 2.2.2, it is clear that a uniformly graded soil, tending to have a
uniform pore size, is less likely to entrap air on wetting up than a more varied soil.

Sun et al and Lumb seem less obviously concerned with the implications that the wetting front has
for pore water pressures or suctions beyond whether the soil has become fully saturated or not.
Little attempt is made to relate the wetting front to the overall behaviour of the soil mass, other
than that resulting from the application of saturated pore pressures.

Lumb (1962a) does link the advance of the wetting front to a loss of suction and hence a loss of
strength, through an expression for the shear strength of unsaturated residual soils in Hong Kong
( see Lumb 1962b ). True fully coupled behaviour is not considered.

Sun et al's 1998 work developed the wetting front approach to determine the long-term variation
in suction within a soil profile. They found that 'partially saturated' wetting bands, where the
advance of a wetting front results in the reduction of the magnitude of the suction, without fully
destroying it, were feasible and, depending on the soil properties, even more likely than a fully
saturated wetting front. This point will be considered further in section 3.8. For infiltration flow

Chapter 3 41
into unsaturated soils, they showed that the critical factors governing behaviour are the
permeability-suction relationship, the initial suction condition, and the duration and intensity of
the rainfall.

3.2. Antecedent rainfall

The effects of initial suction and rainfall duration / intensity are interrelated, and so are considered
together here. If the initial suction within the soil is a controlling factor, it would seem likely that
rainfall immediately prior to the storm event being considered would influence behaviour, since
such rainfall would presumably influence the suctions present in the soil.

Various authors ( e.g. Lumb 1975, Fourie 1996 ) have suggested that rainfall induced failures are
related to the duration and intensity of the antecedent rainfall ( up to 15 days prior to the failure
event ), in addition to the intensity of the ‘trigger’ storm. These findings have been contested by
Brand (1985), who pointed out that Lumb’s work was based on daily rainfall data at the Hong
Kong Royal Observatory, not from data collected local to the landslide locations. Brand used both
1 hour and 24 hour rainfall data, from rain gauges near to recorded land slides, and found that
antecedent rainfall had no significant effect on major landslide events in Hong Kong, although he
conceded that 3 to 4 day antecedent rainfall does seem significant to minor landslides. Brand's
data indicates that major events are due to short duration, high intensity rainfall events, with few
major failures if the hourly rainfall is less than 40mm per hour, and most casualty causing
landslides occurring when rainfall exceeded 70mm per hour.

The 24-hour rainfall data also showed reasonable correlation to frequency of slope failure, and
Brand suggested that this data might be a more useful indicator of impending failure. The
occurrence of a single, high, rainfall event ( i.e., the 1-hour rainfall rate ) is difficult to predict but
after several hours of rain, it is possible to estimate an approximate magnitude for the 24-hour
period.

Brand suggested that the 24-hour rainfall data is significant because a high 1-day rainfall will
normally be indicative of a period of high 1-hour rainfall. Hence, it is the high hourly rainfall that
is actually significant, and the 24-hour rainfall data is merely a marker reflecting this, rather than
being significant in its own right. Au (1998) gives data which supports this idea, stating that 15-
27% of the 24 hour rainfall occurred in 1 hour ( and 50-90% of the 24 hour rainfall occurred in 8
hours ) for a number of ‘cloudburst’ rainstorms.

Chapter 3 42
Where rainfall rates are lesser, Brand found that the antecedent effects built up over 3 or 4 days
may be enough to trigger minor, shallow, failures.

Ng and Shi (1998) carried out a numerical investigation of slope instability due to rainfall. They
found that a “single threshold rainfall intensity” did not provide a reliable warning of land
movement, and that a “critical rainfall duration” proceeding the failure, of between 3 and 7 days,
was also applicable to failure events.

Vaughan (1985) suggested that a decreasing permeability with depth may be partly responsible
for creating threshold event rainstorms which trigger failure in what had previously been
considered safe slopes. This idea does not preclude the possibility of antecedent rainfall also
influencing the slope behaviour, and in his case study from Fiji, Vaughan refers to failure
following “an unusually long period of exceptionally heavy rain” which lasted about 3 days.

Crozier and Eyles (1980) suggest that antecedent rainfall is a controlling factor in slope failure.
Their model generally requires an “antecedent excess rainfall” implying soil saturation must be at
least equal to the field capacity for failure, though an example is given where heavy rain triggered
a slope failure despite high negative pore water pressures. They hypothesised that this was due to
a high infiltration rate combined with low soil permeability, leading to perched groundwater, and
locally high pore water pressures.

Malone and Shelton (1981) looked at landslides in Hong Kong, and concluded that antecedent
rainfall is a relevant factor, in as much as it can affect prevailing soil moisture conditions, and
groundwater recharge. Their work was based on that of Lumb (1975) and Crozier and Eyles
(1980). Kay and Chen (1995) found that both 'hourly' and 'daily' rainfall were significant when
attempting to predict failures of slopes in Hong Kong.

Dai and Lee (2001) looked at rainfall and landslide occurrence in Hong Kong, and found that
‘significant’ landslides ( of 4 cubic metres or greater failure volume ) were best predicted by
reference to the 12 hour cumulative rainfall; however, for ‘large’ landslides ( of 30 cubic metres
or greater failure volume ), the 24 hour cumulative rainfall produced far better predictions. They
make the important point, however, that the occurrence of a landslide is influenced by many
factors other than rainfall, and hence each failure is case specific.

Ayalew (1999) supports this last point. While he found some evidence of a relation between
failure and the ratio of precipitation prior to failure relative to mean annual rainfall, he stated that

Chapter 3 43
“it is well known that rainfall affects the stability of sub-surface materials, but in view of the other
factors that can also influence the mechanisms of landslide generation, the importance of rainfall
is difficult to determine”.

Despite this, Ayalew noted that in Ethiopia, more landslides generally occur in September than in
July, despite July being wetter, while as many failures occur in October as in June, despite “the
amount of rainfall in October is almost negligible”.

This appears to indicate that there may be some time delay between a rainfall event and its effect
on the stability of the ground. Whether this is due to antecedent rainfall leading to a build up of
moisture within the soil, or whether it is a reflection of the movement of water, in some form of
wetting front, is not clear. It is questions such as this that provoked this research project.

Finlay et al (1997) looked at rainfall data for a whole range of time periods, and found that for
Hong Kong, the rainfall over a 1 to 12 hour duration was important for prediction of the number
of failures, with antecedent rainfall also having some influence. They found that the 1 hour
rainfall data gave the best indication of the ‘threshold rainfall’ to cause failure of any ‘single
period of time’ data, but that including the 15 minute, 24 hour, and 30 day rainfall data made their
predictions more accurate.

To determine the number of landslides, the best prediction was achieved using 3-hour rainfall
data, but the estimate was improved if 15-minute through to 30 day data was included. As above,
they found that there is a large degree of uncertainty in predicting landslide failures, due to the
numerous other variables that impact on slope stability, such as slope geometry or the geology.

Dissanayake et al (1999) looked at land sliding in Sri Lanka, and found that there was good
correlation between weekly cumulative movements and one-week cumulative rainfall, particularly
as compared to 2 or 3-week cumulative rainfall. This was, however, for a re-activated slide.

Fredlund and Barbour (1992) presented a model where the suctions developed were dependent on
the initial steady state conditions, and the intensity and duration of the storm event. In this model,
a ‘low intensity’ rainfall was applied over a period of days. This case clearly blurs the distinction
between the trigger event and antecedent rainfall. It raises the question that, if rainfall is
continuous but variable over several days, where is the cut-off between the start of the trigger
event and antecedent rainfall? The Fredlund and Barbour (1992) model also brings in the idea of

Chapter 3 44
‘steady state conditions’. If changes to the soil suctions are to be modelled, it is necessary to first
determine the initial suctions and soil moisture contents.

Fredlund and Rahardjo (1993b) presented an illustrative figure showing how the pore water
pressure / suction profile can vary with depth, reproduced here as Figure 3.3. Given a no-flow
surface boundary, the suction profile will tend in the long term to hydrostatic, in equilibrium with
the position of the phreatic surface. However, the effects of rain, evaporation and transpiration
make no-flow boundaries untypical of real world field situations.

Clearly, the initial pore water pressure / suction conditions selected for an analysis must be
determined on a case by case basis.

3.3 Water flow in an unsaturated soil

The flow of water within a fully saturated soil is normally taken to behave in accordance with
Darcy’s law, which for one-dimensional flow has the form:

∂h
vx = − kx
∂x
Eqn 3.1

Where

vx : Flow ( velocity ) of water in x direction

kx: Coefficient of permeability ( for water flow ) in the x direction


∂h : hydraulic gradient in the x direction
∂x

It is generally accepted ( for example Richards 1967, Childs 1969, Freeze and Cherry 1979,
Koorevaar et al 1983, Ng and Shi 1998 ) that Darcy’s law, and its associated assumptions, remain
applicable to flow through unsaturated soil, except that it now takes the form:

∂h
vx = − kx(ψ )
∂x
Eqn 3.2

Chapter 3 45
Where kx(ψ) is now the coefficient of permeability, as a function of suction. The issue of suction-
dependant permeability is discussed further in section 3.4.

Freeze and Cherry (1979) developed an equation for continuity of flow for transient flow through
an unsaturated soil, incorporating Darcy’s law in its unsaturated form, as shown in Eqn 3.3.

∂  ∂h  ∂  ∂h  ∂  ∂h  ∂θ
 k (ψ )  +  k (ψ )  +  k (ψ )  =
∂x  ∂x  ∂y  ∂y  ∂z  ∂z  ∂t
Eqn 3.3

Where θ represents the volumetric moisture content ( equal to the volume of water divided by the
total volume of the soil unit ).

Equation 3.3 was derived based on the assumption that the water is incompressible, and that the
porosity of the soil is constant ( that is, the soil is rigid ). Neither of these assumptions is
necessarily true, but Freeze and Cherry argue that changes in either water density or soil porosity
are small compared to changes in the degree of saturation, and may therefore be ignored.

A similar equation is presented by Ng and Shi (1998), as shown in Eqn 3.4.

∂  ∂h  ∂  ∂h   ∂θw 
 kx  +  ky  + Q =   Eqn 3.4
∂x  ∂x  ∂y  ∂y   ∂t 

Equation 3.4 is fundamentally the same as equation 3.3 ( the different authors use slightly
different notation ), but it may be noted that Ng and Shi present this equation in its two-
dimensional form, and also that they introduce the additional term Q, to allow for an applied
boundary flux.

Potts and Zdravkovic (1999) present a form of the continuity equation for saturated soils, as
follows:

∂vx ∂vy ∂vz − ∂εv


+ + −Q=
∂x ∂y ∂z ∂t
Eqn 3.5

Substituting Darcy’s law ( equation 3.1 ) gives:

Chapter 3 46
∂  ∂h  ∂  ∂h  ∂  ∂h  ∂εv
 kx  +  ky  +  kz  + Q =
∂x  ∂x  ∂y  ∂y  ∂z  ∂z  ∂t Eqn 3.6

The right hand side of equations 3.5 and 3.6 represents the volumetric strain in the element across
which flow is being measured, and hence, unlike Freeze and Cherry, allows for changes in the soil
porosity ( non-rigid behaviour ). In a rigid soil, this term would go to zero.

Thus it is clear that the fundamental differences in the continuity equation between the saturated
form and the unsaturated form are the change to a suction dependant permeability, and the
addition of a term allowing for changes in the volumetric water content over time.

It is clearly logical that an element of soil within the unsaturated zone can wet up or dry further,
and therefore may be subject to variation in its moisture content ( and hence its degree of
saturation ).

Freeze and Cherry re-wrote equation 3.3 to give equation 3.7, below, in which form it is known as
the Richard’s Equation.

∂  ∂ψ  ∂  ∂ψ  ∂   ∂ψ   ∂ψ

∂x 
k (ψ )  + 
∂x  ∂y 
k ( ψ )  +  k (ψ ) 
∂y  ∂z   ∂z
+ 1  = C
 ∂t
 Eqn 3.7

Where ψ = pore water pressure ( = suction, assuming zero air pressure )

And C = Specific moisture capacity, such that:

∂θ
C=
∂ψ
Eqn 3.8

The right hand side in either equation 3.3 or 3.7 represents the change in water storage within the
soil, and this behaviour is reflected by the slope of the SWCC. The SWCC shows how the water
content ( or degree of saturation, if the curve is in that form ) varies with varying suction. The
slope of the curve indicates the quantity of water that flows in or out of a soil element in response
to a given change in suction ( for example, Lam et al, 1987, Ng and Shi, 1998 ).

Chapter 3 47
In a fully saturated soil, the degree of saturation is constant, and the volumetric water content is
simply equal to the porosity of the soil. Thus, in a rigid element, the volumetric water content is
constant, while in a consolidating analysis, the volumetric water content changes as the soil
porosity does ( assuming solid, incompressible, soil particles ).

It is evident that the saturated continuity of flow equation is really just a specific case of the
unsaturated version of the equation, and that there is no fundamental difference in this aspect of
the flow behaviour.

3.4 The permeability – suction relationship

It has been well established ( for example, Bouwer, 1964, Freeze and Cherry 1979, Ng and Shi,
1998 ) that the coefficient of permeability within the unsaturated region is a function of the pore
water pressure ( i.e. a function of the matric suction ).

As was shown in chapter 2, the desaturation of soil and development of matric suction is
associated with voids within the soil becoming air filled, rather than water filled.

The flow of water in its liquid form through soil occurs through the water filled voids only, so as
the number of air filled voids increase, the available flow path for water flow decreases. Hence,
the effective permeability of the soil to water flow decreases. Bouwer (1964) found that the more
uniform the pore sizes of the soil, the more abrupt is the reduction in permeability with increasing
suction.

Once the water phase within the soil becomes discontinuous, water will no longer be able to flow
in its liquid form; moisture movement will now only be possible in its vapour phase, which occurs
at a much-reduced rate compared to liquid phase flow ( Childs, 1969 ). For this reason vapour
flow is generally ignored while fluid flow is occurring.

Irmay (1954) presented data which supports this, showing that the relative permeability of the soil
to water flow ( that is, the actual permeability divided by the maximum, fully saturated
permeability ) was “practically negligible” below some threshold degree of saturation ( given as
10 to 20% for the data presented ). Irmay further confirmed that flow of this remaining water is
through the vapour phase, requiring an “evaporation process”.

Chapter 3 48
While the general effect of suction on the permeability of a soil is therefore clear, the difficulty
lies in determining the precise relationship between matric suction and water permeability.

Numerous expressions for this relationship have been published, many empirically derived. Such
relationships may give the permeability as a function of suction, a function of degree of
saturation, or a function of a SWCC equation.

For example, Corey (1957) looked at the relation between saturation and permeability ( for both
air and water ) as saturation changed through drainage, in a very sandy sample.

He found that permeability to water flow dropped sharply once air entered the sample, and
became very small at a degree of saturation somewhat greater than zero. This saturation being the
residual degree of saturation, the point at which the water phase becomes discontinuous.

Corey defined the effective saturation, Se, as:

Se = Sw – Sr
1.0 - Sr

Eqn 3.9

Where: Sw = the actual degree of saturation

Sr = the residual degree of saturation

Corey then proposed an expression for the relative permeability of the soil to water flow using the
effective saturation, this simply being Se raised to the power 4.0. Brooks and Corey (1966)
reviewed this work, and suggested that for a completely uniform pore-size distribution, the
exponent would drop to 3.0. They also reviewed a number of similar expressions developed by
others, including Irmay (1954), who found that the exponent should be 3.0. Brooks and Corey
suggest that the value of the exponent typically lies between 3.0 and 4.0 inclusive.

Brooks and Corey also found that for a medium containing two immiscible fluids ( for example
air and water ), the permeability of the wetting phase is a function of the difference in pressure
between the two phases, the relative permeabilities for air and water flow being given by the
‘Burdine’ equations, which are also functions of effective saturation. While these expressions
were derived mathematically, they are based on assumptions that the soil is isotropic, and can be
modelled as a series of small diameter tubes. Brooks and Corey state that for a more general,
anisotropic situation, the expression put forward by Corey (1957) should be applied.

Chapter 3 49
Brooks and Corey (1964) discuss this expression, and state that “The approximations given by
Corey imply that effective saturation, se, is a linear function of 1/Pc2” ( Pc being the capillary
pressure ). They go on to state that this is not generally true, as has been proven experimentally,
but that the use of Se to give relative permeability is still a reasonable approximation. Hence,
while “permeability is a very sensitive function of capillary pressure”, they found that “the
relative permeability as a function of Se is not very sensitive to changes in the actual slope of the
capillary pressure curve.”

Fredlund, Xing and Huang (1994) reviewed the various published permeability functions for
unsaturated soils, before presenting theory based on a statistical pore size variation which is then
used to develop an unsaturated soil permeability function based on an SWCC. This work assumes
that volume change of the soil structure is negligible, and hence is only applicable for ‘rigid’,
rather than ‘fully coupled’ situations.

Fredlund (1998) further discussed the Fredlund, Xing and Huang permeability function. He
showed that this function extended from zero suction (100% saturation) to 1000000kPa suction
(0% saturation), but qualified this, however, by stating that at saturation below the residual state,
the moisture movement will tend to be vapour flow. From this, he suggests that it might therefore
be more appropriate to set a constant, minimum permeability at the residual saturation. Further
increases in the magnitude of the suction would not result in any further decrease in the
permeability. This is consistent with the ideas presented by Irmay (1954) and Childs (1969), as
given earlier.

The Fredlund approach of providing a very low, but non-zero, permeability for saturations at
residual or below is probably the preferable solution to dealing with water permeability at very
low suctions.

Setting permeability to zero at suctions below the residual degree of saturation would prevent any
flow within this region of soil, and hence would make further changes in degree of saturation
( and thus suction ) impossible unless changes are induced in the soil structure, for example by an
imposed load. Thus, in such a situation, to reproduce the real behaviour of a soil it would be
necessary to model the vapour phase transport of moisture.

Vapour phase flow adds an additional layer of complexity to an already complex issue, with flow
due to air flow, and molecular diffusion as described by Fick’s Law ( Childs 1969 ). This latter
process is highly temperature sensitive ( Gardner 1961, Bear 1972 ). Therefore, to accurately

Chapter 3 50
model this behaviour, it would be necessary to include the air phase as a discrete third phase
( which issue was discussed in section 2.2.3 ), and introduce temperature as an additional variable
throughout the half-space of the analytical problem.

While there are certainly situations in which such detailed assessment would be justified
( typically some form of environmental geotechnical analysis ), for a more general engineering
problem such detail adds considerably to the resources required to solve the problem for little
obvious gain. It also increases the amount / range of data needed to formulate the solvable
‘problem’ from the original engineering issue: for example, if temperature variation affects the
modelled behaviour, then it is necessary to know the initial temperature distribution within the
volume of soil being modelled.

Thus, the Fredlund approach of having a very low cut-off to permeability at the residual degree of
saturation seems to offer a practical engineering approach to dealing with this issue. Typical field
suctions in residual soils appear to be less than 100 kPa ( for example, Sweeney 1982, Abdullah
& Ali 1993, Au 1998, Sun et al 1998, Tsaparas et al 2002 ), although the action of vegetation may
lead to higher suctions developing. For most soils, the suction at residual saturation appears to be
considerably higher than these values ( for example Koorevaar et al 1983, Dineen 1997,
Cunningham, 2000 ), though Koorevaar et al do present data which shows that the suction at
residual saturation for sand can be in the region of 100 kPa.

Hence the permeability of soil at suctions greater than that needed to obtain residual saturation is
likely to rarely be of relevance to an analysis. Even when such permeabilities are applicable, the
use of a very low fluid-flow permeability rather than a diffusion flow seems unlikely to be of
sufficient significance to justify the increased complexity that is required to include diffusion
processes into a general unsaturated flow model.

3.5 Suction and soil deformation relationship

In discussing the effects of suction on permeability, it has generally been assumed that changes in
suction have no effect on the soil structure and hence while the volume of water within the soil
varies, the volume of voids remains constant ( that is, the soil is rigid ). However, this is not
actually the case. Changes to the suction within the soil mass will affect the soil structure.

Chapter 3 51
This aspect of unsaturated behaviour has not been widely investigated. With much of the work
published on unsaturated soils being directed towards agronomy, rather than engineering,
deformation behaviour of the soil is rarely considered.

The suction-deformation relationship is, however, an integral part of unsaturated soil behaviour,
so must be considered in any full-coupled approach to this subject. Such literature as is available
on this aspect of unsaturated behaviour forms the basis for much of the development work
undertaken as part of this project, so is reviewed in detail, later in this thesis, see Chapter 5.

3.6 Stress-state variables for unsaturated soil

In examining the fully coupled behaviour of unsaturated soils, it is necessary to consider the
stress-strain-strength behaviour of the soil, and what are the appropriate stress state variables.

In ‘traditional’ saturated soil mechanics, the principle of effective stress is given to hold:

σ = σ’ + u Eqn. 3.10

Where σ is the total stress, σ’ is the effective stress, and u is the pore water pressure. The
mechanical behaviour of the ( saturated ) soil is governed by the effective stress.

In an unsaturated soil, the three-phase nature of the soil makes the situation more complex. A
number of modified effective stress laws for unsaturated soil have been proposed at various times,
see for examples Jennings (1961). Fredlund and Rahardjo (1993) also considered in some detail
the various effective stress expressions that have been proposed for unsaturated soils.

Three stress-state variables are presented, of which any two will adequately describe an
unsaturated soil's mechanical behaviour. The three stress-state variables are (σ - ua), (ua – uw) and
(σ - uw), where σ is the total stress, u is the pore fluid pressure, with the subscript ‘a’ indicating
the air phase, and subscript ‘w’ for the water phase.

Throughout the literature on unsaturated soils, the preferred combination of stress state variables
appears to be (σ - ua) and (ua – uw), the matric suction. If these are combined with the assumption
of zero air pressure ( relative to atmosphere ), the two stress-state variables reduce to σ, the total

Chapter 3 52
stress, and – uw, the negative of the pore water pressure ( which is equivalent to the positive
suction ).

While more advanced stress-strain-strength models exist or are in development for unsaturated
soils, this aspect of unsaturated soil behaviour is outside the scope of this thesis. However, to
develop a realistic coupled unsaturated flow model, some attempt must be made to model the
unsaturated mechanical behaviour.

3.7 Suction-shear strength relationship

The reason for undertaking this research project is that saturated soil mechanics approaches are
not adequate for dealing with slope stability problems in unsaturated soil, since they fail to take
into account the effect of soil suction.

One aspect of the soil’s behaviour that suction impacts on is its shear strength: “ A loss of matrix
(sic) suction is followed by a loss of shear strength which can be responsible for shallow
landslides or initiation of debris flow with a failure surface within the unsaturated zone” ( Cojean
et al, 1994 ).

Clearly then, while this thesis is primarily concerned with the infiltration process, some
consideration must be given to the shear strength of the soil.

Generally, it seems to be accepted that suction within the soil contributes towards its shear
strength.

The standard Mohr Coulomb expression for shear strength in a saturated soil is as shown in
equation 3.11.

τ = c + σ ' tan φ
Eqn. 3.11

Where τ = shear strength, c = apparent cohesion, σ’ = normal effective stress, φ = angle of


shearing resistance

To allow for unsaturated soil suction, equation 3.11 has been modified by a number of authors.

Lumb (1962a) suggested:

Chapter 3 53
τ = cs + σ tan φ
Eqn. 3.12

Where τ = shear strength, cs = apparent cohesion, and σ = Total normal stress.

In equation 3.12, the apparent cohesion, cs, is a function of degree of saturation and voids ratio
( Lumb presented graphs for some typical soils ). This approach is undoubtedly simple, and is
consistent with the use of (σ - ua) and (ua – uw) as stress state variables combined with the zero air
pressure assumption.

Lumb presented a more detailed consideration ( Lumb, 1962b ), which showed that the
unsaturated apparent cohesion, cs, could be defined in terms of the saturated apparent cohesion
and the suction:

τ = c1 + (σ + us) tan φ
Eqn. 3.13

Hence:

cs = c1 + us tanφ
Eqn. 3.14

Where
c1 = the saturated apparent cohesion,
us = -u, the negative value of the pore water pressure
( for unsaturated soils, u is negative, hence us is positive ).

Fredlund et al (1978) presented two forms of a shear strength equation based on the Mohr-
Coulomb form.

The first makes use of (σ - uw) and (ua – uw) as the stress state variables, and takes the form shown
in equation 3.15.

τ = c'+ (σ − uw ) tan φ '+ ( ua − uw) tan φ ''


Eqn. 3.15

However, use of equation 3.15 means that any change in the pore water pressure changes both of
the stress state variables. Therefore, Fredlund et al proposed the alternate form, shown as equation
3.16.

Chapter 3 54
τ = c' '+ (σ − ua ) tan φ a + ( ua − uw) tan φ b
Eqn. 3.16

Where
c’’ = cohesion intercept when the two stress state variables are zero,
φa = Friction angle with respect to changes in (σ - ua) when (ua – uw) is held constant,
φb = Friction angle with respect to changes in (u a – uw) when (σ - ua) is held constant.

They show that equation 3.15 and equation 3.16 are related, such that c’’ ≡ c’ and φa ≡ φ’, giving:

τ = c'+ (σ − ua ) tan φ '+ ( ua − uw) tan φ b


Eqn. 3.17

and

tan φ ' = tan φ b − tan φ ' '


Eqn. 3.18

In equation 3.17, only 1 stress state variable responds to changes in pore water pressure, and only
φb is a ‘new’ parameter, beyond ‘standard’ soil properties, which must be determined.

Vanapalli et al (1996) suggest that at high degrees of saturation ( low values of matric suction )
the negative pore water pressure provides a direct increase in effective stress, thus increasing
shear strength in this manner. At lower degrees of saturation, the effect of suction on strength is
more variable, with strength possibly reducing, and must be explicitly allowed for.

They also found that the saturated angle of shearing resistance, φ’, can be assumed constant, at
least for the practical engineering suction range of 0-500 kPa.

They propose an equation for the shear strength of unsaturated soil derived from Fredlund et al’s
(1978) work, as shown below as equation 3.19:

[
τ = [c'+ (σn − ua ) tan φ '] + ( ua − uw ) (Θ K )( tan φ ') ]
Eqn. 3.19

Where:
Θ = the normalised volumetric water content = θ / θs
θ = Volumetric water content
θs = Volumetric water content at 100% saturation.

Chapter 3 55
K = a fitting parameter, used to match predictions to experimental results.

To avoid the need for a fitting parameter, Vanapalli et al proposed alternate variations for their
equation based on the SWCC, in terms of volumetric water content or degree of saturation:

  θ − θr  
τ = c'+ (σn − ua ) tan φ '+ ( ua − uw )  ( tan φ ')  
  θs − θr  
Eqn. 3.20

θr = Residual volumetric water content

  S − Sr  
τ = c'+ (σn − ua ) tan φ '+ ( ua − uw)  ( tan φ ')  
  100 − Sr  
Eqn. 3.21

S = Degree of saturation

Sr = Residual degree of saturation.

The residual values required by equations 3.20 and 3.21 may be determined from the SWCC,
although Vanapalli et al report that these values may be hard to determine for fine grained soils.

Fredlund and Rahardjo (1993) present data on the shear strength of unsaturated sands which
shows strength increasing with suction while the sands are still saturated ( i.e., still in the tension
saturated zone ), but which then show a decrease in strength before becoming constant with
suction. This conforms with Vanapalli et al (1996).

Fredlund and Rahardjo go on to discuss the shear strength equation, presenting the same equation
as shown here as equation 3.17. They note that the parameter φb ( the “angle indicating the rate of
increase in shear strength relative to the matric suction ( ua-uf )f” ) is always less than or equal to
φ’, the ‘saturated’ angle of friction. They present test data which shows φb = φ’ at low suctions,
with φb then steadily decreasing until a steady value, φb << φ’ is obtained.

They state that equation 3.17 reverts to the saturated shear strength equation ( equation 3.11 )
when suction is zero. This can easily be seen from equation 3.17, since the { (u a-uw)tanφ’ } term
will go to zero, and with the zero air pressure assumption, zero suction means u w equals zero
( hence σ = σ’ ).

Chapter 3 56
Both Fredlund and Barbour (1992) and Fredlund and Rahardjo (1993) show how matric suction,
net normal stress and shear strength interrelate to give a three dimensional failure surface for the
Mohr-Coulomb failure envelope, as shown in Figure 3.4.

This figure shows a planar failure surface, although this need not be the case; as stated above, φb
is not ( generally ) constant with suction. Fredlund and Rahardjo present some typical values for
c’, φb and φ’, for various soils worldwide.

In discussing the failure surface, Fredlund and Rahardjo show how the effect of matric suction
can be shown as an increase in cohesion, when shear strength is presented on a two-dimensional
Mohr-Coulomb plot. However, they emphasise that this does not indicate a ‘true’ cohesion effect
from suction, but that this is merely a presentational effect. They go on to discuss the effects of
stress paths and different types of strength testing, which is beyond the scope of this thesis.

Frydman (2000) reviewed work on the shear strength of Israeli soils, which showed that the angle
of friction, for these soils at least, tends to be fairly constant regardless of degree of saturation.
However, the cohesion of the soil was a direct function of saturation, showing a linear increase in
cohesion with a decrease in degree of saturation.

Frydman discussed the mineralogy of Israeli clays, and suggested that this may account for some
of the behaviour, which may render his conclusions inapplicable to other soils. It may also be
noted that he was working with two-dimensional plots, and did not present data on the three-
dimensional surface suggested by Fredlund and Rahardjo. The apparent linear increase in
cohesion seen by Frydman may well be a reflection of a constant ( for the range of suctions
covered by his test data ) φb angle.

3.8 Published work using numerical modelling to analyse the infiltration


process

Details have been published of a number of attempts that have been made to analyse the
infiltration process using numerical modelling techniques.

Gioda and Desideri (1988) looked at two-dimensional flow through an earth dam on an
impervious base, but with the assumption of a rigid soil ( “assuming…that the deformability of
the soil skeleton can be negated” ). Fundamentally, they approached the problem as a saturated

Chapter 3 57
soil problem, but they did provide for unsaturated flow, in that they incorporated a suction-
permeability relationship that allowed for near-to- but non-zero permeability when pore water
pressures were below atmospheric ( i.e. when PWPs had become suctions ). Their work is of little
direct relevance to this project, since it involves neither infiltration nor consolidating soil, but the
suction-permeability function used is worth noting.

Rubin and Steinhardt (1963) considered rain infiltrating into ( a presumably flat surfaced ) soil.
Their assumptions included that the soil was of “stable structure” ( i.e. rigid ), that the initial
moisture content distribution was uniform, and sufficiently low to give water permeability tending
to zero, and that the air phase was at atmospheric pressure and air was free to flow.

They proposed and then proved using a finite difference technique, that if the infiltration rate was
less than or equal to the fully saturated permeability, all infiltration would enter the soil, and there
would be no surface ponding. Further, if the infiltration rate was less than the fully saturated
permeability, while suctions would be reduced, the soil would remain unsaturated. In this case,
the suctions in the soil after infiltration would correspond to those needed to give a suction-
dependent permeability equal to the infiltration rate.

Only if the infiltration rate exceeded the fully saturated permeability of the soil would ponding
occur. Figure 3.5 shows the results of their analysis: clearly, the higher rainfall rate has led to an
increased moisture content in the soil, yet since neither rainfall rate exceeded the fully saturated
permeability, suctions remain, and the soil is still unsaturated.

Despite using a rigid soil, Rubin and Steinhardt’s results are logical and the general form of
behaviour shown is believable.

Rubin et al (1964) followed up Rubin and Steinhardt’s numerical analysis by undertaking


experimental work to confirm their earlier numerical predictions. Using a 5cm diameter test
column of various heights ( typically around 1m ), they experimentally determined the actual
moisture content profile of the column with time when subject to various rates of infiltration.
Their experimental data generally matches the numerical predictions, as shown in figure 3.6,
although for the higher infiltration rates, their numerical analysis tended to over-predict the
moisture content. Also, the numerical model generally gave a less sharp wetting front than was
observed experimentally, while the experimental data also indicated a shallow surface band of soil
tending towards full saturation, regardless of rainfall rate. Thus, the experimental data shows
close correlation with the wetting profile of Bodman and Cole, as shown in Figure 3.2.

Chapter 3 58
Chapuis et al (2001) considered a number of ‘standard’ problems, both saturated and saturated-
unsaturated, while assessing verification of numerical codes. They undertook their numerical
analysis using the Seep/W code. Within the code, they state that pore air pressure is atmospheric
( i.e. zero ), which is consistent with the discussion given earlier in section 2.2.3. However, they
also state that the total stress, σ, remains constant, and hence their work cannot be considered to
be fully coupled. The majority of their illustrative examples cover flow to wells or similar, and so
are not directly related to the problem under consideration here, but two significant points can be
noted from their work.

Firstly, they show how in an analysis of flow through a dam numerical convergence problems
may occur, caused by using excessively large elements. This resulted in the program incorrectly
determining permeability and volumetric water content ( as functions of suction ), leading to
excessive capillary flow in the region above the phreatic surface. Adjusting the mesh element size
gave better convergence, and more believable results.

This clearly reinforces the point that non-linear numerical analysis can be highly influenced by
the size / geometry of the elements used, and that care must be taken to investigate the sensitivity
of an analysis to the proportions of the mesh used to solve it.

Secondly, Chapuis et al did consider a ( non-coupled ) analysis featuring infiltration into a slope
with a fissured clay surface layer over less permeable clay. As modelled directly, very high pore
pressures are generated in the fissured clay layer as the infiltration is forced into this layer, but is
unable to drain into the less permeable clay beneath. This is presented as an illustration of a
weakness in a numerical code, causing an inaccurate result to be generated; in this case, the
failure of the code to allow for run-off results in incorrect pore water pressures in the near surface
material.

Chapuis et al suggest that this particular problem be dealt with by adding a surface, highly
permeable gravel layer ( which they justify the presence of by claiming it represents the surface
topsoil and vegetation ), which more readily allows lateral flow, thus preventing the build up of
pore pressures.

While this is a reasonable approach to reproducing a more realistic flow regime for this problem
when the code used is flawed and unable to accurately model the real behaviour, it is not the ideal
solution. It is preferable that the code should be formulated such that the real infiltration / run-off
behaviour is modelled.

Chapter 3 59
Tsaparas et al (2002) also used SEEP/W to analyse seepage in an unsaturated soil slope, then
determined the factor of safety by using the seepage results as input data into the computer
program Slope/W.

They investigated the effects of different intensities and distributions of rainfall, antecedent
rainfall, and the initial groundwater conditions ( depth to phreatic surface and pore water pressure
distribution ).

In undertaking their analyses, they found that a very fine element mesh was needed, with
elements 0.25m square on the slope surface. This was due to numerical difficulties in achieving
convergence in the analyses, due to the non-linearity of the problem.

Tsaparas et al found generally that the longer the period of rainfall lasted, the deeper the wetting
front tended to penetrate. Also, the saturated permeability was important, in that pore pressures
within the soil increased more ( became more compressive ), and at greater depths ( a greater
depth of wetting front ) for the soils with higher saturated permeability.

However, the ratio between the saturated permeability and the rainfall rate was also important. A
high rainfall relative to the saturated permeability produced only a shallow wetting front, since
most of the rain became run-off. From this, it can be seen that Tsaparas et al allowed for the slope
surfaces becoming fully saturated, and hence for run-off to develop. This was done by re-setting
the surface pore water pressure to 0 kPa if at any time this surface pressure became compressive
( more compressive than 0 kPa ). As is discussed in Chapter 4, this approach is flawed, in that it
neglects to ‘correct’ the pore pressures below the surface which are likely to be compressive if the
surface pressure became more compressive than 0 kPa.

Tsaparas et al also considered the effects of antecedent rainfall, and found in general that stop-
start rainfall had little long-term effect, particularly where the permeability was relatively high,
since the ground drained during the dry periods. However, prolonged antecedent rainfall would
significantly alter the pore water pressure conditions prior to the main rainfall event, such that the
initial pore water pressure distribution prior to the antecedent rainfall became all but irrelevant.

The effect of rainfall on the stability of the slope seemed to be related to the duration of the
rainfall event, with the longer the rainfall period, the lower the slope factor of safety. Antecedent
rainfall had some impact, especially when continuous, as previously noted, while lower

Chapter 3 60
( saturated ) permeability slopes tended to be more stable, since higher permeabilities permitted
deeper and more rapid penetration of the wetting front, leading to greater loss of suction.

Tsaparas et al make no reference to soil deformation, so it is not clear whether they used a rigid-
soil approach, but by determining the seepage pattern than looking at stability, it is clearly not
fully coupled.

Ng et al (2001) used the finite element program FEMWATER to undertake a three-dimensional


analysis of rainfall infiltration into a cut slope in Hong Kong that was prone to failure. While it is
not clear whether the program has a coupled capability, the analyses undertaken treated the
problem as rigid.

The analyses investigated the effects of various patterns and intensities of rainfall on the pore
water pressure distribution in the soil. While the implications of the analytical results on the
stability of the slope are discussed, no attempt is made to analyse the stability numerically, neither
as part of a coupled analysis including the water flow, nor as a separate analysis using the
determined pore pressure regime as input data.

In applying their precipitation data, Ng et al assumed that, on average, 60% of the actual rainfall
intensity infiltrates into the surface of the soil, with the rest taken as run-off ( a statistic applicable
to Hong Kong, and repeated by Tung et al, 1999 ). This is an attempt to reflect the surface
infiltration process as proposed by Mein and Larson (1973). Mein and Larson showed that in an
unsaturated soil, rainfall results in an initially high hydraulic gradient at the surface, which when
combined with the soil’s ability to increase in moisture content and store water, enables the
infiltration rate to initially exceed the fully saturated permeability.

However, as the surface soil becomes fully saturated, the infiltration rate drops to the fully
saturated permeability, and any ‘excess’ precipitation above this value becomes run-off.

Ng et al do not reproduce this effect directly, but rather simulate it indirectly with their ‘60% of
rainfall as infiltration’ approach. However, this approach seems flawed. Their results show that in
some sections of their mesh, under certain precipitation rates, suctions are completely destroyed
and the soil becomes fully saturated at the surface, but elsewhere and / or for other precipitation
patterns, surface suctions remain, and the soil is unsaturated.

Thus, while their approach might give approximately the correct infiltration into the mesh on
average, the actual pattern of inflow does not reproduce the real world situation. The use of a

Chapter 3 61
percentage reduction to the actual precipitation figure to give a surface boundary infiltration rate
seems reasonable if it is necessary to allow for canopy run-off, or any other form of above surface
interception. It seems inappropriate to pre-determine what percentage of the rain reaching the soil
surface actually penetrates into the ground, and in this respect is no better than Chapuis et al’s
approach of inserting a false gravel layer on the surface of the mesh to allow for lateral run-off.
Rather, the actual infiltration should be determined by what the soil ‘wants’ to absorb; this
requires that the code used can calculate this. An approach that enables this to be achieved was
used in this research project, as explained in Chapter 4.

Ng and Shi (1998) undertook a parametric study of rainfall into “a typical unsaturated hillside
with a steep cut slope in Hong Kong”, the slope being colluvium or completely decomposed
granite ( no distinction being made between the two ).

Their study was done using the finite element programme SEEP/W, with the rainfall modelled
through the application of a specified infiltration rate ( equal to the rainfall data ) on the boundary
surface, to determine the pore water pressure distribution. These pore pressures were then used in
a conventional limit equilibrium analysis to calculate factors of safety.

The results of Ng and Shi’s analysis indicate that rainfall causes a rise in the phreatic surface, and
a reduction in the suctions above the phreatic surface. They found that the stability of the slope
was affected by duration and intensity of rainfall. They also noted that the initial hydraulic
boundary condition was significant to stability, thus firmly implying that antecedent rainfall
directly affects slope stability.

It may be noted in the results that Ng and Shi present that pore water pressure becomes greater
than zero ( i.e. compressive pore water pressures ) at some locations on the surface of the mesh in
some analysis, implying that ponding of surface water is occurring. In the ‘cut’ part of the slope,
which has a horizontal surface, this is perhaps reasonable, but some of their results, for example
those given for anisotropic soils, show non-zero, non-suction pore pressures on sloping surfaces.
Such results seem illogical, and are probably the result of the chosen infiltration boundary
condition, which appears to be simply a specified infiltration rate, regardless of the resulting
surface pore water pressure.

Thus while this work provides a useful guide to some aspects of infiltration behaviour, it is
neither fully-coupled, nor does it appear that the infiltration process is being correctly modelled.

Chapter 3 62
Lam et al (1987) presented an approach to modelling “saturated-unsaturated soil systems” using a
numerical technique ( the finite element computer program ‘TRASEE’ ), and included three
example problems, namely flow through a dam after sudden impounding, vertical flow through an
unsaturated soil from a lagoon to the underlying groundwater, and rain-induced infiltration into a
slope in a multi-layered soil ( in the form of a 1m high experimental model ).

No attempt is made to model the strength-strain-stress behaviour of the soil, and while a ‘water
coefficient of volume change’ is incorporated within their analysis, this represents the change of
water storage with suction, and does not reflect consolidation. Moreover, a fixed value is
specified, implying a constant slope to the suction-volumetric water content relationship, which is
clearly not accurate.

Lam et al demonstrate that the phreatic surface is not a flow line, with their first, dam, example
showing flow across the phreatic surface. This example also shows the importance of flow within
the unsaturated region, with a calculated total flow through the dam being 30% greater than that
calculated using a traditional flow net, with flow limited to the saturated zone.

However, despite this, and their ability to reproduce the pore pressure in the slope model
relatively accurately, the Lam et al approach is fundamentally a ‘rigid soil’ approach, and does
not, therefore, accurately portray the behaviour of real soil.

Kasim et al (1998) undertook a study of the effects of steady rainfall on the long term ( steady
state ) suction within the ground, using the SEEP/W computer program.

They looked at both horizontal ground and slopes of three different gradients, and varied the
SWCC and infiltration rate used. However, no allowance for soil displacement was made, with
the analysis being ‘rigid’.

They found that suctions within the soil would not be fully destroyed unless the infiltration rate
matches the saturated soil permeability. Where the infiltration rate is less than the fully saturated
permeability, the steady state suction profile is a function of the infiltration rate, with lower
infiltration resulting in a higher magnitude of suction remaining in the soil.

Further, they showed that increasing the magnitude of the AEV tended to reduce the effect of
infiltration, see Figure 3.7. Two analyses were conducted of a soil column with a horizontal
ground surface. Saturated permeability was the same in both cases, but there was an order of
magnitude difference in the AEVs ( 10 kPa versus 100 kPa ). When an infiltration rate equal to

Chapter 3 63
the fully saturated permeability rate was placed on the soil surface, both examples showed a
complete loss of all suction. However, for lower infiltration rates, the high AEV soil retained
higher ( steady state ) suction at any given depth, for any given inflow rate, than the low AEV
soil.

They suggest that where the AEV is low ( relatively ), the steady state suction is approximately
constant over a finite depth from the surface. However, their analyses were all carried out using
an initial pore water pressure profile that was hydrostatic, with the ‘groundwater level’ ( phreatic
surface ) at 20m below the ground surface. It seems probable that the tendency for the suction
profile to remain approximately constant over some limited depth towards the ground surface
partly reflects this initial pore pressure profile. The same behaviour might well be observed with a
higher magnitude AEV if the initial PWP profile had the phreatic surface deeper.

This idea is supported by the form of the curves in Figure 3.7b, which match that of the ‘non-
constant’ portion of the curves in Figure 3.7a. For example, the ‘0.3 ks’ flux curve becomes
approximately constant at a suction about equal to the AEV in Figure 3.7a. Extrapolating the
same flux value curve in Figure 3.7b, it is apparent that this curve is also tending to become
constant at the AEV. However, because of the depth of the phreatic surface suctions equal to the
AEV are not present within the mesh, and so the curve cannot develop its ‘constant’ portion.

Kasim et al conclude that the AEV is of more influence than the soil desaturation rate parameter
( which is effectively the gradient of the SWCC ) in determining the steady state suction / PWP
profile, but that neither has any great influence on the rise in the phreatic surface that results from
the infiltration process. The response of the phreatic surface is almost totally dependent on the
magnitude of the infiltration relative to the fully saturated permeability.

It should be noted that the results and conclusions obtained by Kasim et al are for final steady
state conditions. It seems certain that during the transient phase which precedes steady-state,
when the water content of the soil is likely to be varying, the gradient of the SWCC, which
reflects the storage capacity of the soil, will be significant in determining the suction or PWP
profile within the soil.

Sun and Nishigaki (2000) used finite element analysis to look at the combined rain and
evaporation process, and how this generates tension cracks within the ground surface.

Chapter 3 64
Their analysis was of a 35° slope in granite and decomposed granite soil. Of particular note is that
their FE mesh uses very thin elements along the slope surface: approximately 0.25m thick. This
compares to the overall slope height of 10m, and a total mesh depth of 30m. The elements within
the mesh get thicker with depth, with the bottom most layer of elements being 3m thick. Sun and
Nishigaki have clearly concentrated elements in the area where they expected the most change in
the pore water pressure.

Sun and Nishagaki also show how incorporating evaporation in the ‘dry’ periods between rainfall
events can significantly alter the near surface PWP or suction, as compared to the PWP profile
that is generated by treating the ‘dry’ periods as ‘zero rain’. They then go on to compare how the
PWP / suction profile within the slope varies if tension cracks of various depths are introduced.

The effects of such cracks on the stability of the slope are considered, with the permeability of the
soil being multiplied by a factor based on the degree of volumetric shrinkage to determine an
increased permeability within the cracked zone, but the stability is not calculated from the FE
analysis. Rather, the numerical analysis assumes a rigid soil, and this analysis produces a PWP
distribution as an output, which is then incorporated into a limit equilibrium slip circle stability
analysis.

Their work suggests that if evaporation occurs between periods of rainfall, the effects can be
marked. However, it is not clear if their rainfall and evaporation data is based on actual field data.
Thus it is not certain how realistic it is to intersperse periods of evaporation between closely
spaced periods of rainfall.

While ICFEP has the capability to model the effects of evaporation and transpiration, the
approach taken in this research project has been to assume that evaporation occurs predominantly
in the dry season, and may therefore be neglected within the analysis undertaken using ICFEP.
Sun and Nishigaki’s work suggests that this may be a mistake, but in the absence of firm data
indicating that significant evaporation can commence the moment rain has stopped, this simpler
approach is justified.

The approach adopted for the ICFEP analyses is supported by Tsaparas et al (2002), who
attempted to model evaporation, but found their results unrealistic compared to field data. They
thus took the decision to exclude the evaporation process, justifying it on the grounds that no
evaporation occurs during a rainfall event, and it was the pore water pressure response to rainfall,
not evaporation, that they were investigating.

Chapter 3 65
Similarly, it is clear from Sun and Nishigaki’s work that the formation of tension cracks may lead
to significant changes in the slope behaviour. This is unsurprising, since the effects of tension
cracks are well known and generally allowed for in ‘traditional’ analysis of saturated soils. The
value of their work is primarily in giving guidance on the extent that such cracks form due to
evaporation. However, if the assumption is made that no evaporation process occurs during the
period covered by an analysis, then this aspect of their work is similarly non-applicable.

Karnawati (2000) used SEEP/W to model the infiltration process in a slope that failed in
Indonesia, with the slope stability analysis undertaken using GEOSTAR. While not explicitly
stated, this approach clearly implies that the pore water pressure response was determined for a
rigid soil, and then ‘plugged into’ the stability analysis.

Karnawati found that in her analysis, the phreatic surface rose in the lower and middle portion of
the slope, while dropping in the higher portion, despite a constant rainfall throughout the analysis.
She suggests that this was due to down-slope flow, primarily within a sand layer.

Presumably, without the rainfall, the drop in phreatic surface in the upper portion of the slope
would have been more pronounced, and the fact that the phreatic surface did drop while under
precipitation conditions suggests that her initial boundary conditions were not representative of a
‘steady state’. Instead, while she states that the specified PWP boundary conditions are typical of
the wet season, they appear to be a ‘snap-shot’ of the situation during transient conditions. This
emphasises the need to accurately determine the true boundary conditions of any problem prior to
analysing it.

It may be noted that Karnawati modelled rainfall as a specified infiltration ( “constant flux” ) on
the surface boundary throughout the analysis. Her results show that the phreatic surface almost
reached the soil surface. Had this occurred, it would no longer have been correct to impose all the
rainfall as infiltration, since some proportion of the rain would have flowed as surface run-off.
Karnawati does not show how she would have allowed for this, and it is probably that she would
have incorrectly retained the fixed-rate infiltration condition.

Fredlund and Barbour (1992) looked at the effects of rainfall on the suctions within a typical
Hong Kong slope, and how this affected the slope stability, using the computer program SEEP/W.

They modelled the rainfall as a specified infiltration flux on the surface boundary, but recognised
the possibility of the soil becoming fully saturated. However, rather than modifying their

Chapter 3 66
boundary condition to allow for run-off, they instead permitted ponding on the upper portion of
the slope.

The lower steeper portion of the slope was covered with chunam, a cement based slope protection
system, and they modelled this by restricting infiltration to 10% of the specified rainfall. This
made it improbable that compressive pore water pressures would be generated at the surface of
this part of the slope, and Fredlund and Barbour’s results confirm this to be the case. However, it
seems unlikely that ponding could occur on such a steep slope, so it is fortunate that this part of
the slope did not become fully saturated. While their results do not suffer because of it, Fredlund
and Barbour’s work does appear flawed by the absence of any capacity to determine run-off.

Fredlund and Barbour started their analysis with a hydrostatic pore water pressure / suction
distribution, then applied a light rainfall, based on the average annual Hong Kong rainfall, to
generate a steady state condition. This resulted in a relatively constant suction from the soil
surface down several metres, at which point the pressure distribution approximately rejoined the
hydrostatic pressure line, see Figure 3.8. This is comparable to the results obtained by Rubin and
Steinhardt (1963), as shown in Figure 3.5. Fredlund and Barbour’s pore water pressure profile is
from approximately mid-slope, just above the chunam covered portion of the slope.

It again shows how a given infiltration rate can lead to a reduction in the suction in a soil, until the
corresponding suction-dependent permeability matches the infiltration rate. At this point the flow
into the soil is able to pass through, without changing the water content of the soil, and a steady
state is achieved.

Fredlund and Barbour used this steady state situation as the starting pressure condition for two
transient analyses. The first was a reproduction of an intense 2-hour storm, while the second
modelled a prolonged 5-day event, with a lower rainfall rate.

In the first case, the specified inflow rate of 118mm/hr slightly exceeded the saturated
permeability of the soil, which resulted in surface ponding. A fully saturated band of soil
developed at the surface, within which the pore water pressure distribution approximately
approached hydrostatic. Since fluid flow is largely driven by pressure head differences, the PWP
distribution determined from this analysis is certainly influenced by Fredlund and Barbour’s
decision to allow surface ponding, which permits corresponding non-zero compressive pore
pressures to occur at the soil surface, and drive the infiltration process. The application of a run-

Chapter 3 67
off boundary condition with a maximum compressive pore pressure of zero at the surface would
logically result in a different pressure regime within the soil.

The effects of the intensive rainfall appear to be limited to a relatively thin band of soil, which
tends to move downward and disperse with time, once the rainfall event had ended.

The second case involved a five-day storm, with rainfall varying on a daily basis. While the
maximum rainfall specified of 7.2mm/hr was considerably less than case 1, the total rainfall for
the event amounted to 275mm, which exceeded that of case 1.

The effect of the longer but less intense storm was less extreme but more widespread. The
suctions within the soil were reduced throughout its depth, such that the general form of the
steady state profile was retained, but with lower magnitude suctions.

The results of the two different storm events modelled by Fredlund and Barbour clearly show that
the suction response to rain is at least in part dependent on the relative magnitude of the rainfall.

Fredlund and Barbour then took the pore water pressure / suction distributions generated by the
two transient analyses, and undertook a limit equilibrium analysis to investigate how the results
affected the factor of safety of the slope. Thus, while their work is of interest, it is still a non-
coupled approach. Also notably is that they incorporated into their work variation in the suction-
dependent shear strength of the soil, as discussed in section 3.7.

Ng and Pang (2000) looked at the dependence of the SWCC on the applied stress state, and then
considered the implications of this on slope stability.

They undertook a numerical analysis of a “typical cut slope” in Hong Kong, using SEEP/W to
determine the PWP/suction distribution. As they state, the SEEP/W analysis was uncoupled, with
soil deformations ignored. The results of their numerical analysis were then used in a limit-
equilibrium analysis to determine factors of safety.

To investigate the effects of stress on the SWCC, Ng and Pang undertook two sets of transient
analysis, with a short intense rain and a longer less intense rain in each, using a common multi-
layered slope mesh. The first assigned a single SWCC to all the soil layers, while the second
allocated different, stress-dependent, SWCCs to each layer. The transient analyses were
commenced from a steady state determined in a similar manner to that used by Fredlund and
Barbour (1992).

Chapter 3 68
Ng and Pang specified an infiltration flux equal to 60% of the actual rainfall data, with the other
40% being assumed to be run-off. It is assumed that this is not representing canopy intercept, but
rather is an attempt to reflect the average amount of rainfall that reaches the ground that will
penetrate into the ground. As such, it is clearly a less than satisfactory approach. During the early
stages of a storm event, when all of the rain should be absorbed into the ground, this approach
will result in insufficient flow entering the soil, while it does not prevent unlimited ponding to
develop later in the event, if the surface soil becomes fully saturated.

The effects of the different SWCCs were immediate, in that the steady state profiles varied. This
difference carried over into the transient analyses. Ng and Pang’s work also showed the same
general difference in pressure response between short intense rain and longer, less intense rain
that was shown by Fredlund and Barbour (1992). This occurred whether stress-dependent SWCCs
were used or not, so clearly both factors are important in determining the response of unsaturated
soil to rainfall.

Clearly, while Ng and Pang were investigating the effects of applied stress on the SWCC and how
this affected the soil’s response, the conclusions of the work could be extrapolated to cover multi-
layered soils. Ng and Pang effectively showing that a layered soil responds differently than a
homogeneous soil; an unsurprising conclusion. The important aspect of this work is that they
demonstrate stress-dependency in the SWCC ( through laboratory tests ), and then numerically
use this data to show how it may be necessary to treat an apparently homogeneous soil as multi-
layered.

The fact that the SWCC is affected by applied stress proves that it is not a unique property of the
soil. The SWCC reflects the water storage capacity of the soil, which is obviously a function of
the soil void’s ratio. The fact that the void’s ratio can change when subject to applied stress is
unsurprising, so the idea that the SWCC is affected by stress is quite logical. The nature of the
SWCC is considered further in Chapter 6.

Wong et al (1998) present one of the few attempts to develop a fully-coupled approach to the
behaviour of unsaturated soils. They make use of Biot (1941) and Dakshanamurthy et al (1984) to
generate coupled consolidation equations for unsaturated behaviour. These also form the basis for
the theory developed within this thesis ( see Chapter 5 ).

Chapter 3 69
Wong et al make use of the computer programs SEEP/W and SIGMA/W to undertake their
analyses, the first being used to undertake the seepage analyses while the second deals with stress-
deformation behaviour.

They present analyses of an unsaturated triaxial test, to investigate the effects of the gradient of
the SWCC, then go on to model a 2m high column of soil, with the phreatic surface at mid-height,
subjected to an applied surface load. Their results indicate that while applied loads cause a
significant change in the pore water pressure within saturated soil, the effects on the suctions
within the unsaturated soil are all but negligible, the most significant effect being the movement
of the phreatic surface. Conversely, vertical deformations within the unsaturated zone occur
rapidly in response to applied load, whereas in the saturated zone, they occur much more slowly,
as the consolidation process occurs.

Wong et al’s work included a number of simplifications, such as having a constant permeability,
not suction-dependent, and does not attempt to model any part of the infiltration process.
However, it is valuable as an attempt to reproduce coupled behaviour in unsaturated soil, and
provides a comparison against which the new coding within ICFEP can be compared.

De Campos et al (1992) undertook stability analysis of a slope in Brazil, using a computer code
developed at the Catholic University of Rio de Janeiro. They make the point that the information
required to undertake unsaturated analyses, specifically in their case the suction dependent
hydraulic conductivity relationship, is frequently scarce, requiring estimates to be made of such
properties.

Clearly, the accuracy of a numerical analysis is likely to suffer if parameters are estimated, and
this emphasises the need for numerical analysts to work closely with their laboratory-based
colleagues in determining what properties of the soil are needed and ensuring suitable tests are
developed to determine them.

Despite using their own software, de Campos et al still treat the stability problem in two parts,
with an initially rigid soil assumed in which the flow analysis is carried out, the results of which
are then utilised in the second, limit equilibrium analysis, to determine the factor of safety.

Kim (2000) more usefully attempts a fully coupled analysis of an unsaturated soil problem. He
looked at the variation in the position of the phreatic surface within a partly saturated / partly
unsaturated soil column which was subject to a surface load.

Chapter 3 70
His results show that there is an instantaneous rise in the position of the phreatic surface on
application of the load. Depending on the initial depth to the phreatic surface, its position may
continue to rise with time for a short period, before in all cases the position of the surface begins
to drop. The final depth of the phreatic surface lies above its initial position, but below the
minimum depth to which it rose.

Vertical displacements of the column surface also reflect the initial position of the phreatic
surface. The greater the initial depth to the surface, and hence the more of the column that starts
unsaturated, the more rapid is the development of the surface displacements.

Both of these two observations are consistent with the results of Wong et al’s (1998) work.

Kim states that “less final steady-state vertical displacement occurs in the partially saturated soil
columns…These effects of the unsaturated zone are more conspicuous as the initial water table is
located higher and the magnitude of the surface loading increases”.

This and the results he presents seem to suggest that if the column analysed is fully unsaturated,
vertical displacements are apparently equal to those of a fully saturated column, but as the
phreatic surface is raised within the soil, the calculated displacement deviates from the fully
saturated value. The biggest deviation between the calculated value for the saturated and partially
saturated column displacements occurs for the case with the phreatic surface at 2m below the
surface ( i.e., for the case where the phreatic surface is closest to the ground surface, without
actually being at the surface ). Thus the closer the column is to full saturation, the greater the
deviation of the calculated displacement from the fully saturated value.

Kim accounts for the displacement behaviour by saying “the unsaturated zone absorbs a portion
of the mechanical loading stress permanently”, but this does not seem to adequately explain the
behaviour he presents.

He also extends his work to a two-dimensional situation, and showed that the behaviour seen in
the one-dimensional column was also generally applicable to the two-dimensional case.

While Kim demonstrated a fully-coupled analysis, he did not look at the infiltration process, so
while his work is useful in developing the theory necessary to code unsaturated behaviour into
ICFEP, it does not address the central concern of this research project.

Chapter 3 71
Ng and Small (2000) use Biot (1941) to develop the fully coupled equations for unsaturated soil
analysis. However, they are primarily interested in soils that, while unsaturated, have a high
degree of saturation, such that the pore water pressure and the pore air pressure are very close in
value. They state that this is typically at degrees of saturation of 80% or higher.

This appears to lead them to make a number of assumptions, most notably that strain can be
assumed to be small, and hence that porosity of the soil remains approximately constant. Thus any
change in volumetric water content due to a change in suction manifests itself solely as a change
in the degree of saturation, with porosity unchanged. Hence, while the soil is not rigid and may
undergo normal consolidation through loading, it is effectively rigid in response to changes in
matric suction.

Ng and Small’s finite element formulation for unsaturated soil is thus basically the consolidating
full-saturated equations, with the flow matrix modified to allow for suction dependent
permeability, and with an added term to allow for the water storage capacity of the soil. As is
shown in Chapter 5, this does not fully allow for all aspects of unsaturated behaviour, and hence
their formulation is incomplete.

Ng and Small primarily seem to be interested in how ‘almost saturated’ soils respond to imposed
loading, and how this changes the saturation within the soil. They present results of several
analyses using their formulation, coded in FORTRAN, and compare them to laboratory test data,
and there is generally good correlation.

The one example analysis they present that comes close to the situation of infiltration into
unsaturated soils is of rapid draw down affecting a dam. However, in this example they admit that
their formulation for the unsaturated equations is only valid at high degrees of saturation. To
analyse the draw down problem, they are forced to treat the soil as rigid.

Ng and Small’s approach fails to provide the general, fully-coupled unsaturated soil capability
that is necessary to complete this research project. However, it did provide useful guidance in
aspects of the theory developed in Chapter 5.

Forsyth (1988) developed two methods for dealing with transient flow in unsaturated soils,
however neither was fully coupled. His work is notable, however, since while one method made
the common assumption of zero air pressure, the other allowed for the air phase as a separate
phase ( i.e. non-zero air pressures possible ).

Chapter 3 72
Forsyth showed that the zero air pressure assumption often gives much the same results as the full
three phase soil ( two-phase flow approach ), but that under certain circumstances significant
errors may be given if the air phase is not specifically modelled. Such errors occurred when the
air permeability was very low and not changing much, for example when airflow permeability is a
power function of air saturation, and air saturation is very low. They also tended to occur if the
zone of soil suction was small relative to the mesh element size used in the analysis: where the
zone of soil suction lay entirely within the thickness of one element, discrepancies between the
‘standard’ two phase and more detailed three phase approach occurred. If the element size was
adjusted such that the zone of suction was spread over several elements thickness, and hence
changes in the suction between integration points in the mesh were less sharp, both methods gave
comparable results. It is probable, however, that any analysis where the unsaturated zone lay
entirely within the thickness of one element would give results of dubious accuracy. Simple good
practice would suggest that a finer mesh would be required.

Forsyth’s work thus emphasises the need to select element size within a finite element mesh
carefully, when dealing with unsaturated soil behaviour.

The discrepancy between the two methods’ results resulting from low air permeability is less easy
to avoid without actually modelling airflow as a separate phase. However, as discussed in section
2.2.3, it seems reasonable to accept some inaccuracies in the results of the analysis, particularly if
this is limited to a narrow range of saturation when air permeability is low, in return for the
reduction in effort and complexity that can be achieved by maintaining a two-phase approach with
the assumption of zero air pressure.

Lloret and Alonso (1980) were one of the first to attempt to generate a model for unsaturated
behaviour that included consolidation behaviour.

They provided a general formulation for three-dimensional behaviour that allowed for air and
water as separate phases. However, the data for their analyses was determined from one-
dimensional tests, and their examples are limited to such.

They simulated saturated consolidation, infiltration leading to swelling in an unsaturated soil, and
collapse due to loading then wetting of an unsaturated soil, all one-dimensionally.

Their results show the importance of variable permeabilities in governing deformation in


unsaturated soil. They appear also to support the use of a three-phase approach, with the air phase

Chapter 3 73
explicitly modelled as a separate phase. However, since they did not repeat their analyses with a
two-phase model, it is not possible to ascertain the full significance of the three-phase approach.
Additionally, as discussed in Chapter 2, a one-dimensional analysis is not typical of a field
situation. Hence, a two-phase, zero air pressure assumption remains a credible approach to the
analysis of unsaturated soils.

Li and Zienkiewicz (1992) studied the problem of multi-phase flow in deforming porous media.
They looked at the case of two immiscible fluids, but also accepted the advantages of the zero air
pressure assumption, and presented a simplified single-phase flow approach for unsaturated soils.
However, the examples they present are for two-phase flow, dealing with water injection into an
oil formation.

Their theory is largely a development of Biot (1941), and, as with Ng and Small (2000), provided
useful guidance in the formulation of the unsaturated theory developed in this thesis ( see Chapter
5 ). However, they were clearly addressing a different issue to that covered by this research
project, with no reference being made to precipitation-infiltration or slope stability.

Alonso et al (1988) present a coupled approach for the stress-strain-flow behaviour of partly
saturated soils. While not explicitly stated as such, their work appears to be a progression from
Lloret and Alonso (1980), and again is a full 3-phase approach.

They use their formulation to look at pore water pressure response and deformation in an earth
dam under construction, with either a ‘dry soil’ fill, with an initial degree of saturation of 80%, or
a ‘wet soil’ fill, initially at 90% saturation.

Their results show that the wet soil responds to the construction of subsequent layers above, with
consolidation and an increase in the pore water pressure occurring. This behaviour is not observed
in the dry soil.

Alonso et al (1988) do not give an SWCC for the material used in their analyses, but rather
present state surfaces to show the behaviour of the soil. From this, the AEV appears to be equal to
0 kPa, with the saturation of the soil decreasing noticeably immediately the soil develops suctions.
However, the results of their analyses suggest that the wet soil responds to loading in a manner
reminiscent of saturated soil. In contrast, the dry soil shows markedly different behaviour
compared to a fully saturated soil. This supports the idea that there is some ‘cut-off’ degree of

Chapter 3 74
saturation, below 100%, at which point, the soil’s behaviour changes from a saturated, or at least
pseudo-saturated soil to unsaturated behaviour.

Alonso et al’s (1988) work is also interesting in that to obtain the fully coupled solution they
‘cheat’. Rather than solve the flow and stress-strain behaviour as a single combined equation,
these two aspects are solved by two separate computer programs, which are linked and work
iteratively in tandem to determine the coupled solution.

Additionally, the concept of state surfaces for unsaturated soils provides a useful insight into how
different aspects of an unsaturated soil’s behaviour are inter-related.

As with the other attempts at coupled analysis, Alonso et al (1988) do not address the problem of
rainfall infiltration and slope stability.

However, Alonso et al (1995) use the work of Alonso et al (1988) as the basis to undertake a two-
dimensional analysis of unsaturated flow due to rainfall in a slope, determining slope stability
through limit equilibrium. Thus they were able to show the evolution of the factor of safety with
time, and how various analysis input factors impacted on the safety factor.

Despite their previous work using a three-phase approach, Alonso et al (1995) make the zero air
pressure assumption for their slope stability analyses. They also distinguish between two different
rainfall boundary conditions. Either a specified pressure boundary condition, with suction equal
zero, is applied, representing a flooded surface with run-off, or a specified infiltration rate is
applied, where the infiltration rate equals the rainfall rate, and is less than the permeability of the
surface soil. However, these two conditions are either/or. Consequently, they do not have a
‘dynamic’ boundary condition that responds to the pore pressures at the boundary surface, and
adjusts the boundary condition to reflect the current pressure.

The results of Alonso et al’s (1995) analyses of a flooded-surface slope revealed the sensitivity to
soil permeability. The higher the permeability, the more rapid was the decline in the factor of
safety, which reflected the quicker penetration of the wetting front into a more permeable soil.

The nature of the SWCC was also shown to be important. If the SWCC shows a sudden drop in
saturation, the factor of safety variation with time would also show a sudden drop. A gentler
SWCC would result in the factor of safety dropping more slowly with time. This is illustrated in
Figure 3.9.

Chapter 3 75
The analyses carried out using a specified infiltration rate demonstrated that the factor of safety of
a slope can continue to decline even after the rainfall event has ended. This is a response to the
continued percolation of the wetting front into the soil, under gravity. This behaviour also reflects
the permeability of the soil, it being noted that the post-rainstorm reduction in Factor of Safety is
more prevalent in lower permeability soils, as the penetration of the wetting front takes longer.
However, the total reduction in factor of safety was greater in the higher permeability soils,
reflecting the greater ability of the water to penetrate into the soil.

Alonso et al (1995) also looked at infiltration into a layered overconsolidated clay slope, with the
‘flooded surface’ boundary condition. It is worth noting that for the ‘non-rain’ period at the end of
their analysis, they assumed negligible evaporation ( in contrast to Sun and Nishigaki (2000), as
described earlier ). Their analysis showed a reduction in safety factor and progressive lateral
displacement with time, particularly in the surface most layer of the soil. Due to differences in
permeability, a perched water table developed in this layer.

This form of lateral spreading, without a clear failure surface, is consistent with some unsaturated
failures, such as Pak Kong ( see section 3.9 ), and suggests that the development of perched water
tables may be highly significant in the behaviour of unsaturated soil slopes.

Neuman (1973) presented an early attempt to codify unsaturated behaviour and used his work to
look at transient seepage through an earth dam

While his work has been generally superseded by more recent publications, Neuman makes an
interesting observation regarding the flow pattern. If the dam is constructed of ‘equally wet’
material, there will be a uniform pore water pressure throughout, and hence a pore water total
head gradient will exist.

Thus, without an external flow being applied, flow will occur as the pore pressures within the
analysis mesh redistribute towards a hydrostatic profile. This flow will be occur regardless of the
imposed boundary conditions.

This again emphasises the need for care in determining initial conditions for a numerical analysis.

Forsyth et al (1995) developed a numerical model for unsaturated flow that included two-phase
flow ( i.e. a three phase – air:soil:water – model ). However, while it allowed for both the water
phase and soil particle compressibility, it dealt only with flow, and was not fully coupled. They

Chapter 3 76
also show how their expression can be developed into the special case of zero air pressure ( two
phase model / one phase flow ).

They were primarily interested in flow in very dry soils, specifically looking at waste cover
designs in arid regions, and they present a number of examples, mostly involving infiltration over
a limited portion of the surface boundary into unsaturated soil. Their examples are run using both
the two-phase flow and one-phase flow versions of their expression.

Their results confirm that generally a one-phase flow analysis gives similar results to the more
detailed two-phase flow approach. The exception was for high degree of saturations ( S = 90% or
higher ), when significant differences occurred, due to the trapping of the air phase. This is
consistent with Forsyth’s earlier 1988 work.

It should also be noted that Forsyth et al (1995) were demonstrating an alternative numerical
technique, and much of their work dealt with numerical aspects, such as required CPU time,
rather than the geotechnical processes at work.

Sun et al (1998) used the SEEP/W program to analyse one-dimensional columns of two different
soils, exposed to various rates of rainfall, and their results were discussed briefly in section 3.1.

A steady state seepage condition for each soil type was generated by applying a low-level
infiltration on the surface boundary. These steady states were then used as the initial pore water
pressure distributions for a series of transient analyses. The transient analyses used three different
rates of infiltration, each lasting for a different duration.

The results of the analyses showed that if the infiltration rate equals or exceeds the saturated
permeability of the soil, the suctions would be fully destroyed, and a 100% saturation wetting
front would penetrate into the soil. A lesser infiltration rate would still generate a wetting front.
However, this would be less than 100% saturation, and hence some amount of suction would
remain in the soil. If the suction left in the soil is compared to the suction-permeability
relationship, the corresponding permeability is found to be equal to the specified infiltration rate.

This is consistent with the work or many others, for example Rubin and Steinhardt (1963) and
Kasim et al (1998), as presented earlier.

Sun et al’s results show that the wetting front tends to penetrate down into the soil quicker for
soils with a higher saturated permeability. However, it is probable that the rate of penetration of

Chapter 3 77
the wetting front reflects the maximum possible infiltration rate: in the less permeable soil, the
maximum infiltration into the soil was set by the saturated permeability, not by the specified
infiltration rate. The difference in speed of penetration by the wetting front for the case where the
infiltration rate was less than the saturated permeability of both soils was less marked.

Sun et al further note that the wetting front will advance into the soil more quickly if the water
storage capacity of the soil is small ( that is, if the gradient of the SWCC is small ). They also
observed that if the supply of water to the soil surface exceeds the saturated permeability, surface
runoff must be generated. No slope stability analyses were undertaken.

3.9 The Pak Kong failure

Part of the ‘inspiration’ for this research project was the failure of a slope at Pak Kong, in Hong
Kong.

Following heavy rainfall in June 1992, a cut slope at the Pak Kong water treatment works
underwent movement, leading to damage at the toe of the slope and to properties in O Long
village at the top of the slope.

A number of limit equilibrium analyses were undertaken, all of which indicated a factor of safety
for the failed slope of greater than 1.0, when ‘realistic’ water pressures, based on the field data,
where used. Only by applying a considerable increase in the height of the phreatic surface,
leading to almost full saturation, could failure be simulated ( Watson Hawksley, 1993 ).

It was noted that the actual failure was more of a lateral spreading, and that it moved
intermittently, apparently in response to rainfall events, with the movement stopping of its own
accord. Hence it did not appear to be a clear cut, ‘rigid body’ slip moving along a well-defined
slip surface. This may account for the inability of the limit equilibrium analyses to correctly
predict failure, and because of this, a series of numerical analyses were undertaken.

The numerical analyses were undertaken using the Imperial College Finite Element Program,
ICFEP ( GCG, 1993 ). At the time, while tensile pore water pressures could be incorporated into
an analysis, the soil could only be treated as saturated.

Chapter 3 78
It was recognised that to accurately model the deformations of the slope it would be necessary to
model the initial condition as accurately as possible. To this end, the stresses in the slope were
determined by allowing them to form from the self-weight of the soil in the original natural slope
profile. The excavation to form the cut slope was then modelled.

The initial phreatic surface was determined from the field observations of water level in the slope.
The effects of precipitation, however, were not modelled directly. Rather, the phreatic surface was
raised by various amounts, to simulate the effects of rainfall.

The analyses were conducted in the form of a parametric study, and failure of the slope was only
achieved when the angle of friction of one of the soil layers was reduced to below that which was
likely, and the phreatic surface was raised to the slope surface.

In all the analyses, there was a tendency for lateral spreading to occur as the phreatic surface was
raised to the soil surface. However, this spreading was of lesser magnitude than was observed in
the real slope, was generally more deep seated, and tended to involve rigid body movement, rather
than the more dispersed strains implied by the field data from the real slope.

The ICFEP analyses of the Pak Kong slope emphasised the need for a better understanding of the
infiltration process into unsaturated soil, and through the unsuccessful attempt to reproduce the
field observed deformations, demonstrated why a fully coupled approach is required. In this
respect, it was the ‘ancestor’ of this research project.

Chapter 3 79
0 So Sf 1.0
So = Initial saturation
Sf = Final degree of
saturation

Depth
h

Figure 3.1: Lumb wetting front ( after Lumb, 1962a )

Ground surface Saturated zone

n Moisture content, c

Transition zone
Air dry soil

Transmission
zone
n = porosity
Depth

Saturation

Wetting zone

Wetting front

Z ’

Figure 3.2: Bodman and Cole wetting front ( after Bear, 1972 )

80
Steady state Steady state
evaporation infiltration
+ qwy - qwy
Cover

+ qwy Steady state flow


Steady state
downward Vadose
flow upwards
- qwy zone

Static equilibrium with water


table ( i.e. ‘hydrostatic’ ) ( Negative pore-
( qwy = 0 ) water pressures )
y

Water table ( Phreatic surface ) x


Datum

Figure 3.3: Variation in pore water pressure distribution with


depth ( after Fredlund and Rahardjo, 1993b )

81
,
φb

oni
ct
u su
Extended Mohr-Coulomb
φ’

a - ic
w)
failure envelope

(u atr
M
φb
Shear strength, τ

(ua - uw)f tanφb


c’

φ’

c’

0 Net normal stress, (σ - ua)

Figure 3.4: Shear strength failure surface for unsaturated soils


( after Fredlund and Rahardjo, 1993a )

82
Water content, cm3 / cm3
.00 .02 .04 .06 .08 .10 .12 .14 .16 .18 .20 .22 .24
0
5625 2400

10 9000 3600

12375 4800
20
15750 6000
X, cm

30
7200
19125
40 8400
22500

50

60

For Rehovot sand.


Numbers on curves indicate infiltration duration, in seconds.
The profile series on the right ( 2400 to 8400 seconds ) is for rain intensity of 0.001306
cm/sec.
The profile series on the left ( 5625 to 22500 seconds ) is for rain intensity of 0.0003528
cm/sec.

Figure 3.5: Soil moisture content profiles predicted for rain infiltration
( after Rubin and Steinhardt 1963 )

83
Water content cm3/cm3
.000 .060 .120 .180 .000 .060 .120 .180 .240 .300

12

18
Soil depth, cm

24

30

36

42

48

54
A B
60

Data for Rehovot sand.


A represents infiltration of 13.4 mm/hr
B represents infiltration of 47.0 mm/hr
Crosses represent experimental data from short infiltration durations
Circles represent experimental data from long infiltration durations
Continuous lines represent the theoretical curves.

Rehout sand porosity = 0.40; therefore at 100% saturation,


volumetric water content = 0.40

Figure 3.6: Comparison of theoretical and experimental


data ( after Rubin et al, 1964 )

84
20 3.7a:
AEV: 10 kPa
15 Surface flux
No flux
Elevation (m) 1 ks
10 0.3 ks
0.1 ks
5 0.01 ks
0.001 ks
0
-200 -150 -100 -50 0 50 100 150

Pore-water pressure (kPa)

20 3.7b:
AEV: 100 kPa
15
Surface flux
Elevation (m)

No flux
10 1 ks
0.3 ks
0.1 ks
5
0.01 ks
0.001 ks
0
-200 -150 -100 -50 0 50 100 150

Pore-water pressure (kPa)

ks for both 3.7a and 3.7b = 1E-05 m/s

Figure 3.7: Effect of AEV on suction profile under infiltration


( after Kasim et al, 1998 )

85
80

75

70

65

60

55
ELEVATION ( m )

50

45

40

35 Hydrostatic pressure
profile
Simulated steady
30
state case

25

20
-30 -20 -10 0 10 20
SUCTION ( meters of water )

Figure 3.8: Suction profile showing steady state simulation


( after Fredlund and Barbour 1992 )

86
106

Fig 3.7a: SWCC


Curve A
105

Suction (Pa)
104
Curve B

103

102
0 0.2 0.4 0.6 0.8 1.0
Degree of saturation

2.2
Fig 3.7b: FofS
2.0
Global safety factor

1.8 Curve B
Curve A
1.6

1.4

1.2
0 0.2 0.4 0.6 0.8 1.0
6
Time ( 10 seg )

Figure 3.9 Effect of the form of the SWCC


( after Alonso et al, 1995 )

87

You might also like