You are on page 1of 22

CHAPTER 7

7.1

Co
py
rig
hte
dM
ate
ria
l

Effective, Intergranular, and


Total Stress

INTRODUCTION

The compressibility, deformation, and strength properties of a soil mass depend on the effort required to
distort or displace particles or groups of particles relative to each other. In most engineering materials,
resistance to deformation is provided by internal
chemical and physicochemical forces of interaction
that bond the atoms, molecules, and particles together.
Although such forces also play a role in the behavior
of soils, the compression and strength properties depend primarily on the effects of gravity through self
weight and on the stresses applied to the soil mass.
The state of a given soil mass, as indicated, for example, by its water content, structure, density, or void
ratio, reects the inuences of stresses applied in the
past, and this further distinguishes soils from most
other engineering materials, which, for practical purposes, do not change density when loaded or unloaded.
Because of the stress dependencies of the state, a
given soil can exhibit a wide range of properties. Fortunately, however, the stresses, the state, and the properties are not independent, and the relationships
between stress and volume change, stress and stiffness,
and stress and strength can be expressed in terms of
denable soil parameters such as compressibility and
friction angle. In soils with properties that are inuenced signicantly by chemical and physicochemical
forces of interaction, other parameters such as cohesion may be needed.
Most problems involving volume change, deformation, and strength require separate consideration of the
stress that is carried by the grain assemblage and that
carried by the uid phases. This distinction is essential
because an assemblage of grains in contact can resist
both normal and shear stress, but the uid and gas

phases (usually water and air) can carry normal stress


but not shear stress. Furthermore, whenever the total
head in the uid phases within the soil mass differs
from that outside the soil mass, there will be uid ow
into or out of the soil mass until total head equality is
reached.
In this chapter, the relationships between stresses in
a soil mass are examined with particular reference to
stress carried by the assemblage of soil particles and
stress carried by the pore uid. Interparticle forces of
various types are examined, the nature of effective
stress is considered, and physicochemical effects on
pore pressure are analyzed.

7.2

PRINCIPLE OF EFFECTIVE STRESS

The principle of effective stress is the keystone of


modern soil mechanics. Development of this principle
was begun by Terzaghi about 1920 and extended for
several years (Skempton, 1960a). Historical accounts
of the development are described in Goodman (1999)
and de Boer (2000). A lucid statement of the principle
was given by Terzaghi (1936) at the First International
Conference on Soil Mechanics and Foundation Engineering. He wrote:
The stresses in any point of a section through a mass of
soil can be computed from the total principal stresses, 1,
2, 3, which act in this point. If the voids of the soil are
lled with water under a stress u, the total principal
stresses consist of two parts. One part, u, acts in the water
and in the solid in every direction with equal intensity. It
is called the neutral stress (or the pore water pressure).
The balance 1 1 u, 2 2 u, and 3 3
u represents an excess over the neutral stress u, and it has
its seat exclusively in the solid phase of the soil.

173

Copyright 2005 John Wiley & Sons

Retrieved from: www.knovel.com

174

EFFECTIVE, INTERGRANULAR, AND TOTAL STRESS

ticle forces in a soil mass. Interparticle forces at the


microscale can be separated into the following three
categories (Santamarina, 2003):
1. Skeletal Forces Due to External Loading These
forces are transmitted through particles from the
forces applied externally [e.g., foundation loading) (Fig. 7.1a)].
2. Particle Level Forces These include particle
weight force, buoyancy force when a particle is
submerged under uid, and hydrodynamic forces
or seepage forces due to pore uid moving
through the interconnected pore network (Fig.
7.1b).
3. Contact Level Forces These include electrical
forces, capillary forces when the soil becomes
unsaturated, and cementation-reactive forces (Fig.
7.1c).

Co
py
rig
hte
dM
ate
ria
l

This fraction of the total principal stresses will be called


the effective principal stresses . . . . A change in the neutral
stress u produces practically no volume change and has
practically no inuence on the stress conditions for failure
. . . . Porous materials (such as sand, clay, and concrete)
react to a change of u as if they were incompressible and
as if their internal friction were equal to zero. All the measurable effects of a change of stress, such as compression,
distortion and a change of shearing resistance are exclusively due to changes in the effective stresses 1, 2 and
3. Hence every investigation of the stability of a saturated
body of soil requires the knowledge of both the total and
the neutral stresses.

In simplest terms, the principle of effective stress


asserts that (1) the effective stress controls stress
strain, volume change, and strength, independent of the
magnitude of the pore pressure, and (2) the effective
stress is given by   u for a saturated soil.1
There is ample experimental evidence to show that
these statements are essentially correct for soils. The
principle is essential to describe the consolidation of a
liquid-saturated deformable porous solid, as was done
for the one-dimensional case by Terzaghi and further
developed for the three-dimensional case by others
such as Biot (1941). It is also an essential concept for
the understanding of soil liquefaction behavior during
earthquakes.
The total stress  can be directly measured or computed using the external forces and the body force due
to weight of the soilwater mixture. A pore water pressure, denoted herein by u0, can be measured at a point
remote from the interparticle zone. The actual pore water pressure in the interparticle zone is u. We know
that at equilibrium the total potential or head of the
water at the two points must be equal, but this does
not mean that u u0, as discussed in Section 7.7. The
effective stress is a deduced quantity, which in practice
is taken as   u0.

7.3 FORCE DISTRIBUTIONS IN A


PARTICULATE SYSTEM

The term intergranular stress has become synonymous


with effective stress. Whether or not the intergranular
stress i is indeed equal to  u cannot be ascertained
without more detailed examination of all the interpar-

1
The terms  and   are the principal total and effective stresses.
For general stress conditions, there are six stress components (11,
22, 33, 12, 23, and 31), where the rst three are the normal stresses
and the latter three are the shear stresses. In this case, the effective
 11 u, 22
 22 u, 33
 33
stresses are dened as 11
u,  12 12,  23
 23, and  31
 31.

Copyright 2005 John Wiley & Sons

When external forces are applied, both normal and


tangential forces develop at particle contacts. All particles do not share the forces or stresses applied at the
boundaries in equal manner. Each particle has different
skeletal forces depending on the position relative to the
neighboring particles in contact. The transfer of forces
through particle contacts from external stresses was
shown in Fig. 5.15 using a photoelastic model. Strong
particle force chains form in the direction of major
principal stress. The evolution and distribution of interparticle skeletal forces in soils govern the macroscopic stressstrain behavior, volume change, and
strength. As the soil approaches failure, buckling of
particle force chains occurs and shear bands develop
due to localization of deformation. Further discussion
of microbehavior in relation to deformation and
strength is given in Chapter 11.
Particle weights act as body forces in dry soil and
contribute to skeletal forces, observed in the photoelastic model shown in Fig. 5.15. When the pores are
lled with uids, the weight of the uids adds to the
body force of the soiluids mixture. However, hydrostatic pressure results from the uid weight, and the
uplift force due to buoyancy reduces the effective
weight of a uid-lled soil. This leads to smaller skeletal forces for submerged soil compared to dry soil.
Seepage forces that result from additional uid pressures applied externally produce hydrodynamic forces
on particles and alter the skeletal forces.

7.4

INTERPARTICLE FORCES

Long-range particle interactions associated with electrical double layers and van der Waals forces are dis-

Retrieved from: www.knovel.com

INTERPARTICLE FORCES

175

Body Force

External Load

Buoyancy Force
if Saturated
Viscous Drag by
Seepage Flow
Interparticle
Forces

Capillary Force or
Cementation-reactive
Force

Co
py
rig
hte
dM
ate
ria
l

Interparticle
Forces

Seepage

(a)

Electrical Forces

(b)

(c)

Figure 7.1 Interparticle forces at the particle level: (a) skeletal forces by external loading,

(b) particle level forces, and (c) contact level forces (after Santamarina, 2003).

cussed in Chapter 6. These interactions control the


occulationdeocculation behavior of clay particles
in suspension, and they are important in swelling soils
that contain expanding lattice clay minerals. In denser
soil masses, other forces of interaction become important as well since they may inuence the intergranular
stresses and control the strength at interparticle contacts, which in turn controls resistance to compression
and strength. In a soil mass at equilibrium, there must
be a balance among all interparticle forces, the pressure in the water, and the applied boundary stresses.
Interparticle Repulsive Forces

Electrostatic Forces Very high repulsion, the Born


repulsion, develops at contact points between particles.
It results from the overlap between electron clouds,
and it is sufciently great to prevent the interpenetration of matter.
At separation distances beyond the region of direct
physical interference between adsorbed ions and hydration water molecules, double-layer interactions provide the major source of interparticle repulsion. The
theory of these forces is given in Chapter 6. As noted
there, this repulsion is very sensitive to cation valence,
electrolyte concentration, and the dielectric properties
of the pore uid.
Surface and Ion Hydration The hydration energy
of particle surfaces and interlayer cations causes large
repulsive forces at small separation distances between
unit layers (clear distance between surfaces up to about
2 nm). The net energy required to remove the last few

Copyright 2005 John Wiley & Sons

layers of water when clay plates are pressed together


may be 0.05 to 0.1 J/m2. The corresponding pressure
required to squeeze out one molecular layer of water
may be as much as 400 MPa (4000 atm) (van Olphen,
1977).
Thus, pressure alone is not likely to be sufcient to
squeeze out all the water between parallel particle surfaces in naturally occurring clays. Heat and/or high
vacuum are needed to remove all the water from a negrained soil. This does not mean, however, that all the
water may not be squeezed from between interparticle
contacts. In the case of interacting particle corners,
edges, and faces of interacting asperities, the contact
stress may be several thousand atmospheres because
the interparticle contact area is only a very small proportion ( 1%) of the total soil cross-sectional area
in most cases. The exact nature of an interparticle contact remains largely a matter for speculation; however,
there is evidence (Chapter 12) that it is effectively solid
to solid.
Hydration repulsions decay rapidly with separation
distance, varying inversely as the square of the distance.
Interparticle Attractive Forces

Electrostatic Attractions When particle edges and


surfaces are oppositely charged, there is attraction due
to interactions between double layers of opposite sign.
Fine soil particles are often observed to adhere when
dry. Electrostatic attraction between surfaces at different potentials has been suggested as a cause. When the

Retrieved from: www.knovel.com

176

EFFECTIVE, INTERGRANULAR, AND TOTAL STRESS

gap between parallel particle surfaces separated by distance d at potentials V1 and V2 is conductive, there is
an attractive force per unit area, or tensile strength,
given by (Ingles, 1962)
F

4.4 106 (V1 V2)2


N/m2
d2

(7.1)

Co
py
rig
hte
dM
ate
ria
l

where F is the tensile strength, d is in micrometers,


and V1 and V2 are in millivolts. This force is independent of particle size and becomes signicant (greater
than 7 kN/m2 or 1 psi) for separation distances less
than 2.5 nm.
Electromagnetic Attractions Electromagnetic attractions caused by frequency-dependent dipole interactions (van der Waals forces) are described in Section
6.12. Anandarajah and Chen (1997) proposed a method
to quantify the van der Waals force between particles
specically for ne-grained soils with various geometric parameters such as particle length, thickness, orientation, and spacing.
Primary Valence Bonding Chemical interactions
between particles and between the particles and adjacent liquid phase can only develop at short range. Covalent and ionic bonds occur at spacings less than 0.3
nm. Cementation involves chemical bonding and can
be considered as a short-range attraction.
Whether primary valence bonds, or possibly hydrogen bonds, can develop at interparticle contacts without the presence of cementing agents is largely a
matter of speculation. Very high contact stresses between particles could squeeze out adsorbed water and
cations and cause mineral surfaces to come close together, perhaps providing opportunity for cold welding. The activation energy for soil deformation is high,
in the range characteristic for rupture of chemical
bonds, and strength behavior appears in reasonable
conformity with the adhesion theory of friction (Chapter 11). Thus, interatomic bonding between particles
seems possible. On the other hand, the absence of cohesion in overconsolidated silts and sands argues
against such pressure-induced bonding.
Cementation Cementation may develop naturally
from precipitation of calcite, silica, alumina, iron oxides, and possibly other inorganic or organic compounds. The addition of stabilizers such as cement and
lime to a soil also leads to interparticle cementation. If
two particles are not cemented, the interparticle force
cannot become tensile; they loose contact. However, if
a particle contact is cemented, it is possible for some
interparticle forces to become negative due to the tensile resistance (or strength) of the cemented bonds.

There is also an increase in resistance to tangential


force at particle contacts. However, when the bond
breaks, the shear capacity at a contact reduces to that
of the uncemented contacts.
An analysis of the strength of cemented bonds
should consider three cases: (i) failure in the cement,
(ii) failure in the particle and (iii) failure at the cementparticle interface. The following equation can be
derived (Ingles, 1962) for the tensile strength T per
unit area of soil cross section:

Copyright 2005 John Wiley & Sons

T Pk

1e

(7.2)

Ai

where P is the bond strength per contact zone, k is the


mean coordination number of a grain, e is the void
ratio, n is the number of grains in an ideal breakage
plane at right angles to the direction of T, and Ai is
the total surface area of the ith grain.
For a random and isotropic assembly of spheres of
diameter d, Eq. (7.2) becomes
T

Pk
d (1 e)
2

(7.3)

For a random and isotropic assembly of rods of length


l and diameter d
T

Pk
d(l d/2)(1 e)

(7.4)

Bond strength P is evaluated in the following way (Fig.


7.2) for two cemented spheres of radius R. It may be
shown that
 cosh

(R cos )
R sin 


(7.5)

so for known ,  can be computed. Then, for cement


failure,
P c  2

(7.6)

where c is the tensile strength of the cement; for


sphere failure,
P s  ()2

(7.7)

where  R sin , and s is the tensile strength of


the sphere, and for failure at the interface

Retrieved from: www.knovel.com

INTERPARTICLE FORCES

177

Co
py
rig
hte
dM
ate
ria
l

mented natural materials, if the soil is unloaded from


high overburden stress, elastic rebound may disrupt cemented bonds.
Cementation allows interparticle normal forces to
become negative, and, therefore, the distribution and
evolution of skeletal forces may be different than in
uncemented soils, even though the applied external
stresses are the same. Thus, the stiffness and strength
properties of a soil are likely to differ according to
when and how cementation was developed. How to
account for this in terms of effective stress is not yet
clear.
Capillary Stresses Because water is attracted to
soil particles and because water can develop surface
tension, suction develops inside the pore uid when a
saturated soil mass begins to dry. This suction acts like
a vacuum and will directly contribute to the effective
stress or skeletal forces. The negative pore pressure is
usually considered responsible for apparent and temporary cohesion in soils, whereas the other attractive
forces produce true cohesion.
When the soil continues to dry, air starts to invade
into the pores. The air entry pressure is related to the
pore size and can be estimate using the following equation, assuming a capillary tube as shown in Fig. 7.3a:

Figure 7.2 Contact zone failures for cemented spheres.

P 1

sin 


2R2(1 cos )

(7.8)

where 1 is the tensile strength of the interface bond.


In principle, Eq. (7.6), (7.7), or (7.8) can be used to
obtain a value for P in Eq. (7.2) enabling computation
of the tensile strength T of a cemented soil.
The behavior of cemented soils can depend on the
timing of cementation development. Articially cemented soils are often loaded after cementation has
developed, whereas cementation develops during or after overburden loading in natural soils. In the former
case, the particles and cementation bonding are loaded
together and contact forces can become negative depending on the tensile resistance of cementation bonding. The distribution and magnitude of skeletal forces
are therefore inuenced by both geometric arrangement of particles and the cementation bonding at the
particle contacts. In the latter case, on the other hand,
the contact forces induced by external loading are developed before cementation coats the already loaded
particles. In this case, it is possible that cementation
creates extra forces at particle contacts. In some ce-

Copyright 2005 John Wiley & Sons

P c

2aw cos
rp

(7.9)

where P c is the capillary pressure at air entry, aw is


the airwater interfacial tension, is contact angle dened in Fig. 7.3, and rp is the tube radius. For pure
water and air, aw depends on temperature, for example, it is 0.0756 N/m at 0C, 0.0728 N/m at 20C, and
0.0589 N/m at 100C. If the capillary pressure Pc
( ua uw, where ua and uw are the air and water
pressures, respectively) is larger than P c, then air invades the pore.2 Since soil has pores with various sizes,
the water in the largest pores is displaced rst followed
by smaller pores. This leads to a macroscopic model
of the soilwater characteristic curve (or the capillary
pressuresaturation relationship), as discussed in Section 7.11.
If the water surrounding the soil particles remains
continuous [termed the funicular regime by Bear
(1972)], the interparticle force acting on a particle with
radius r can be estimated from

2
It is often assumed that ua 0 (for gauge pressure) or 1 atm (for
absolute pressure). However, this may not be true in cases such as
rapid water inltration when air in the pores cannot escape or the air
boundary is completely blocked.

Retrieved from: www.knovel.com

178

EFFECTIVE, INTERGRANULAR, AND TOTAL STRESS

Capillary Tube
Representing a Pore
2 rp

ua

uw Pc = w gdc =

Co
py
rig
hte
dM
ate
ria
l

dc

2aw cos
rp

(a)

(b)

Figure 7.3 Capillary tube concept for air entry estimation: (a) capillary tube and (b) bundle

of capillary tubes to represent soil pores with different sizes.

2 r 2 aw cos
Fc  r 2 P c
rp

(7.10)

where rp is the size of the pore into which the air has
entered. Since the uid acts like a membrane with negative pressure, this force contributes directly to the
skeletal forces like the water pressure as shown in Fig.
7.4a.
As the soil continues to dry, the water phase becomes disconnected and remains in the form of menisci or liquid bridges at the interparticle contacts
[termed the pendular regime by Bear (1972)]. The
curved airwater interface produces a pore water tension, which, in turn, generates interparticle compressive forces. The force only acts at particle contacts in
contrast to the funicular regime, as shown in Fig. 7.4b.
The interparticle force generally depends on the separation between the two particles, the radius of the liquid bridge, interfacial tension, and contact angle (Lian
et al., 1993). Once the water phase becomes discontinuous, evaporation and condensation are the primary
mechanisms of water transfer. Hence, the humidity of
the gas phase and the temperature affect the water vapor pressure at the surface of water menisci, which in
turn inuences the air pressure ua.

7.5

INTERGRANULAR PRESSURE

Several different interparticle forces were described in


the previous section. Quantitative expression of the in-

Copyright 2005 John Wiley & Sons

teractions of all these forces in a soil is beyond the


present state of knowledge. Nonetheless, their existence bears directly on the magnitude of intergranular
pressure and the relationship between intergranular
pressure and effective stress as dened by 
 u.
A simplied equation for the intergranular stress in
a soil may be developed in the following way. Figure
7.5 shows a horizontal surface through a soil at some
depth. Since the stress conditions at contact points,
rather than within particles, are of primary concern, a
wavy surface that passes through contact points (Fig.
7.5a) is of interest. The proportion of the total wavy
surface area that is comprised of intergrain contact area
is very small (Fig. 7.5c).
The two particles in Fig. 7.5 that contact at point A
are shown in Fig. 7.6, along with the forces that act in
a vertical direction. Complete saturation is assumed.
Vertical equilibrium across wavy surface xx is considered.3 The effective area of interparticle contact is
ac; its average value along the wavy surface equals the
total mineral contact area along the surface divided by
the number of interparticle contacts. Dene area a as

3
Note that only vertical forces at the contact are considered in this
simplied analysis. It is evident, however, that applied boundary normal and shear stresses each induce both normal and shear forces at
interparticle contacts. These forces contribute both to the development of soil strength and resistance to compression and to the slipping and sliding of particles relative to each other. These interparticle
movements are central to compression, shear deformations, and creep
as discussed in Chapters 10, 11, and 12.

Retrieved from: www.knovel.com

INTERGRANULAR PRESSURE

179

Continuous
Water Film

Interparticle
Forces
Soil Particles
Soil Particles

Co
py
rig
hte
dM
ate
ria
l

Air
Liquid
Bridges

Pores of Radius
rp Filled with Air

Negative pore pressure acts all


around the particles
(a)

Suction forces act only at particle


contacts and the magnitude of the
forces depends on the size of liquid
bridges.
(b)

Figure 7.4 Microscopic watersoil interaction in unsaturated soils: (a) funicular regime and
(b) pendular regime.

Figure 7.6 Forces acting on interparticle contact A.

the average total cross-sectional area along a horizontal


plane served by the contact. It equals the total horizontal area divided by the number of interparticle contacts along the wavy surface. The forces acting on area
a in Fig. 7.6 are:

Figure 7.5 Surfaces through a soil mass.

Copyright 2005 John Wiley & Sons

1. a, the force transmitted by the applied stress ,


which includes externally applied forces and
body weight from the soil above.

Retrieved from: www.knovel.com

180

EFFECTIVE, INTERGRANULAR, AND TOTAL STRESS

where  aw /a. Although it is clear that for a dry soil


 0, and for a saturated soil  1.0, the usefulness
of Eq. (7.15) has been limited in practice because of
uncertainties about  for intermediate degrees of saturation. Further discussion of the effective stress concept for unsaturated soils is given in Section 7.12.
Limiting the discussion to saturated soils, two questions arise:
1. How does the intergranular pressure i relate to
the effective stress as dened for most analyses,
that is,   u?
2. How does the intergranular pressure i relate to
the measured quantity, m  u0, that is taken
as the effective stress, recalling (Section 7.2) that
pore pressure can only be measured at points outside the true interparticle zone?

Co
py
rig
hte
dM
ate
ria
l

2. u(a ac), the force carried by the hydrostatic


pressure u. Because a ac and ac is very small,
the force may be taken as ua. Long-range,
double-layer repulsions are included in ua.
3. A(a ac) Aa, the force caused by the longrange attractive stress A, that is, van der Waals
and electrostatic attractions.
4. Aac, the force developed by the short-range attractive stress A, resulting from primary valence
(chemical) bonding and cementation.
5. Cac, the intergranular contact reaction that is generated by hydration and Born repulsion.
Vertical equilibrium of forces requires that
a Aa Aac ua Cac

(7.11)

Division of all terms by a converts the forces to


stresses per unit area of cross section,
 (C A)

ac
uA
a

Answers to these questions require a more detailed


consideration of the meaning of uid pressures in soils.

(7.12)

7.6

The term (C A)ac /a represents the net force across


the contact divided by the total cross-sectional area
(soil plus water) that is served by the contact. In other
words, it is the intergrain force divided by the gross
area or the intergranular pressure in common soil mechanics usage. Designation of this term by i gives
i  A u

(7.13)

Equations analogous to Eqs. (7.11), (7.12), and (7.13)


can be developed for the case of a partly saturated soil.
To do so requires consideration of the pressures in the
water uw and in the air ua and the proportions of area
a contributed by water aw and by air aa with the condition that
a w aa a

i.e., ac 0

WATER PRESSURES AND POTENTIALS

Pressures in the pore uid of a soil can be expressed


in several ways, and the total pressure may involve
several contributions. In hydraulic engineering, problems are analyzed using Bernoullis equation for the
total heads and head losses associated with ow between two points, that is,
Z1

p1
v2
p
v2
1 Z2 2 2 h12
w 2g
w 2g

where Z1 and Z2 are the elevations of points 1 and 2,


p1 and p2 are the hydrostatic pressures at points 1 and
2, v1 and v2 are the ow velocities at points 1 and 2,
w is the unit weight of water, g is the acceleration due
to gravity, and h12 is the loss in head between points
1 and 2. The total head H (dimension L) is
HZ

The resulting equation is

i  A ua

aw
(u ua)
a w

(7.14)

In the absence of signicant long-range attractions,


this equation is similar to that proposed by Bishop
(1960) for partially saturated soils
i  ua  (ua uw)

(7.15)

Copyright 2005 John Wiley & Sons

(7.16)

p
v2

w 2g

(7.17)

Flow results only from differences in total head;


conversely, if the total heads at two points are the
same, there can be no ow, even if Z1 Z2 and p1
p2. If there is no ow, there is no head loss and h12
0.
The ow velocity through soils is low, and as a result v 2 /2g 0, and in most cases it may be neglected.
Therefore, the relationship

Retrieved from: www.knovel.com

WATER PRESSURE EQUILIBRIUM IN SOIL

Z1

p1
p
Z2 2 h12
w
w

(7.18)

is the basis for evaluation of pore pressures and analysis of seepage through soils and other porous media.
Although the absence of velocity terms is a factor
that seems to simplify the analysis of ows and pressures in soils, there are other considerations that tend
to complicate the problem. These include:

1. Gravitational potential g (head Z, pressure pz)


corresponds to elevation head in normal hydraulic usage.
2. Matrix or capillary potential m (head hm, pressure p) is the work per unit quantity of water to
transport reversibly and isothermally an innitesimal quantity of water to the soil from a pool
containing a solution identical in composition to
the soil water at the same elevation and external
gas pressure as that of the point under consideration in the soil. This component corresponds to
the pressure head in normal hydraulic usage. It
results from that part of the boundary stresses
that is transmitted to the water phase, from pressures generated by capillarity menisci, and from
water adsorption forces exerted by particle surfaces. A piezometer measures the matrix potential if it contains uid of the same composition
as the soil water.
3. Osmotic (or solute) potential s (head hs, pressure ps) is the work per unit quantity of water to
transport reversibly and isothermally an innitesimal quantity of water from a pool of pure water
at a specied elevation and atmospheric pressure
to a pool containing a solution identical in composition to the soil water, but in all other respects
identical to the reference pool. This component
is, in effect, the osmotic pressure of the soil water, and it depends on the composition and ability
of the soil particles to restrain the movement of
adsorbed cations. The osmotic potential is negative, that is, water tends to ow in the direction
of increasing concentration.

Co
py
rig
hte
dM
ate
ria
l

1. The use of several terms to describe the status of


water in soils, for example, potential, pressure,
and head.
2. The possible existence of tensions in the pore water.
3. Compositional differences in the water from
point-to-point and adsorptive force elds from
particle surfaces.
4. Differences in interparticle forces and the energy
state of the pore uid from point to point owing
to thermal, electrical, and chemical gradients.
Such gradients can cause uid ows, deformations, and volume changes, as considered in more
detail in Chapter 9.
Some formalism in denition and terminology is
necessary to avoid confusion. The status of water in a
soil can be expressed in terms of the free energy relative to free, pure water (Aitchison, et al., 1965). The
free energy can be (and is) expressed in different ways,
including
1. Potential (dimensionsL2T2: J/kg)
2. Head (dimensionsL: m, cm, ft)
3. Pressure (dimensionsML1 T2: kN/m2, dyn/
cm2, tons/m2, atm, bar, psi, psf)

If the free energy is less than that of pure water


under the ambient air pressure, the terms suction and
negative pore water pressure are used.
The total potential (head, pressure) of soil water is
the potential (head, pressure) in pure water that will
cause the same free energy at the same temperature as
in the soil water. An alternative denition of total potential is the work per unit quantity to transport reversibly and isothermally an innitesimal amount of
pure water from a pool at a specied elevation at atmospheric pressure to the point in soil water under
consideration.
The selection of the components of the total potential  (total head H, total pressure P) is somewhat
arbitrary (Bolt and Miller, 1958); however, the following have gained acceptance for geotechnical work
(Aitchison, et al., 1965):

Copyright 2005 John Wiley & Sons

181

The total potential, head, and pressure then become


 g m s

(7.19)

H Z hm hs

(7.20)

P pz p ps

(7.21)

At equilibrium and no ow there can be no variations in , H, or P within the soil.


7.7

WATER PRESSURE EQUILIBRIUM IN SOIL

Consider a saturated soil mass as shown in Fig. 7.7.


Conditions at several points will be analyzed in terms
of heads for simplicity, although potential or pressure
could also be used with the same result. The system is
assumed at constant temperature throughout. At point
0, a point inside a piezometer introduced to measure

Retrieved from: www.knovel.com

EFFECTIVE, INTERGRANULAR, AND TOTAL STRESS

Co
py
rig
hte
dM
ate
ria
l

182

Figure 7.7 Schematic representation of a saturated soil for analysis of pressure conditions.

pore pressure, Z 0, hm hm0, and hs0 0 if pure


water is used in the piezometer. Thus,
H0 0 hm0 0 hm0

It follows that

At point 2, which is between the same two clay particles as point 1 but closer to a particle surface, there
will be a different ion concentration than at 1. Thus,
at equilibrium, and assuming Z2 0,
hm2 hs2 hm1 hs1 hm0

P0 hm0 w u0

(7.22)

the measured pore pressure.


Point 1 is at the same elevation as point 0, except it
is inside the soil mass and midway between two clay
particles. At this point, Z1 0, but hs 0 because the
electrolyte concentration is not zero. Thus,
H1 0 hm1 hs1

A similar analysis could be applied to any point in the


system. If point 3 were midway between two clay particles spaced the same distance apart as the particles
on either side of point 1, then hs3 hs1, but Z3 0.
Thus,
u0
Z3 hm3 hs3 Z3 hm3 hs1
w

(7.24)

A partially saturated system can also be analyzed,


but the inuences of curved airwater interfaces must
be taken into account in the development of the hm
terms.
The conclusions that result from the above analysis
of component potentials are:

If no water is owing, H1 H0, and


hm1 hs1 hm0

Also, because p1 p0 u0
u0 hm1 w hs1 w

u0
w

(7.23)

Copyright 2005 John Wiley & Sons

1. As the osmotic and gravitational components


vary from point to point in a soil at equilibrium,

Retrieved from: www.knovel.com

MEASUREMENT OF PORE PRESSURES IN SOILS

3.

reach equilibrium, and the suction can be determined by the water content of the lter paper.
These techniques are used for measurement of
pore pressures less than atmospheric.
Pressure-Membrane Devices An exposed soil
sample is placed on a membrane in a sealed
chamber. Air pressure in the chamber is used to
push water from the pores of the soil through the
membrane. The relationship between water content and pressure is used to establish the relationship between soil suction and water content.
Consolidation Tests The consolidation pressure
on a sample at equilibrium is the soil water suction. If the consolidation pressure were instantaneously removed, then a negative water pressure
or suction of the same magnitude would be
needed to prevent water movement into the soil.
Vapor Pressure Methods The relationship between relative humidity and water content is used
to establish the relationship between suction and
water content.
Osmotic Pressure Methods Soil samples are
equilibrated with solutions of known osmotic
pressure to give a relationship between water
content and water suction.
Dielectric Sensors Such as Capacitance Probes
and Time Domain Reectometry Soil moisture
can be indirectly determined by measuring the
dielectric properties of unsaturated soil samples.
With the knowledge of soil water characteristics
relationship (Section 7.11), the negative pore
pressure corresponding to the measured soil
moisture can be determined. The capacitance
probe measures change in frequency response of
the soils capacitance, which is related to dielectric constants of soil particle, water, and air. The
capacitance is largely inuenced by water content, as the dielectric constant of water is large
compared to the dielectric constants of soil
particle and air. Time domain reectrometry
measures the travel time of a high-frequency,
electromagnetic pulse. The presence of water in
the soil slows down the speed of the electromagnetic wave by the change in the dielectric properties. Volumetric water content can therefore be
indirectly measured from the travel time measurement.

Co
py
rig
hte
dM
ate
ria
l

the matrix or capillary component must also vary


to maintain equal total potential. The concept that
hydrostatic pressure must vary with elevation to
maintain equilibrium is intuitive; however, the
idea that this pressure must vary also in response
to compositional differences is less easy to visualize. Nonetheless, this underlies the whole
concept of water ow by chemical osmosis.
2. The total potential, head, and pressure are measurable, and separation into components is possible experimentally, although it is difcult.
3. A pore pressure measurement using a piezometer
containing pure water gives a pressure u0 wh,
where h is the pressure head at the piezometer.
When referred back to points between soil particles, u0 is seen to include contributions from
osmotic pressures as well as matrix pressures.
Since osmotic pressures are the cause of longrange repulsions due to double-layer interactions,
measured pore water pressures may include contributions from long-range interparticle repulsive
forces.

7.8 MEASUREMENT OF PORE PRESSURES IN


SOILS

Several techniques for the measurement of pore water


pressures are available. Some are best suited for laboratory use, whereas others are intended for use in the
eld. Some yield the pore pressure or suction by direct
measurement, while others require deduction of the
value using thermodynamic relationships.

1. Piezometers of Various Types Water in the piezometer communicates with the soil through a
porous stone or lter. Pressures are determined
from the water level in a standpipe, by a manometer, by a pressure gauge, or by an electronic
pressure transducer. A piezometer used to measure pressures less than atmospheric is usually
termed a tensiometer.
2. Gypsum Block, Porous Ceramic, and Filter
Paper The electrical properties across a specially prepared gypsum block or porous ceramic
block are measured. The water held by the block
determines the resistance or permittivity, and the
moisture tension in the surrounding soil determines the amount of moisture in the block
(Whalley et al., 2001). The same principle can be
applied by placing a dry lter paper on a soil
specimen and allowing the soil moisture to absorb into the paper. When the suction in the lter
paper is equal to the suction in the soil, the two

Copyright 2005 John Wiley & Sons

4.

5.

6.

7.

183

Piezometer methods are used when positive pore


pressures are to be measured, as is usually the case in
dams, slopes, and foundations on soft clays. The other
methods are suitable for measurement of negative pore
pressures or suction. Pore pressures are often negative
in expansive and partly saturated soils. More detailed

Retrieved from: www.knovel.com

184

EFFECTIVE, INTERGRANULAR, AND TOTAL STRESS

descriptions and comparisons of these and other methods are given by Croney et al. (1952), Aitchison et al.
(1965), Richards and Peter (1987), and Ridley et al.
(2003).

dened effective stress   u0 differ by the net


interparticle stress due to physicochemical contributions,
i  A R

7.9 EFFECTIVE AND INTERGRANULAR


PRESSURE

i  A u

(7.25)

where u is the hydrostatic pressure between particles


(or hm w in the terminology of Section 7.7). Generalized forms of Eq. (7.24) are

and

When A and R are both small, as would be true in


granular soils, silts, and clays of low plasticity, or in
cases where A R, the intergranular and effective
stress are approximately equal. Only in cases where
either A or R is large, or both are large but of signicantly different magnitude, would the intergranular and
effective stress be signicantly different. Such a condition appears not to be common, although it might be
of importance in a well-dispersed sodium montmorillonite, where compression behavior can be accounted
for reasonably well in terms of double-layer repulsions
(Chapter 10).4
The derivation of Eq. (7.30) assumed vertical equilibrium, with contributing forces parallel to each other,
that is, the intergranular stress i is the sum of the
skeletal forces (dened as   u0) and the electrochemical stress (A R), as illustrated in Fig. 7.8a.
This implies that the deformation induced by the electrochemical stress (A R) is equal to the deformation
induced by the skeletal forces at contacts [i.e., a parallel model as described by Hueckel (1992)]. The
change in pore uid chemistry at constant connement
() leads to changes in intergranular stresses (i), resulting in changes in shear strength, for example.
An alternative assumption can be made; the total
deformation of soil is the sum of the deformations of
the particles and in the double layers as illustrated in
Fig. 7.8b. The effective stress  is then equal to the
electrochemical stress (R A):

Co
py
rig
hte
dM
ate
ria
l

In Section 7.5, it was shown that the intergranular pressure is given by

u0 Z w hm w hs w

(7.26)

u hm w u0 Z w hs w

(7.27)

Thus, Eq. (7.25) becomes, for the case of no elevation


difference between a piezometer and the point in question (i.e., Z 0),
i  A u0 hs w

(7.28)

Because the quantity hs w is an osmotic pressure and


the salt concentration between particles will invariably
be greater than at points away from the soil (such as
in a piezometer), hs w will be negative. This pressure
reects double-layer repulsions. It has been termed R
in some previous studies (Lambe, 1960; Mitchell,
1962). If hs w in Eq. (7.28) is replaced by the absolute
value of R, we obtain
i  A u0 R

(7.30)

i R A   u0

(7.31)

(7.29)

From Eq. (7.25), it was seen that the intergranular


pressure was dependent on long-range interparticle attractions A as well as on the applied stress  and the
pore water pressure between particles u. Equation
(7.29) indicates that if intergranular pressure i is to
be expressed in terms of a measured pore pressure u0,
then the long-range repulsion R must also be taken into
account. The actual hydrostatic pressure between particles u u0 R includes the effects of long-range
repulsions as required by the condition of constant total potential for equilibrium.
In the general case, therefore, the true intergranular
pressure i  A u0 R and the conventionally

Copyright 2005 John Wiley & Sons

This is called the series model (Hueckel, 1992), and


the model can be applicable for very ne soils at high
water content, in which particles are not actually in
contact with each other but are aligned in a parallel
arrangement. Increase in intergranular stress i or effective stress  changes the interparticle spacing,
which may contribute to changes in strength properties
upon shearing.

4
A detailed analysis of effective stress in clays is presented by Chattopadhyay (1972), which leads to similar conclusions, including Eq.
(7.29). i was termed the true effective stress and it governed the
volume change behavior of Namontmorillonite.

Retrieved from: www.knovel.com

ASSESSMENT OF TERZAGHIS EQUATION

Skeletal Force

Skeletal Force
Electrochemical Force

185

Electrochemical Force
Skeletal Force

Skeletal Force

Electrochemical Force
Electrochemical Force

i

Co
py
rig
hte
dM
ate
ria
l

i
Skeletal Force

Skeletal Force

Electrochemical
Force A _ R

 = _ u0

 = _ u0

Particle Deformation
by Skeletal Force

Electrochemical
Force A _ R

Deformation at
the Contact

i

i = _ u0 + A _ R
(a)

Total Deformation
at the Contact

i

i = _ u0 = A _ R
(b)

Figure 7.8 Contribution of skeletal force ( u0) and electrochemical force (A R) to


intergranular force i: (a) parallel model and (b) series model.

Since the particles are arranged in parallel as well


as nonparallel manner, the chemomechanical coupling
behavior of actual soils can be far from the predictions
made by the above two models. In fact, Santamarina
(2003) argues that the impact of skeletal forces by external forces, particle-level forces, and contact-level
forces on soil behavior is different, and mixing both
types of forces in a single algebraic expression in terms
of effective stress can lead to incorrect prediction [e.g.,
Eq. (7.15) for unsaturated soils and Eq. (7.30) for soils
with measurable interparticle repulsive and attractive
forces].

of saturated soils. Skempton proposed three possible


relationships for effective stress in saturated soils:
1. The true intergranular pressure for the case when
AR0
  (1 ac)u

(7.32)

in which ac is the ratio of contact area to total


cross-sectional area.
2. The solid phase is treated as a real solid that has
compressibility Cs and shear strength given by
i k  tan 

7.10

ASSESSMENT OF TERZAGHIS EQUATION

The preceding equations and discussion do not conrm


that Terzaghis simple equation is indeed the effective
stress that governs consolidation and strength behavior
of soils. However, its usefulness has been established
from the experience of many years of successful application in practice. Skempton (1960b) showed that
the Terzaghi equation does not give the true effective
stress but gives an excellent approximation for the case

Copyright 2005 John Wiley & Sons

(7.33)

where  is an intrinsic friction angle and k is a


true cohesion. The following relationships were
derived: For shear strength,

  1

ac tan 
u
tan 

(7.34)

where  is the effective stress angle of shearing


resistance. For volume change,

Retrieved from: www.knovel.com

186

EFFECTIVE, INTERGRANULAR, AND TOTAL STRESS

  1

Cs
u
C

(7.35)

where C is the soil compressibility.


3. The solid phase is a perfect solid, so that  0
and Cs 0. This gives
  u

(7.36)

Compressibility a
per kN/m2 106
Material

Quartzitic sandstone
0.059
Quincy granite (30 m deep)
0.076
Vermont marble
0.18
Concrete (approx.)
0.20
Dense sand
18
Loose sand
92
London clay (over cons.)
75
Gosport clay (normally cons.) 600

Co
py
rig
hte
dM
ate
ria
l

To test the three theories, available data were studied


to see which related to the volume change of a system
acted upon by both a total stress and a pore water
pressure according to

Table 7.1 Compressibility Values for Soil, Rock,


and Concrete

C 

(7.37)

and also satised the Coulomb equation for drained


shear strength d :
d c  tan 

Cs /C

0.027
0.019
0.014
0.025
0.028
0.028
0.020
0.020

0.46
0.25
0.08
0.12
0.0015
0.0003
0.00025
0.00003

After Skempton (1960b).


a
Compressibilities at p 98 kN/m2; water Cw 0.49
106 per kN/m2.

(7.38)

when both a total stress and a pore pressure are acting.


It may be noted that this approach assumes that the
Coulomb strength equation is valid a priori.
The results of Skemptons analysis showed that Eq.
(7.32) was not a valid representation of effective stress.
Equations (7.34) and (7.35) give the correct results for
soils, concrete, and rocks. Equation (7.36) accounts
well for the behavior of soils but not for concrete and
rock. The reason for this latter observation is that in
soils Cs /C and ac tan  /tan  approach zero, and,
thus, Eqs. (7.34) and (7.35) reduce to Eq. (7.36). In
rock and concrete, however, Cs /C and ac tan  /tan 
are too large to be neglected. The value of tan  /tan
 may range from 0.1 to 0.3, ac clearly is not negligible, and Cs /C may range from 0.1 to 0.5 as indicated
in Table 7.1.
Effective stress equations of the form of Eqs. (7.32),
(7.34), (7.35), and (7.36) can be generalized to the general form (Lade and de Boer, 1997):
  u

Cs

(7.39)

where  is the fraction of the pore pressure that gives


the effective stress.5 Different expressions for  proposed by several researchers are listed in Table 7.2.

A more general expression has been proposed as ij ij iju,


where ij is the tensor that accounts for the constitutive characteristics
of the solid such as complex kinematics associated with anisotropic
elastic materials (Carroll and Katsube, 1983; Coussy, 1995; Didwania, 2002).
5

Copyright 2005 John Wiley & Sons

A more rigorous evaluation of the contribution of


soil particle compressibility to effective stress was
made by Lade and de Boer (1997) using a two-phase
mixture theory. The volume change of the soil skeleton
can be separated into that due to pore pressure increment u and that due to the change in conning pressure ( u) (or  u). The effective stress
increment  is dened as the stress that produces the
same volume change,
CV0  Vsks Vsku CV0(  u)
CuV0 u

(7.40)

where Vsks is the volume change of soil skeleton due


to change in conning pressure, Vsku is the volume
change of soil skeleton due to pore pressure change,
V0 is the initial volume, C is the compressibility of the
soil skeleton by conning pressure change, and Cu is
the compressibility of the soil skeleton by pore pressure change. Rearranging Eq. (7.40) leads to
 

Cu
u
C

(7.41)

Lade and de Boer (1997) used this equation to derive an effective stress equation for granular materials
under drained conditions. Consider a condition in
which the total conning pressure is constant [ (

Retrieved from: www.knovel.com

ASSESSMENT OF TERZAGHIS EQUATION

Table 7.2

187

Expressions for to Dene Effective Stress

Pore Pressure Fraction 


1
n

Note

Reference

n porosity
ac grain contact area per unit area of plane
Equation (7.34)

1 ac
tan 
1 ac
tan 
C
1 s
C

Biot and Willis (1957), Skempton


(1960b), Nur and Byerlee (1971), Lade
and de Boer (1997)

Co
py
rig
hte
dM
ate
ria
l

Equation (7.35); for isotropic elastic


deformation of a porous material; for solid
rock with small interconnected pores and
low porosity (Lade and de Boer, 1997)
Equation (7.43)

Terzaghi (1925b)
Biot (1955)
Skempton and Bishop (1954)
Skempton (1960b)

1 (1 n)

Cs
C

Suklje (1969); Lade and de Boer (1997)

After Lade and de Boer (1997).

u) 0], but the pore pressure changes by u.6 The


volume change of soil skeleton caused by change in
pore pressure ( Vsku) is attributed solely from the volumetric compression of the solid grains ( Vgu). Hence,
Vsku CuV0 u Cs(1 n)V0 u Vgu

Cu Cs(1 n)

or

(7.42)

where Cs is the compressibility of soil grains due to


pore pressure change and n is the porosity. Substituting
Eq. (7.42) into (7.41) gives
 

1 (1 n)

 1 (1 n)

Cs
C

Cs
u
C

or

(7.43)

Figure 7.9 shows the variations of  with stress for


quartz sand and gypsum sand (Lade and de Boer,
1997). For a stress level less than 20 MPa,  is essentially one. Thus, Terzaghis effective stress equation,
while not rigorously correct, is again shown to be an
excellent approximation in almost all cases for saturated soils (i.e., solid grains and pore uid are considered to be incompressible compared to soil skeleton
compressibility).

An example of this condition is a soil under a seabed, in which the


sea depth varies. This condition is often called the unjacked condition.

Copyright 2005 John Wiley & Sons

Figure 7.9 Variation of  with stress for quartz sand and

gypsum sand (Lade and de Boer, 1997).

Can the effective stress concept also be applied for


undrained conditions where drainage is prevented?
That is, when an isotropic total stress load of iso is
applied, is u equal to iso? Using a two-phase mixture theory, the total stress increment ( iso) is separated into partial stress increments for the solid phase
( s) and the uid phase ( ) (Oka, 1996). Considering that the macroscopic volumetric strains by two
phases are equal but of opposite sign for undrained

Retrieved from: www.knovel.com

188

EFFECTIVE, INTERGRANULAR, AND TOTAL STRESS

conditions, Oka (1996) showed that the partial stresses


are related to the total stress as follows:


C Cs
iso
(C/n) (1 1/n)Cs Cl

Solid Surface

(7.44)

(a)

[(1/n) 1]C (Cs /n) Cl


s
iso
(C/n) (1 1/n)Cs Cl

Water
(reference fluid)

Co
py
rig
hte
dM
ate
ria
l

where n is the porosity, C is the compressibility of soil


skeleton, Cs is the compressibility of soil particles, and
Cl is the compressibility of pore uid.
If the excess pore pressure generated by undrained
isotropic loading  is u, the partial stress increment
for the uid phase becomes (Oka, 1996)
 n u

(7.45)

C Cs
iso
C Cs n(Cl Cs)

Air

Solid surface
(b)

Water

Air

Combining Eqs. (7.45) and (7.46),

Solid

(7.46)

The multiplier in the right-hand side of the above


equation is in fact Bishops pore water pressure coefcient B (Bishop and Eldin, 1950).7 For typical soils
(Cs 1.9 2.7 108 m2 / kN, Cl 4.9 109
m2 /kN, C 105 104 m2 /kN), so the values of B
are roughly equal to 1. Hence, it can be concluded that
Terzaghis effective stress equation is also applicable
for undrained conditions for most soils.

7.11

Air

Water
(reference fluid)

WATERAIR INTERACTIONS IN SOILS

Wettability refers to the afnity of one uid for a solid


surface in the presence of a second or third uid or
gas. A measure of wettability is the contact angle,
which was introduced in Eq. (7.9). Figure 7.10 illustrates a drop of the reference liquid (water for Fig.
7.10a and air for Fig. 7.10b) resting on a solid surface
in the presence of another uid (air for Fig. 7.10a and
water for 7.10b). The interface between the two uids
meets the solid surface at a contact angle . If the angle
is less than 90, the reference uid is referred to as the
wetting uid for a given solid surface. If the angle is
greater than 90, the reference liquid is referred to as
the nonwetting phase. The gure shows that water and

A similar equation for B value has been proposed by Lade and de


Boer (1997).

Copyright 2005 John Wiley & Sons

(c)

Figure 7.10 Wettability of two uids (water and air) on a


solid surface: (a) contact angle less than 90, (b) contact angle more than 90, and (c) unsaturated sand with water as the

wetting uid and air as the nonwetting uid.

air are the wetting and nonwetting uid, respectively.8


The environmental SEM photos in Fig. 5.27 showed
that water can be either wetting or nonwetting uid
depending soil mineralogy.
The contact angle is a property related to interactions of solid and two uids (water and air, in this
case).
cos

as ws
aw

(7.47)

where as is the interfacial tension between air and


solid, ws is the interfacial tension between water
and solid, and aw is the interfacial tension between

8
Some contaminated sites contain non-aqueous-phase liquids
(NAPLs). In general, NAPLS can be assumed to be nonwetting with
respect to water since the soil particles are in general primarily
strongly water-wet. Above the water table, it is usually appropriate
to assume that the water is the wetting uid with respect to NAPL
and that NAPL is a wetting uid with respect to air, implying that
the wettability order is water NAPL air. Below the water table,
water is the wetting uid and NAPL is the nonwetting uid.

Retrieved from: www.knovel.com

WATERAIR INTERACTIONS IN SOILS

air and water. The microscopic scale distribution of


water and air is illustrated in Fig. 7.10c, whereby it is
assumed that water is wetting the grain surfaces.
The aforementioned discussion on wettability and
contact angle assumes static water drops on solid surfaces. It has been observed for movement of water relative to soil that the dynamic contact angle formed
by the receding edge of a water droplet is generally
less than the angle formed by its advancing edge.
Matric suction (or capillary pressure) refers to the
pressure discontinuity across a curved interface separating two uids. This pressure difference exists because of the interfacial tension present in the uid
uid interface. Matric suction is a property that causes
porous media to draw in the wetting uid and repel
the nonwetting uid and is dened as the difference
between the nonwetting uid pressure and the wetting
uid pressure. For a two-phase system consisting of
water and air, the matric suction  is

1 Dune Sand
2 Loamy Sand
3 Calcareous Fine Sandy Loam
4 Calcareous Loam
5 Silt Loam Derived from Loess
6 Young Oligotrophous Peat Soil
7 Marine Clay

105
7
6

103

Co
py
rig
hte
dM
ate
ria
l

Matric suction ua uw (kPa)

106

104

 un uw

102

101

189

100

10-1
0.0

0.1

0.2

0.3

0.4

0.5

0.6

Volumetric Water Content w

(7.48)

Figure 7.11 Soilwater characteristic curves for some Dutch

soils (from Koorevaar et al., 1983; copied from Fredlund and


Rahardjo, 1993).

where un is the pressure of the nonwetting uid (air)


and uw is the pressure of the wetting uid (water).
Assuming that the soil pores have a cylindrical
shape, like a bundle of capillary tubes as illustrated in
Fig 7.3b, the interface between two liquids in each tube
forms a subsection of a sphere. The capillary pressure
is then related to the tube radius, contact angle, and
the interfacial tension between the two liquids. The
pressure drop across the interface is directly proportional to the interfacial tension and inversely proportional to the radius of curvature. It follows that higher
air pressure is required for air to enter water-saturated
ne-grained than coarse-grained materials.
Soil contains a range of different pore sizes, which
will drain at different capillary pressure values. This
leads to a soilwater characteristic relationship in
which the matric suction is plotted against the volumetric water content (or sometimes water saturation
ratio) such as shown in Fig. 7.11.9 The curves are often
determined during air invasion into a previously watersaturated soil. As the volumetric water content decreases, as a result of drainage or evaporation, the
matric suction increases. When water inltrates into
the soil (wetting or imbibition), the conditions reverse,
with the volumetric water content increasing and matric suction decreasing. Usually drainage and wetting

processes do not follow the same curve and the volumetric water content versus matric suction curves exhibit hysteresis during cycles of drainage and wetting
as shown in Fig. 7.12a. One cause of hysteresis is the
existence of ink bottle neck pores at the microscopic
scale as shown in Fig. 7.12b. Larger water-lled pores
can remain owing to the inability of water to escape
through smaller openings below in the case of drainage
or above in the case of evaporation. Another cause is
irreversible change in soil fabric and shrinkage during
drying.
The curves in Fig. 7.11 have two characteristic
pointsthe air entry pressure a and residual volumetric water content r as dened in Fig. 7.12a. The
entry pressure is the matric suction at which the air
begins to enter the pores and the pores become interconnected (Corey, 1994). At this point, the air permeability becomes greater than zero. Corey (1994)
also introduced the term displacement pressure (d
in Fig. 7.12b) and dened it as the matric suction at
which the rst water desaturation occurs during a
drainage cycle.10 The entry pressure is always slightly

10

The soilwater characteristic curve is referred to by a variety of


names depending on different disciplines. They include moisture retention, soilwater retention, specic retention, and moisture characteristic.

Copyright 2005 John Wiley & Sons

For the Dense NAPLwater two-phase system (often Dense NAPL


is the nonwetting uid and water is the wetting uid), the displacement pressure may be important to examine the potential of DNAPL
invading into a noncontaminated water-lled porous media.

Retrieved from: www.knovel.com

EFFECTIVE, INTERGRANULAR, AND TOTAL STRESS

Scanning
Curve

Suction

190

Hysteresis

Scanning
Curve
Initial drainage
Curve

a
d

Draining

Co
py
rig
hte
dM
ate
ria
l

Main Drying Curve

Main Wetting
Curve

Water Content

Wetting

r Residual Water Content

a Air Entry Value

d Displacement pressure

(b)

(a)

Figure 7.12 Hysteresis of a soilwater characteristic curve: (a) effect of hysteresis and (b)
ink bottle effect: a possible physical explanation for the hysteresis.

greater than the displacement pressure because pore


throats smaller than the maximum must be penetrated
to establish air connectivity. The air entry pressure is
much greater for ne-grained than for coarse-grained
soils because of their smaller pore sizes.
Residual water content r is dened as the water
content that cannot be further reduced by the increase
in matric suction. At this stage, the water phase
becomes essentially discontinuous and the regime
changes from the funicular to pendular state, as described in Section 7.4. However, this does not mean
that the soil cannot have a degree of saturation less
that the residual saturation because residual water can
continue to evaporate. Hence, it is important to note
that the residual saturation dened here is a mathematical tting parameter without a specic quantitative
value.
The shape of the soilwater characteristic curve depends on many factors, including the grain size distribution, soil fabric, the contact angle, and the interfacial
tension [see Eq. (7.11)]. If the material is uniform with
a narrow range of pore sizes, the curve has three distinct parts: a straight part up to the air entry pressure,
a relatively horizontal middle part, and an end part that
is almost vertical (soil 1 in Fig. 7.11). On the other
hand, if the material is well graded, the curve is
smoother (soils 3, 4, and 5 in Fig. 7.11). The capillary
pressure increases gradually as the water saturation decreases and the middle part is not horizontal. Many

Copyright 2005 John Wiley & Sons

algebraic formulas have been proposed to t the measured soil-water characteristic relations. The most popular ones are (a) the BrooksCorey (1966) equation:
m

 d

when   d

r
m r

(7.49)

1/

when   d

(7.50)

where m is the volumetric water content at full


saturation and is the curve-tting parameter called
the pore size distribution index and (b) the van Genuchten equation (1980):

 0

r
m r

1 / m

1m

(7.51)

where 0 and m are curve-tting parameters.


Various modications have been proposed to these
equations to include behaviors such as hysteresis, nonwetting uid trapping, and three-phase conditions.

7.12 EFFECTIVE STRESS IN UNSATURATED


SOILS

Although it seems clear that the volume change and


strength behavior of partly saturated soils are con-

Retrieved from: www.knovel.com

EFFECTIVE STRESS IN UNSATURATED SOILS

trolled by an effective stress that is not the same as the


total stress, the appropriate formulation for the effective stress is less certain than for a fully saturated soil.
As noted earlier, Bishop (1960) proposed Eq. (7.15)
(assuming  i ):
  ua  (ua uw)

(7.52)

Limitations in Bishops equation were highlighted


by Jennings and Burland (1962) in their experiments
investigating the volume change characteristics of unsaturated soils. Figure 7.14 shows that the oedometer
compression curve of air-dry silt falls above that of
saturated silt. Also, as shown in the gure, some airdry samples were consolidated at four different pressures (200, 400, 800, and 1600 kPa) and then soaked.

The term  ua is the net total stress. The term


ua uw represents the soil water suction that adds to
the effective stress since uw is negative. Thus, the
Bishop equation is appealing intuitively because negative pore pressures are known to increase strength and
decrease compressibility. Using Eq. (7.52), the shear
strength of unsaturated soil can be expressed as

0.80

0.76

Void Ratio e

Initially Soaked Test

0.72

Air Dry (8 specimens)

0.68

Soaked at Constant Void Ratio


Soaked at Constant Applied
Pressure

0.64
10

100

Figure 7.14 Oedometer compression curves of unsaturated


silty soils (after Jennings and Burland, 1962 in Leroueil and
Hight, 2002).

1. Compacted
Boulder Clay
2. Compacted Shale
3. Breadhead silt
4. Silt
5. Silty clay
6. Sterrebeek silt
7. White clay

(ua uw)
(ua uw)

0.55

(ua_uw)b = Air Entry Value

Degree of SaturationS(%)

(ua_uw)/(ua_uw)b

(a)

(b)

Figure 7.13 Variation of parameter  with the degree of water saturation Sr for different
soils: (a)  versus water saturation (after Gens, 1996) and (b)  versus suction (after Khalili
and Khabbaz, 1998).

Copyright 2005 John Wiley & Sons

1000

Applied Pressure (kPa )

Coefficient

(7.53)

where  is the effective friction angle of the soil.


However, difculties in the evaluation of the parameter
, its dependence on saturation ( 1 for saturated
soils and  0 for dry soils), and that the relationship
between  and saturation is soil dependent, as shown
in Fig. 7.13a, all introduce problems in the application
of Eq. (7.53). Since water saturation is related to matric
suction as described in Section 7.11, it is possible that
 depends on matric suction as shown in Fig. 7.13b.
Nonetheless, because of the complexity in determining
, the attempt to couple total stress and suction together into a single equivalent effective stress is uncertain (Toll, 1990).

Coefficient

0.84

Co
py
rig
hte
dM
ate
ria
l

 {( ua) (ua uw)}tan 

191

Retrieved from: www.knovel.com

192

EFFECTIVE, INTERGRANULAR, AND TOTAL STRESS

 a( ua) b(ua uw)

1.25

Void Ratio e

1.20
1.15
1.10
1.05
1.00
0.95
25

in which a and b are material parameters that may also


depend on degree of saturation and stress. For example, Fredlund et al. (1978) propose the following equation:
 ( ua)tan  (ua uw)tan  b

(7.54)

Preconsolidation
pressure

ua _ uw (kPa)
300 kPa

Curves are Averages of


Several Tests
50

100

200 kPa

100 kPa
0 kPa

200

(7.55)

where  b is the angle dening the rate of increase in


shear strength with respect to soil suction. An example
of this parameter as a function of water content, friction angle, and matric suction is given by Fredlund et
al. (1995).
Similarly, the change in void ratio e of an unsaturated soil can be given by (Fredlund, 1985)

Co
py
rig
hte
dM
ate
ria
l

The void ratio decreased upon soaking and the nal


state was very close to the compression curve of the
saturated silt. Additional tests in which constant volume during soaking was maintained by adjusting the
applied load were also done. Again, after equilibrium,
the state of soaked samples was close to the compression curve of the saturated silt. Soaking reduces the
suction and, hence, Bishops effective stress decreases.
This decrease in effective stress should be associated
with an increase in void ratio. However, the experimental observations gave the opposite trend (i.e., a decrease in void ratio is associated with irreversible
compression). The presence of meniscus water lenses
in the soil before wetting was stabilizing the soil structure, which is not taken into account in Bishops equation (7.52).
An alternative approach is to describe the shear
strength/deformation and volume change behavior of
unsaturated soil in terms of the two independent stress
variables  ua and ua uw (Coleman, 1962; Bishop
and Blight, 1963; Fredlund and Morgenstern, 1977;
Fredlund, 1985; Toll, 1990, Fredlund and Rahardjo,
1993; Tarantino et al., 2000). Figure 7.15 shows the
results of isotropic compression tests of compacted kaolin. Different compression curves are obtained for
constant suction conditions, and relative effects of 
ua and ua uw on volume change behavior can be
observed. Furthermore, the preconsolidation pressure
(or yield stress) increases with suction.
On this basis, the dependence of shear strength  on
stress is given by equations of the form

 at ( ua) am (ua uw)

(7.56)

where at is the coefcient of compressibility with respect to changes in  ua and am is the coefcient of
compressibility with respect to changes in capillary
pressure. A similar equation, but with different coefcients, can be written for change in water content.
For a partly saturated soil, change in water content and
change in void ratio are not directly proportional.
The two stress variables, or their modications that
include porosity and water saturation, have been used
in the development of elasto-plastic-based constitutive
models for unsaturated soils (e.g., Alonso et al., 1990;
Wheeler and Sivakumar, 1995; Houlsby, 1997; Gallipoli et al., 2003). The choice of stress variables is still
in debate; further details on this issue can be found in
Gens (1996), Wheeler and Karube (1996), Wheeler et
al. (2003), and Jardine et al. (2004).
Bishops  parameter in Eq. (7.52) is a scalar quantity, but microscopic interpretation of water distribution
in pores can lead to an argument that  is directional
dependent (Li, 2003; Molenkamp and Nazemi,
2003).11 During the desaturation process, the number
of soil particles under a funicular condition decreases,
and they change to a pendular condition with further
drying. For particles in the funicular region, the suction
pressure acts all around the soil particles like the water
pressure as illustrated in Fig. 7.4a. Hence, the effect is
isotropic even at the microscopic level. However, once
the microscopic water distribution of a particle changes
to the pendular condition, the capillary forces only act
on a particle at locations where water bridge forms and
the contribution to the interparticle forces becomes

400

_ ua (kPa)
11

Figure 7.15 Isotropic compression tests of compacted kaolin

(after Wheeler and Sivakumar, 1995 in Leroueil and Hight,


2002).

Copyright 2005 John Wiley & Sons

A microstructural analysis by Li (2003) suggests the following effective stress expression:


ij ij uaij ij (ua uw)

Retrieved from: www.knovel.com

QUESTIONS AND PROBLEMS

7.13

in the pendular regime) in the macroscopic effective


stress equations.

QUESTIONS AND PROBLEMS

1. A sand in the ground has porosity n of 0.42 and


specic gravity Gs of 2.6. It is assumed that these
values remain constant throughout the depth. The
water table is 4 m deep and the groundwater is under hydrostatic condition. The suctionvolumetric
water content relation of the sand is given by soil
1 in Fig. 7.11.
a. Calculate the saturated unit weight and dry unit
weight.
b. Evaluate the unit weights at different saturation
ratios Sw.
c. Plot the hydrostatic pore pressures with depth
down to a depth of 10 m and evaluate the saturation ratios above the water table.
d. Along with the hydrostatic pore pressure plot,
sketch the vertical total stress with depth using
the unit weights calculated in parts (a) and (b).
e. Estimate the vertical effective stress with depth.
Use Bishops equation (7.52) with  Sw. Comment on the result.

Co
py
rig
hte
dM
ate
ria
l

more or less point wise, as shown in Fig. 7.4b. As


described in Section 7.3, the magnitude of capillary
force depends on the size of the water bridge and the
separation of the two particles, and hence, the contact
force distribution in the particle assembly becomes dependent not only on pore size location and distribution
but also on the relative locations of particles to one
another (or soil fabric). It is therefore possible that the
distribution of the pendular-type capillary forces becomes directional dependent.
In clayey soils, water is attracted to clay surface by
electrochemical forces, creating large matric suction.
Although uw u0 is used in practice, the actual pore
pressure u acting at interparticle contacts may be different from u0, as discussed in Section 7.9. The contribution of the long-range interparticle forces to
mechanical behavior of unsaturated clayey soils remains to be fully evaluated.

CONCLUDING COMMENTS

The concepts in this chapter provide insight into the


meanings of intergranular pressure, effective stress,
and pore water pressure and the factors controlling
their values. Because soils behave as particulate materials and not as continua, knowledge of these stresses
and of the factors inuencing them is a necessary prerequisite to the understanding and quantication of
compressibility, deformation, and strength in constitutive relationships for behavior. Various interparticle
forces have been identied and their possible effects
on soil behavior are highlighted.
The effective stress in a soil is a function of its state,
which depends on the water content, density, and soil
structure. These factors are, in turn, inuenced by the
composition and ambient conditions. The relationships
between soil structure and effective stress are developed further in Chapter 8. Chemical, electrical, and
thermal inuences on effective pressures and uid
pressures in soils have not been considered in the developments in this chapter. They may be signicant,
however, as regards soil structure stability uid ow,
volume change, and strength properties. They are analyzed in more detail in subsequent chapters.
An understanding of the components of pore water
pressure is important to the proper measurement of
pore pressure and interpretation of the results. Inclusion of the effect of pore water suction and air or gas
pressure on the mechanical behavior of unsaturated
soils requires modication of the effective stress equation used for saturated soils. Complications arise from
the difculty in the choice of stress variables and in
treatment of contact-level forces (i.e., capillary forces

Copyright 2005 John Wiley & Sons

193

2. Repeat the calculations done in Question 1 with soil


5 in Fig. 7.11. The specic gravity of the soil is
2.65. Comment on the results by comparing them
to the results from Question 1.
3. Using Eq. (7.3), estimate the tensile strength of a
soil with different values of tensile strengths of cement, sphere, and interface. The soil has a particle
diameter of 0.2 mm and the void ratio is 0.7. Assume k/(1 e) 3.1. Consider the following two
cases: (a)  0.0075 mm and  5 and (b)
 0.025 and  30. Comment on the results.
4. Compute the following contact forces at different
particle diameters d ranging from 0.1 to 10 mm.
Comment on the results in relation to the effective
and intergranular pressure described in Section 7.9.
a. Weight of the sphere, W 61 Gs wd 3, where Gs
is the specic gravity (say 2.65) and w is the
unit weight of water.
b. Contact force by external load, N d 2, where
 is the external conning pressures applied.
The equation is approximate for a simple cubic
packing of equal size spheres (Santamarina,
2003). Consider two cases, (i)  1 kPa (
depth of 0.1 m) and (ii)  100 kPa ( depth
of 10 m).

Retrieved from: www.knovel.com

194

EFFECTIVE, INTERGRANULAR, AND TOTAL STRESS

c. Long-range van der Waals attraction force, A


Ahd/(24t 2), where Ah is the Hamaker constant
(Section 6.12) and t is the separation between
particles (Israelachvili, 1992, from Santamarina,
.
2003). Use Ah 1020 N-m and t 30 A

8. Clay particles in unsaturated soils often aggregate


creating matrix pores and intraaggregate pores. Air
exists in the matrix pores, but the intraaggregate
pores are often saturated by strong water attraction
to clay surfaces. The total potential of unsaturated
soil can be extended from Eq. (7.19) to  g
m s p, where p is the gas pressure potential.12 Discuss the values of each component of the
above equation in the matrix pores and the intraaggregate pores.

Co
py
rig
hte
dM
ate
ria
l

5. Discuss why it is difcult to measure suction using


a piezometer-type tensiometer for long-term monitoring of pore pressures. Describe the advantages of
other indirect measurement techniques such as porous ceramic and dielectric sensors.

7. Give a microscopic interpretation for why an unsaturated soil can collapse and decrease its volume
upon wetting as shown in Fig. 7.14 even though the
Bishops effective stress decreases.

6. For the following cases, compare the effective


stresses calculated by the conventional Terzaghis
equation and by the modied equation (7.39) with
values presented in Fig. 7.8. Discuss the possible
errors associated with effective stress estimation by
Terzaghis equation.
a. Pile foundation at a depth of 20 m.
b. A depth of 5 km from the sea level where the
subsea soil surface is 1 km deep.

Copyright 2005 John Wiley & Sons

12

This was proposed by a Review Panel in the Symposium on Moisture Equilibrium and Moisture Changes in Soils Beneath Covered
Areas in 1965.

Retrieved from: www.knovel.com

You might also like