SlideShare une entreprise Scribd logo
1  sur  22
Télécharger pour lire hors ligne
173 
CHAPTER 7 
Copyrighted Material 
Effective, Intergranular, and 
Total Stress 
7.1 INTRODUCTION 
The compressibility, deformation, and strength prop-erties 
of a soil mass depend on the effort required to 
distort or displace particles or groups of particles rel-ative 
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 de-pend 
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 ex-ample, 
by its water content, structure, density, or void 
ratio, reflects the influences of stresses applied in the 
past, and this further distinguishes soils from most 
other engineering materials, which, for practical pur-poses, 
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. For-tunately, 
however, the stresses, the state, and the prop-erties 
are not independent, and the relationships 
between stress and volume change, stress and stiffness, 
and stress and strength can be expressed in terms of 
definable soil parameters such as compressibility and 
friction angle. In soils with properties that are influ-enced 
significantly by chemical and physicochemical 
forces of interaction, other parameters such as cohe-sion 
may be needed. 
Most problems involving volume change, deforma-tion, 
and strength require separate consideration of the 
stress that is carried by the grain assemblage and that 
carried by the fluid phases. This distinction is essential 
because an assemblage of grains in contact can resist 
both normal and shear stress, but the fluid and gas 
phases (usually water and air) can carry normal stress 
but not shear stress. Furthermore, whenever the total 
head in the fluid phases within the soil mass differs 
from that outside the soil mass, there will be fluid flow 
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 fluid. 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 Engi-neering. 
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 
filled 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. 
Copyright © 2005 John Wiley  Sons Retrieved from: www.knovel.com
174 7 EFFECTIVE, INTERGRANULAR, AND TOTAL STRESS 
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 influence 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 meas-urable 
effects of a change of stress, such as compression, 
distortion and a change of shearing resistance are exclu-sively 
Copyrighted Material 
due to changes in the effective stresses 1, 2 and 
. Hence every investigation of the stability of a saturated 3 
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 com-puted 
using the external forces and the body force due 
to weight of the soil–water mixture. A pore water pres-sure, 
denoted herein by u0, can be measured at a point 
remote from the interparticle zone. The actual pore wa-ter 
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 is indeed equal to   u cannot be ascertained i 
without more detailed examination of all the interpar- 
1The 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 first three are the normal stresses 
and the latter three are the shear stresses. In this case, the effective 
stresses are defined as 11  11  u, 22  22  u, 33  33  
u,     ,     , and     . 12 12 23 23 31 31 
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 load-ing) 
(Fig. 7.1a)]. 
2. Particle Level Forces These include particle 
weight force, buoyancy force when a particle is 
submerged under fluid, and hydrodynamic forces 
or seepage forces due to pore fluid 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). 
When external forces are applied, both normal and 
tangential forces develop at particle contacts. All par-ticles 
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 in-terparticle 
skeletal forces in soils govern the macro-scopic 
stress–strain 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 photo-elastic 
model shown in Fig. 5.15. When the pores are 
filled with fluids, the weight of the fluids adds to the 
body force of the soil–fluids mixture. However, hydro-static 
pressure results from the fluid weight, and the 
uplift force due to buoyancy reduces the effective 
weight of a fluid-filled soil. This leads to smaller skel-etal 
forces for submerged soil compared to dry soil. 
Seepage forces that result from additional fluid pres-sures 
applied externally produce hydrodynamic forces 
on particles and alter the skeletal forces. 
7.4 INTERPARTICLE FORCES 
Long-range particle interactions associated with elec-trical 
double layers and van der Waals forces are dis- 
Copyright © 2005 John Wiley  Sons Retrieved from: www.knovel.com
INTERPARTICLE FORCES 175 
External Load 
Interparticle 
Forces 
Material 
Copyrighted Interparticle 
Forces 
(a) 
Body Force 
Buoyancy Force 
if Saturated 
Viscous Drag by 
Seepage Flow 
Seepage 
(b) 
Capillary Force or 
Cementation-reactive 
Force 
Electrical Forces 
(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 
flocculation–deflocculation 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 impor-tant 
as well since they may influence the intergranular 
stresses and control the strength at interparticle con-tacts, 
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 pres-sure 
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 sufficiently great to prevent the interpenetra-tion 
of matter. 
At separation distances beyond the region of direct 
physical interference between adsorbed ions and hy-dration 
water molecules, double-layer interactions pro-vide 
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 fluid. 
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 
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 sufficient to 
squeeze out all the water between parallel particle sur-faces 
in naturally occurring clays. Heat and/or high 
vacuum are needed to remove all the water from a fine-grained 
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 pro-portion 
( 1%) of the total soil cross-sectional area 
in most cases. The exact nature of an interparticle con-tact 
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 dis-tance. 
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 differ-ent 
potentials has been suggested as a cause. When the 
Copyright © 2005 John Wiley  Sons Retrieved from: www.knovel.com
176 7 EFFECTIVE, INTERGRANULAR, AND TOTAL STRESS 
gap between parallel particle surfaces separated by dis-tance 
d at potentials V1 and V2 is conductive, there is 
an attractive force per unit area, or tensile strength, 
given by (Ingles, 1962) 
4.4  106 (V  V )2 F  1 2 N/m2 (7.1) d2 
Copyrighted Material 
where F is the tensile strength, d is in micrometers, 
and V1 and V2 are in millivolts. This force is indepen-dent 
of particle size and becomes significant (greater 
than 7 kN/m2 or 1 psi) for separation distances less 
than 2.5 nm. 
Electromagnetic Attractions Electromagnetic at-tractions 
caused by frequency-dependent dipole inter-actions 
(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 
specifically for fine-grained soils with various geomet-ric 
parameters such as particle length, thickness, ori-entation, 
and spacing. 
Primary Valence Bonding Chemical interactions 
between particles and between the particles and adja-cent 
liquid phase can only develop at short range. Co-valent 
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 hydro-gen 
bonds, can develop at interparticle contacts with-out 
the presence of cementing agents is largely a 
matter of speculation. Very high contact stresses be-tween 
particles could squeeze out adsorbed water and 
cations and cause mineral surfaces to come close to-gether, 
perhaps providing opportunity for cold weld-ing. 
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 (Chap-ter 
11). Thus, interatomic bonding between particles 
seems possible. On the other hand, the absence of co-hesion 
in overconsolidated silts and sands argues 
against such pressure-induced bonding. 
Cementation Cementation may develop naturally 
from precipitation of calcite, silica, alumina, iron ox-ides, 
and possibly other inorganic or organic com-pounds. 
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 ten-sile 
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 ce-ment– 
particle interface. The following equation can be 
derived (Ingles, 1962) for the tensile strength T per 
unit area of soil cross section: 
 1  n 
  Pk(7.2) T 1  e n 
 Ai 
1 
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 
Pk 
  (7.3) T d2(1  e) 
For a random and isotropic assembly of rods of length 
l and diameter d 
Pk 
  (7.4) T d(l  d/2)(1  e) 
Bond strength P is evaluated in the following way (Fig. 
7.2) for two cemented spheres of radius R. It may be 
shown that 
(R  cos ) 
 cosh  R sin  (7.5) 
 
so for known ,  can be computed. Then, for cement 
failure, 
P    2 (7.6) c 
where c is the tensile strength of the cement; for 
sphere failure, 
P    ()2 (7.7) s 
where   R sin , and s is the tensile strength of 
the sphere, and for failure at the interface 
Copyright © 2005 John Wiley  Sons Retrieved from: www.knovel.com
INTERPARTICLE FORCES 177 
Copyrighted Material 
Figure 7.2 Contact zone failures for cemented spheres. 
sin  P     2R2(1  cos ) (7.8) 1  
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. Artificially ce-mented 
soils are often loaded after cementation has 
developed, whereas cementation develops during or af-ter 
overburden loading in natural soils. In the former 
case, the particles and cementation bonding are loaded 
together and contact forces can become negative de-pending 
on the tensile resistance of cementation bond-ing. 
The distribution and magnitude of skeletal forces 
are therefore influenced by both geometric arrange-ment 
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 de-veloped 
before cementation coats the already loaded 
particles. In this case, it is possible that cementation 
creates extra forces at particle contacts. In some ce-mented 
natural materials, if the soil is unloaded from 
high overburden stress, elastic rebound may disrupt ce-mented 
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 fluid 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 tem-porary 
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 equa-tion, 
assuming a capillary tube as shown in Fig. 7.3a: 
2 cos
aw ˆP (7.9) c rp 
where is the capillary pressure at air entry, ˆc 
P 
is aw the air–water interfacial tension,
is contact angle de-fined 
in Fig. 7.3, and rp is the tube radius. For pure 
water and air, aw depends on temperature, for exam-ple, 
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 then air in- ˆP 
, c 
vades the pore.2 Since soil has pores with various sizes, 
the water in the largest pores is displaced first followed 
by smaller pores. This leads to a macroscopic model 
of the soil–water characteristic curve (or the capillary 
pressure–saturation relationship), as discussed in Sec-tion 
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 infiltration when air in the pores cannot escape or the air 
boundary is completely blocked. 
Copyright © 2005 John Wiley  Sons Retrieved from: www.knovel.com
178 7 EFFECTIVE, INTERGRANULAR, AND TOTAL STRESS 
Capillary Tube 
Representing a Pore 
2rp 
dc 
Material 
Copyrighted θ 
ua 
uw 
^ 2σaw cosθ 
Pc = ρw gdc = 
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. 
2r2  cos
F  r2 Pˆ  aw (7.10) c c rp 
where rp is the size of the pore into which the air has 
entered. Since the fluid acts like a membrane with neg-ative 
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 be-comes 
disconnected and remains in the form of me-nisci 
or liquid bridges at the interparticle contacts 
[termed the ‘‘pendular’’ regime by Bear (1972)]. The 
curved air–water interface produces a pore water ten-sion, 
which, in turn, generates interparticle compres-sive 
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 sep-aration 
between the two particles, the radius of the liq-uid 
bridge, interfacial tension, and contact angle (Lian 
et al., 1993). Once the water phase becomes discontin-uous, 
evaporation and condensation are the primary 
mechanisms of water transfer. Hence, the humidity of 
the gas phase and the temperature affect the water va-por 
pressure at the surface of water menisci, which in 
turn influences the air pressure ua. 
7.5 INTERGRANULAR PRESSURE 
Several different interparticle forces were described in 
the previous section. Quantitative expression of the in-teractions 
of all these forces in a soil is beyond the 
present state of knowledge. Nonetheless, their exis-tence 
bears directly on the magnitude of intergranular 
pressure and the relationship between intergranular 
pressure and effective stress as defined by   
  u. 
A simplified 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 x–x is con-sidered. 
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. Define area a as 
3Note that only vertical forces at the contact are considered in this 
simplified analysis. It is evident, however, that applied boundary nor-mal 
and shear stresses each induce both normal and shear forces at 
interparticle contacts. These forces contribute both to the develop-ment 
of soil strength and resistance to compression and to the slip-ping 
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. 
Copyright © 2005 John Wiley  Sons Retrieved from: www.knovel.com
INTERGRANULAR PRESSURE 179 
Material 
(a) 
Copyrighted Soil Particles 
Continuous 
Water Film 
Negative pore pressure acts all 
around the particles 
Suction forces act only at particle 
contacts and the magnitude of the 
forces depends on the size of liquid 
bridges. 
(b) 
Liquid 
Bridges 
Soil Particles 
Interparticle 
Forces 
Pores of Radius 
rp Filled with Air 
Air 
Figure 7.4 Microscopic water–soil interaction in unsaturated soils: (a) funicular regime and 
(b) pendular regime. 
Figure 7.6 Forces acting on interparticle contact A. 
Figure 7.5 Surfaces through a soil mass. 
the average total cross-sectional area along a horizontal 
plane served by the contact. It equals the total hori-zontal 
area divided by the number of interparticle con-tacts 
along the wavy surface. The forces acting on area 
a in Fig. 7.6 are: 
1. a, the force transmitted by the applied stress , 
which includes externally applied forces and 
body weight from the soil above. 
Copyright © 2005 John Wiley  Sons Retrieved from: www.knovel.com
180 7 EFFECTIVE, INTERGRANULAR, AND TOTAL STRESS 
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 long-range 
attractive stress A, that is, van der Waals 
and electrostatic attractions. 
4. Aac, the force developed by the short-range at-tractive 
stress A, resulting from primary valence 
Material 
Copyrighted (chemical) bonding and cementation. 
5. Cac, the intergranular contact reaction that is gen-erated 
by hydration and Born repulsion. 
Vertical equilibrium of forces requires that 
a  Aa  Aa  ua  Ca (7.11) c c 
Division of all terms by a converts the forces to 
stresses per unit area of cross section, 
ac   (C  A)  u  A (7.12) 
a 
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 me-chanics 
usage. Designation of this term by gives i 
    A  u (7.13) i 
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 con-dition 
that 
a  a  a i.e., a → 0 w a c 
The resulting equation is 
aw     A  u  (u  u ) (7.14) i a a w a 
In the absence of significant long-range attractions, 
this equation is similar to that proposed by Bishop 
(1960) for partially saturated soils 
    u  (u  u ) (7.15) i a a w 
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 sat-uration. 
Further discussion of the effective stress con-cept 
for unsaturated soils is given in Section 7.12. 
Limiting the discussion to saturated soils, two ques-tions 
arise: 
1. How does the intergranular pressure relate to i 
the effective stress as defined for most analyses, 
that is,     u? 
2. How does the intergranular pressure relate to i 
the measured quantity,    u0m , that is taken 
as the effective stress, recalling (Section 7.2) that 
pore pressure can only be measured at points out-side 
the true interparticle zone? 
Answers to these questions require a more detailed 
consideration of the meaning of fluid pressures in soils. 
7.6 WATER PRESSURES AND POTENTIALS 
Pressures in the pore fluid of a soil can be expressed 
in several ways, and the total pressure may involve 
several contributions. In hydraulic engineering, prob-lems 
are analyzed using Bernoulli’s equation for the 
total heads and head losses associated with flow be-tween 
two points, that is, 
p v2 p v2 Z  1  1  Z  2  2  
h (7.16) 1 2 1–2 	 2g 	 2g w w 
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 flow velocities at points 1 and 2, 
	w is the unit weight of water, g is the acceleration due 
to gravity, and 
h1–2 is the loss in head between points 
1 and 2. The total head H (dimension L) is 
p v2 
H  Z  (7.17) 
	 2g w 
Flow results only from differences in total head; 
conversely, if the total heads at two points are the 
same, there can be no flow, even if Z1 	 Z2 and p1 	 
p2. If there is no flow, there is no head loss and 
h1–2 
 0. 
The flow velocity through soils is low, and as a re-sult 
v2/2g → 0, and in most cases it may be neglected. 
Therefore, the relationship 
Copyright © 2005 John Wiley  Sons Retrieved from: www.knovel.com
WATER PRESSURE EQUILIBRIUM IN SOIL 181 
p p Z  1  Z  2  
h (7.18) 1 2 1–2 	w 	w 
is the basis for evaluation of pore pressures and anal-ysis 
of seepage through soils and other porous media. 
Although the absence of velocity terms is a factor 
that seems to simplify the analysis of flows and pres-sures 
in soils, there are other considerations that tend 
Copyrighted Material 
to complicate the problem. These include: 
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 wa-ter. 
3. Compositional differences in the water from 
point-to-point and adsorptive force fields from 
particle surfaces. 
4. Differences in interparticle forces and the energy 
state of the pore fluid from point to point owing 
to thermal, electrical, and chemical gradients. 
Such gradients can cause fluid flows, deforma-tions, 
and volume changes, as considered in more 
detail in Chapter 9. 
Some formalism in definition and terminology is 
necessary to avoid confusion. The status of water in a 
soil can be expressed in terms of the free energy rel-ative 
to free, pure water (Aitchison, et al., 1965). The 
free energy can be (and is) expressed in different ways, 
including 
1. Potential (dimensions—L2T2: J/kg) 
2. Head (dimensions—L: m, cm, ft) 
3. Pressure (dimensions—ML1 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 definition of total po-tential 
is the work per unit quantity to transport re-versibly 
and isothermally an infinitesimal amount of 
pure water from a pool at a specified elevation at at-mospheric 
pressure to the point in soil water under 
consideration. 
The selection of the components of the total poten-tial 
 (total head H, total pressure P) is somewhat 
arbitrary (Bolt and Miller, 1958); however, the follow-ing 
have gained acceptance for geotechnical work 
(Aitchison, et al., 1965): 
1. Gravitational potential g (head Z, pressure pz) 
corresponds to elevation head in normal hydrau-lic 
usage. 
2. Matrix or capillary potential m (head hm, pres-sure 
p) is the work per unit quantity of water to 
transport reversibly and isothermally an infinites-imal 
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 consider-ation 
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 pres-sures 
generated by capillarity menisci, and from 
water adsorption forces exerted by particle sur-faces. 
A piezometer measures the matrix poten-tial 
if it contains fluid of the same composition 
as the soil water. 
3. Osmotic (or solute) potential s (head hs, pres-sure 
ps) is the work per unit quantity of water to 
transport reversibly and isothermally an infinites-imal 
quantity of water from a pool of pure water 
at a specified elevation and atmospheric pressure 
to a pool containing a solution identical in com-position 
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 wa-ter, 
and it depends on the composition and ability 
of the soil particles to restrain the movement of 
adsorbed cations. The osmotic potential is nega-tive, 
that is, water tends to flow in the direction 
of increasing concentration. 
The total potential, head, and pressure then become 
       (7.19) g m s 
H  Z  h  h (7.20) m s 
P  p  p  p (7.21) z s 
At equilibrium and no flow there can be no varia-tions 
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 
Copyright © 2005 John Wiley  Sons Retrieved from: www.knovel.com
182 7 EFFECTIVE, INTERGRANULAR, AND TOTAL STRESS 
Material 
Copyrighted 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, 
H  0  h  0  h 0 m0 m0 
It follows that 
P  h 	  u (7.22) 0 m0 w 0 
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, 
H  0  h  h 1 m1 s1 
If no water is flowing, H1  H0, and 
h  h  h m1 s1 m0 
Also, because p1  p0  u0 
u  h 	  h 	 (7.23) 0 m1w s1 w 
At point 2, which is between the same two clay par-ticles 
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, 
u h  h  h  h  h  0 m2 s2 m1 s1 m0 	w 
A similar analysis could be applied to any point in the 
system. If point 3 were midway between two clay par-ticles 
spaced the same distance apart as the particles 
on either side of point 1, then hs3  hs1, but Z3 	 0. 
Thus, 
u0  Z  h  h  Z  h  h (7.24) 3 m3 s3 3 m3 s1 	w 
A partially saturated system can also be analyzed, 
but the influences of curved air–water 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: 
1. As the osmotic and gravitational components 
vary from point to point in a soil at equilibrium, 
Copyright © 2005 John Wiley  Sons Retrieved from: www.knovel.com
MEASUREMENT OF PORE PRESSURES IN SOILS 183 
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 vi-sualize. 
Nonetheless, this underlies the whole 
concept of water flow by chemical osmosis. 
Copyrighted Material 
2. The total potential, head, and pressure are meas-urable, 
and separation into components is possi-ble 
experimentally, although it is difficult. 
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 par-ticles, 
u0 is seen to include contributions from 
osmotic pressures as well as matrix pressures. 
Since osmotic pressures are the cause of long-range 
repulsions due to double-layer interactions, 
measured pore water pressures may include con-tributions 
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 lab-oratory 
use, whereas others are intended for use in the 
field. 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 pi-ezometer 
communicates with the soil through a 
porous stone or filter. Pressures are determined 
from the water level in a standpipe, by a manom-eter, 
by a pressure gauge, or by an electronic 
pressure transducer. A piezometer used to mea-sure 
pressures less than atmospheric is usually 
termed a tensiometer. 
2. Gypsum Block, Porous Ceramic, and Filter 
Paper The electrical properties across a spe-cially 
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 deter-mines 
the amount of moisture in the block 
(Whalley et al., 2001). The same principle can be 
applied by placing a dry filter paper on a soil 
specimen and allowing the soil moisture to ab-sorb 
into the paper. When the suction in the filter 
paper is equal to the suction in the soil, the two 
reach equilibrium, and the suction can be deter-mined 
by the water content of the filter paper. 
These techniques are used for measurement of 
pore pressures less than atmospheric. 
3. 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 con-tent 
and pressure is used to establish the relation-ship 
between soil suction and water content. 
4. Consolidation Tests The consolidation pressure 
on a sample at equilibrium is the soil water suc-tion. 
If the consolidation pressure were instanta-neously 
removed, then a negative water pressure 
or suction of the same magnitude would be 
needed to prevent water movement into the soil. 
5. Vapor Pressure Methods The relationship be-tween 
relative humidity and water content is used 
to establish the relationship between suction and 
water content. 
6. Osmotic Pressure Methods Soil samples are 
equilibrated with solutions of known osmotic 
pressure to give a relationship between water 
content and water suction. 
7. Dielectric Sensors Such as Capacitance Probes 
and Time Domain Reflectometry 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 soil’s capacitance, which is related to dielec-tric 
constants of soil particle, water, and air. The 
capacitance is largely influenced by water con-tent, 
as the dielectric constant of water is large 
compared to the dielectric constants of soil 
particle and air. Time domain reflectrometry 
measures the travel time of a high-frequency, 
electromagnetic pulse. The presence of water in 
the soil slows down the speed of the electromag-netic 
wave by the change in the dielectric prop-erties. 
Volumetric water content can therefore be 
indirectly measured from the travel time mea-surement. 
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 
Copyright © 2005 John Wiley  Sons Retrieved from: www.knovel.com
184 7 EFFECTIVE, INTERGRANULAR, AND TOTAL STRESS 
descriptions and comparisons of these and other meth-ods 
are given by Croney et al. (1952), Aitchison et al. 
(1965), Richards and Peter (1987), and Ridley et al. 
(2003). 
7.9 EFFECTIVE AND INTERGRANULAR 
PRESSURE 
In Section 7.5, it was shown that the intergranular pres-sure 
Copyrighted Material 
is given by 
    A  u (7.25) i 
where u is the hydrostatic pressure between particles 
(or hm	w in the terminology of Section 7.7). General-ized 
forms of Eq. (7.24) are 
u  Z	  h 	  h 	 (7.26) 0 w mw sw 
and 
u  h 	  u  Z	  h 	 (7.27) m w 0 w sw 
Thus, Eq. (7.25) becomes, for the case of no elevation 
difference between a piezometer and the point in ques-tion 
(i.e., Z  0), 
    A  u  h 	 (7.28) i 0 s w 
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 
reflects 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 
    A  u  R (7.29) i 0 
From Eq. (7.25), it was seen that the intergranular 
pressure was dependent on long-range interparticle at-tractions 
A as well as on the applied stress  and the 
pore water pressure between particles u. Equation 
(7.29) indicates that if intergranular pressure is to i 
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 par-ticles 
u  u0  R includes the effects of long-range 
repulsions as required by the condition of constant to-tal 
potential for equilibrium. 
In the general case, therefore, the true intergranular 
pressure    A  u0  R and the conventionally i 
defined effective stress     u0 differ by the net 
interparticle stress due to physicochemical contribu-tions, 
    A  R (7.30) i 
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 signifi-cantly 
different magnitude, would the intergranular and 
effective stress be significantly different. Such a con-dition 
appears not to be common, although it might be 
of importance in a well-dispersed sodium montmoril-lonite, 
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 equi-librium, 
with contributing forces parallel to each other, 
that is, the intergranular stress is the sum of the i skeletal forces (defined as     u0) and the elec-trochemical 
stress (A  R), as illustrated in Fig. 7.8a. 
This implies that the deformation induced by the elec-trochemical 
stress (A  R) is equal to the deformation 
induced by the skeletal forces at contacts [i.e., a ‘‘par-allel’’ 
model as described by Hueckel (1992)]. The 
change in pore fluid chemistry at constant confinement 
() leads to changes in intergranular stresses (), re- i 
sulting 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): 
  R  A      u (7.31) i 0 
This is called the ‘‘series’’ model (Hueckel, 1992), and 
the model can be applicable for very fine 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 or ef- i 
fective stress  changes the interparticle spacing, 
which may contribute to changes in strength properties 
upon shearing. 
4A detailed analysis of effective stress in clays is presented by Chat-topadhyay 
(1972), which leads to similar conclusions, including Eq. 
(7.29). was termed the true effective stress and it governed the i 
volume change behavior of Na–montmorillonite. 
Copyright © 2005 John Wiley  Sons Retrieved from: www.knovel.com
ASSESSMENT OF TERZAGHI’S EQUATION 185 
Electrochemical Force Electrochemical Force 
Skeletal Force 
σ = σ _ u0 
Material 
Copyrighted Skeletal Force 
Electrochemical 
Force A _ R 
σi 
σi 
Deformation at 
the Contact 
σi = σ _ u0 + A _ R 
(a) 
Skeletal Force 
σi 
Skeletal Force 
σ = σ _ u0 
Total Deformation 
at the Contact σi 
σi = σ _ u0 = A _ R 
(b) 
Particle Deformation 
by Skeletal Force 
Electrochemical 
Force A _ R 
Electrochemical Force 
Skeletal Force Electrochemical Force 
Skeletal Force 
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 ex-ternal 
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]. 
7.10 ASSESSMENT OF TERZAGHI’S EQUATION 
The preceding equations and discussion do not confirm 
that Terzaghi’s 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 ap-plication 
in practice. Skempton (1960b) showed that 
the Terzaghi equation does not give the true effective 
stress but gives an excellent approximation for the case 
of saturated soils. Skempton proposed three possible 
relationships for effective stress in saturated soils: 
1. The true intergranular pressure for the case when 
A  R  0 
    (1  a )u (7.32) c 
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 
  k   tan  (7.33) i 
where  is an intrinsic friction angle and k is a 
true cohesion. The following relationships were 
derived: For shear strength, 
ac tan      1  u (7.34) tan  
where  is the effective stress angle of shearing 
resistance. For volume change, 
Copyright © 2005 John Wiley  Sons Retrieved from: www.knovel.com
186 7 EFFECTIVE, INTERGRANULAR, AND TOTAL STRESS 
Table 7.1 Compressibility Values for Soil, Rock, 
and Concrete 
Material 
Compressibilitya 
per kN/m2  106 
C Cs Cs /C 
Quartzitic sandstone 0.059 0.027 0.46 
Quincy granite (30 m deep) 0.076 0.019 0.25 
Vermont marble 0.18 0.014 0.08 
Concrete (approx.) 0.20 0.025 0.12 
Dense sand 18 0.028 0.0015 
Loose sand 92 0.028 0.0003 
London clay (over cons.) 75 0.020 0.00025 
Gosport clay (normally cons.) 600 0.020 0.00003 
After Skempton (1960b). 
aCompressibilities at p  98 kN/m2; water Cw  0.49 
 106 per kN/m2. 
    1  Csu (7.35) C 
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) 
Copyrighted Material 
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 

V 
 C 
 (7.37) 
V 
and also satisfied the Coulomb equation for drained 
shear strength d : 
  c   tan  (7.38) d 
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 Skempton’s 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 negli-gible, 
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 gen-eral 
form (Lade and de Boer, 1997): 
    u (7.39) 
where  is the fraction of the pore pressure that gives 
the effective stress.5 Different expressions for  pro-posed 
by several researchers are listed in Table 7.2. 
5A more general expression has been proposed as ij ijij u, 
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; Did-wania, 
2002). 
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 incre-ment 

u and that due to the change in confining pres-sure 

(  u) (or 
  
u). The effective stress 
increment 
 is defined as the stress that produces the 
same volume change, 
CV 
 
 
V  
V  CV (
  
u) 0 sks sku 0 
 C V 
u (7.40) u 0 
where 
Vsks is the volume change of soil skeleton due 
to change in confining 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 confining pressure change, and Cu is 
the compressibility of the soil skeleton by pore pres-sure 
change. Rearranging Eq. (7.40) leads to 
Cu 
  
  1   
u (7.41) C 
Lade and de Boer (1997) used this equation to de-rive 
an effective stress equation for granular materials 
under drained conditions. Consider a condition in 
which the total confining pressure is constant [
(  
Copyright © 2005 John Wiley  Sons Retrieved from: www.knovel.com
ASSESSMENT OF TERZAGHI’S EQUATION 187 
Table 7.2 Expressions for  to Define Effective Stress 
Pore Pressure Fraction  Note Reference 
1 Terzaghi (1925b) 
n n porosity Biot (1955) 
1  ac ac  grain contact area per unit area of plane Skempton and Bishop (1954) 
1  ac 
tan  
tan  
Equation (7.34) Skempton (1960b) 
1  
CEquation (7.35); for isotropic elastic 
s 
C 
deformation of a porous material; for solid 
rock with small interconnected pores and 
low porosity (Lade and de Boer, Material 
1997) 
Copyrighted Biot and Willis (1957), Skempton 
(1960b), Nur and Byerlee (1971), Lade 
and de Boer (1997) 
1  (1  n) 
Cs 
C 
Equation (7.43) Suklje (1969); Lade and de Boer (1997) 
After Lade and de Boer (1997). 
Figure 7.9 Variation of  with stress for quartz sand and 
gypsum sand (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 vol-umetric 
compression of the solid grains (
Vgu). Hence, 

V 
 C V 
u  C (1  n)V 
u 
 
V or sku u 0 s 0 gu 
C  C (1  n) u s (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) Cs 
u or C 
Cs   
1  (1  n)  (7.43) C 
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 essen-tially 
one. Thus, Terzaghi’s effective stress equation, 
while not rigorously correct, is again shown to be an 
excellent approximation in almost all cases for satu-rated 
soils (i.e., solid grains and pore fluid are consid-ered 
to be incompressible compared to soil skeleton 
compressibility). 
6An example of this condition is a soil under a seabed, in which the 
sea depth varies. This condition is often called the ‘‘unjacked con-dition.’’ 
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 mix-ture 
theory, the total stress increment (
iso) is sepa-rated 
into partial stress increments for the solid phase 
(
s) and the fluid phase (
ƒ) (Oka, 1996). Consid-ering 
that the macroscopic volumetric strains by two 
phases are equal but of opposite sign for undrained 
Copyright © 2005 John Wiley  Sons Retrieved from: www.knovel.com
188 7 EFFECTIVE, INTERGRANULAR, AND TOTAL STRESS 
θ 
Solid Surface 
Material 
θ 
Copyrighted Air 
Water 
(reference fluid) 
(a) 
Air 
Water 
(reference fluid) 
Solid surface 
(b) 
(c) 
Air 
Water 
Solid 
Figure 7.10 Wettability of two fluids (water and air) on a 
solid surface: (a) contact angle less than 90, (b) contact an-gle 
more than 90, and (c) unsaturated sand with water as the 
wetting fluid and air as the nonwetting fluid. 
conditions, Oka (1996) showed that the partial stresses 
are related to the total stress as follows: 
C  Cs 
ƒ  
iso (C/n)  (1  1/n)C  C s l (7.44) 
[(1/n)  1]C  (C /n)  C s l 
s  
iso (C/n)  (1  1/n)C  C s 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 fluid. 
If the excess pore pressure generated by undrained 
isotropic loading 
 is 
u, the partial stress increment 
for the fluid phase becomes (Oka, 1996) 

  n 
u (7.45) ƒ 
Combining Eqs. (7.45) and (7.46), 
C  Cs 
u  
 (7.46) C  C  n(C  C ) iso s l s 
The multiplier in the right-hand side of the above 
equation is in fact Bishop’s pore water pressure coef-ficient 
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 
Terzaghi’s effective stress equation is also applicable 
for undrained conditions for most soils. 
7.11 WATER–AIR INTERACTIONS IN SOILS 
Wettability refers to the affinity of one fluid for a solid 
surface in the presence of a second or third fluid or 
gas. A measure of wettability is the contact angle, 
which was introduced in Eq. (7.9). Figure 7.10 illus-trates 
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 fluid (air for Fig. 7.10a and 
water for 7.10b). The interface between the two fluids 
meets the solid surface at a contact angle
. If the angle 
is less than 90, the reference fluid is referred to as the 
wetting fluid for a given solid surface. If the angle is 
greater than 90, the reference liquid is referred to as 
the nonwetting phase. The figure shows that water and 
7A similar equation for B value has been proposed by Lade and de 
Boer (1997). 
air are the wetting and nonwetting fluid, respectively.8 
The environmental SEM photos in Fig. 5.27 showed 
that water can be either wetting or nonwetting fluid 
depending soil mineralogy. 
The contact angle is a property related to interac-tions 
of solid and two fluids (water and air, in this 
case). 
as  ws cos
(7.47) 
aw 
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 
8Some 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 fluid with respect to NAPL 
and that NAPL is a wetting fluid with respect to air, implying that 
the wettability order is water  NAPL  air. Below the water table, 
water is the wetting fluid and NAPL is the nonwetting fluid. 
Copyright © 2005 John Wiley  Sons Retrieved from: www.knovel.com
WATER–AIR INTERACTIONS IN SOILS 189 
106 
105 
104 
Material 
103 
102 
101 
100 
Copyrighted 7 
5 
3 
2 
0.0 0.1 0.2 0.3 0.4 0.5 0.6 
Volumetric Water Content θw 
10-1 
1 
4 
6 
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 
Matric suction ua – uw (kPa) 
Figure 7.11 Soil–water characteristic curves for some Dutch 
soils (from Koorevaar et al., 1983; copied from Fredlund and 
Rahardjo, 1993). 
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 sur-faces. 
It has been observed for movement of water rel-ative 
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 sepa-rating 
two fluids. This pressure difference exists be-cause 
of the interfacial tension present in the fluid– 
fluid interface. Matric suction is a property that causes 
porous media to draw in the wetting fluid and repel 
the nonwetting fluid and is defined as the difference 
between the nonwetting fluid pressure and the wetting 
fluid pressure. For a two-phase system consisting of 
water and air, the matric suction  is 
  u  u (7.48) n w 
where un is the pressure of the nonwetting fluid (air) 
and uw is the pressure of the wetting fluid (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 propor-tional 
to the interfacial tension and inversely propor-tional 
to the radius of curvature. It follows that higher 
air pressure is required for air to enter water-saturated 
fine-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 soil–water characteristic relationship in 
which the matric suction is plotted against the volu-metric 
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 water-saturated 
soil. As the volumetric water content de-creases, 
as a result of drainage or evaporation, the 
matric suction increases. When water infiltrates into 
the soil (wetting or imbibition), the conditions reverse, 
with the volumetric water content increasing and ma-tric 
suction decreasing. Usually drainage and wetting 
9The soil–water characteristic curve is referred to by a variety of 
names depending on different disciplines. They include moisture re-tention, 
soil–water retention, specific retention, and moisture char-acteristic. 
processes do not follow the same curve and the volu-metric 
water content versus matric suction curves ex-hibit 
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-filled 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 
points—the air entry pressure a and residual volu-metric 
water content

Contenu connexe

Tendances

Settlement of shallow foundations
Settlement of shallow foundationsSettlement of shallow foundations
Settlement of shallow foundationsAmirmasoud Taghavi
 
Understanding Permeability of Hydraulic Fracture Networks A Sandbox Analog Mo...
Understanding Permeability of Hydraulic Fracture Networks A Sandbox Analog Mo...Understanding Permeability of Hydraulic Fracture Networks A Sandbox Analog Mo...
Understanding Permeability of Hydraulic Fracture Networks A Sandbox Analog Mo...Renee Heldman
 
Effect of soil structure interaction on high rise r.c regular frame structur...
Effect of soil  structure interaction on high rise r.c regular frame structur...Effect of soil  structure interaction on high rise r.c regular frame structur...
Effect of soil structure interaction on high rise r.c regular frame structur...eSAT Journals
 
Structure soil-structure interaction literature review
Structure soil-structure interaction literature reviewStructure soil-structure interaction literature review
Structure soil-structure interaction literature reviewÁngel J. Alicea, EIT
 
Goetech. engg. Ch# 03 settlement analysis signed
Goetech. engg. Ch# 03 settlement analysis signedGoetech. engg. Ch# 03 settlement analysis signed
Goetech. engg. Ch# 03 settlement analysis signedIrfan Malik
 
6 effective stress capillarity_permeability
6 effective stress capillarity_permeability6 effective stress capillarity_permeability
6 effective stress capillarity_permeabilitySaurabh Kumar
 
Seismic ssi effects and liquification
Seismic ssi effects and liquificationSeismic ssi effects and liquification
Seismic ssi effects and liquificationArpan Banerjee
 
Seismic Behaviour of Multi-Storied Building by Using Tuned Mass Damper and Ba...
Seismic Behaviour of Multi-Storied Building by Using Tuned Mass Damper and Ba...Seismic Behaviour of Multi-Storied Building by Using Tuned Mass Damper and Ba...
Seismic Behaviour of Multi-Storied Building by Using Tuned Mass Damper and Ba...IJERA Editor
 
Soil Structure Interaction
Soil Structure InteractionSoil Structure Interaction
Soil Structure InteractionAmmar Motorwala
 
Deformation activity
Deformation activityDeformation activity
Deformation activityDNHSGRADE8
 
Case study on effect of water table on bearing capacity
Case study on effect of water table on bearing capacityCase study on effect of water table on bearing capacity
Case study on effect of water table on bearing capacityAbhishek Mangukiya
 
INFLUENCE OF SOIL-STRUCTURE INTERACTION ON RESPONSE OF A MULTI-STORIED BUILDI...
INFLUENCE OF SOIL-STRUCTURE INTERACTION ON RESPONSE OF A MULTI-STORIED BUILDI...INFLUENCE OF SOIL-STRUCTURE INTERACTION ON RESPONSE OF A MULTI-STORIED BUILDI...
INFLUENCE OF SOIL-STRUCTURE INTERACTION ON RESPONSE OF A MULTI-STORIED BUILDI...SJ BASHA
 
Soil Amplification and Soil-Structure Interaction
Soil Amplification and Soil-Structure InteractionSoil Amplification and Soil-Structure Interaction
Soil Amplification and Soil-Structure InteractionBappy Kamruzzaman
 
Soil structure interaction amec presentation-final
Soil structure interaction amec presentation-finalSoil structure interaction amec presentation-final
Soil structure interaction amec presentation-finalAhmad Hallak PEng
 
Soil structure interaction effect on dynamic behavior of 3 d building frames ...
Soil structure interaction effect on dynamic behavior of 3 d building frames ...Soil structure interaction effect on dynamic behavior of 3 d building frames ...
Soil structure interaction effect on dynamic behavior of 3 d building frames ...eSAT Journals
 

Tendances (20)

Settlement of shallow foundations
Settlement of shallow foundationsSettlement of shallow foundations
Settlement of shallow foundations
 
Chapter 12
Chapter 12Chapter 12
Chapter 12
 
Settlement of soil/foundation
Settlement of soil/foundationSettlement of soil/foundation
Settlement of soil/foundation
 
Ch9 settlement (1)
Ch9 settlement (1)Ch9 settlement (1)
Ch9 settlement (1)
 
Chapter 18
Chapter 18Chapter 18
Chapter 18
 
liquifaction
liquifactionliquifaction
liquifaction
 
Understanding Permeability of Hydraulic Fracture Networks A Sandbox Analog Mo...
Understanding Permeability of Hydraulic Fracture Networks A Sandbox Analog Mo...Understanding Permeability of Hydraulic Fracture Networks A Sandbox Analog Mo...
Understanding Permeability of Hydraulic Fracture Networks A Sandbox Analog Mo...
 
Effect of soil structure interaction on high rise r.c regular frame structur...
Effect of soil  structure interaction on high rise r.c regular frame structur...Effect of soil  structure interaction on high rise r.c regular frame structur...
Effect of soil structure interaction on high rise r.c regular frame structur...
 
Structure soil-structure interaction literature review
Structure soil-structure interaction literature reviewStructure soil-structure interaction literature review
Structure soil-structure interaction literature review
 
Goetech. engg. Ch# 03 settlement analysis signed
Goetech. engg. Ch# 03 settlement analysis signedGoetech. engg. Ch# 03 settlement analysis signed
Goetech. engg. Ch# 03 settlement analysis signed
 
6 effective stress capillarity_permeability
6 effective stress capillarity_permeability6 effective stress capillarity_permeability
6 effective stress capillarity_permeability
 
Seismic ssi effects and liquification
Seismic ssi effects and liquificationSeismic ssi effects and liquification
Seismic ssi effects and liquification
 
Seismic Behaviour of Multi-Storied Building by Using Tuned Mass Damper and Ba...
Seismic Behaviour of Multi-Storied Building by Using Tuned Mass Damper and Ba...Seismic Behaviour of Multi-Storied Building by Using Tuned Mass Damper and Ba...
Seismic Behaviour of Multi-Storied Building by Using Tuned Mass Damper and Ba...
 
Soil Structure Interaction
Soil Structure InteractionSoil Structure Interaction
Soil Structure Interaction
 
Deformation activity
Deformation activityDeformation activity
Deformation activity
 
Case study on effect of water table on bearing capacity
Case study on effect of water table on bearing capacityCase study on effect of water table on bearing capacity
Case study on effect of water table on bearing capacity
 
INFLUENCE OF SOIL-STRUCTURE INTERACTION ON RESPONSE OF A MULTI-STORIED BUILDI...
INFLUENCE OF SOIL-STRUCTURE INTERACTION ON RESPONSE OF A MULTI-STORIED BUILDI...INFLUENCE OF SOIL-STRUCTURE INTERACTION ON RESPONSE OF A MULTI-STORIED BUILDI...
INFLUENCE OF SOIL-STRUCTURE INTERACTION ON RESPONSE OF A MULTI-STORIED BUILDI...
 
Soil Amplification and Soil-Structure Interaction
Soil Amplification and Soil-Structure InteractionSoil Amplification and Soil-Structure Interaction
Soil Amplification and Soil-Structure Interaction
 
Soil structure interaction amec presentation-final
Soil structure interaction amec presentation-finalSoil structure interaction amec presentation-final
Soil structure interaction amec presentation-final
 
Soil structure interaction effect on dynamic behavior of 3 d building frames ...
Soil structure interaction effect on dynamic behavior of 3 d building frames ...Soil structure interaction effect on dynamic behavior of 3 d building frames ...
Soil structure interaction effect on dynamic behavior of 3 d building frames ...
 

Similaire à 63027 07

Comparision of Analysis of Overhead Intze Water Tank by Finite Element Method...
Comparision of Analysis of Overhead Intze Water Tank by Finite Element Method...Comparision of Analysis of Overhead Intze Water Tank by Finite Element Method...
Comparision of Analysis of Overhead Intze Water Tank by Finite Element Method...IRJET Journal
 
Comparison between static and dynamic analysis of elevated water tank 2
Comparison between static and dynamic analysis of elevated water tank 2Comparison between static and dynamic analysis of elevated water tank 2
Comparison between static and dynamic analysis of elevated water tank 2IAEME Publication
 
Module 2 jif 104 liquid
Module 2 jif 104 liquidModule 2 jif 104 liquid
Module 2 jif 104 liquidKurenai Ryu
 
Consolidation seminar......
Consolidation seminar......Consolidation seminar......
Consolidation seminar......Akshata shettar
 
SOIL STRENGTH AND SOIL FORCES
SOIL STRENGTH AND SOIL FORCESSOIL STRENGTH AND SOIL FORCES
SOIL STRENGTH AND SOIL FORCESAfendiAriff
 
Introduction to Reservoir Geomechanics
Introduction to Reservoir GeomechanicsIntroduction to Reservoir Geomechanics
Introduction to Reservoir GeomechanicsFarida Ismayilova
 
Surface Tension Experiment No. 4.pdf
Surface Tension Experiment No. 4.pdfSurface Tension Experiment No. 4.pdf
Surface Tension Experiment No. 4.pdfKaiwan B. Hamasalih
 
Soil Water Energy Concept
Soil Water Energy ConceptSoil Water Energy Concept
Soil Water Energy ConceptAhmad Hassan
 
Engineering materials related terms .
Engineering materials related terms .Engineering materials related terms .
Engineering materials related terms .Tapon Chakrabarti
 
23. The effect of plastic fines on the pore pressure generation characteristi...
23. The effect of plastic fines on the pore pressure generation characteristi...23. The effect of plastic fines on the pore pressure generation characteristi...
23. The effect of plastic fines on the pore pressure generation characteristi...PinakRay2
 
Theoretical models for surface forces and adhesion and their measurement usin...
Theoretical models for surface forces and adhesion and their measurement usin...Theoretical models for surface forces and adhesion and their measurement usin...
Theoretical models for surface forces and adhesion and their measurement usin...Grupo de Pesquisa em Nanoneurobiofisica
 
3 Surface Tension and Its MeasurementSina Ebnesajjad3.1 In.docx
3 Surface Tension and Its MeasurementSina Ebnesajjad3.1 In.docx3 Surface Tension and Its MeasurementSina Ebnesajjad3.1 In.docx
3 Surface Tension and Its MeasurementSina Ebnesajjad3.1 In.docxlorainedeserre
 

Similaire à 63027 07 (20)

Effective stress
Effective stressEffective stress
Effective stress
 
63027 11a
63027 11a63027 11a
63027 11a
 
Considering soil
Considering soilConsidering soil
Considering soil
 
Sullivan
SullivanSullivan
Sullivan
 
Comparision of Analysis of Overhead Intze Water Tank by Finite Element Method...
Comparision of Analysis of Overhead Intze Water Tank by Finite Element Method...Comparision of Analysis of Overhead Intze Water Tank by Finite Element Method...
Comparision of Analysis of Overhead Intze Water Tank by Finite Element Method...
 
Comparison between static and dynamic analysis of elevated water tank 2
Comparison between static and dynamic analysis of elevated water tank 2Comparison between static and dynamic analysis of elevated water tank 2
Comparison between static and dynamic analysis of elevated water tank 2
 
Module 2 jif 104 liquid
Module 2 jif 104 liquidModule 2 jif 104 liquid
Module 2 jif 104 liquid
 
Consolidation seminar......
Consolidation seminar......Consolidation seminar......
Consolidation seminar......
 
SOIL STRENGTH AND SOIL FORCES
SOIL STRENGTH AND SOIL FORCESSOIL STRENGTH AND SOIL FORCES
SOIL STRENGTH AND SOIL FORCES
 
Introduction to Reservoir Geomechanics
Introduction to Reservoir GeomechanicsIntroduction to Reservoir Geomechanics
Introduction to Reservoir Geomechanics
 
Surface Tension Experiment No. 4.pdf
Surface Tension Experiment No. 4.pdfSurface Tension Experiment No. 4.pdf
Surface Tension Experiment No. 4.pdf
 
Surface tension
Surface tensionSurface tension
Surface tension
 
Soil Water Energy Concept
Soil Water Energy ConceptSoil Water Energy Concept
Soil Water Energy Concept
 
Stress in rock
Stress in rockStress in rock
Stress in rock
 
Engineering materials related terms .
Engineering materials related terms .Engineering materials related terms .
Engineering materials related terms .
 
23. The effect of plastic fines on the pore pressure generation characteristi...
23. The effect of plastic fines on the pore pressure generation characteristi...23. The effect of plastic fines on the pore pressure generation characteristi...
23. The effect of plastic fines on the pore pressure generation characteristi...
 
N013148591
N013148591N013148591
N013148591
 
Theoretical models for surface forces and adhesion and their measurement usin...
Theoretical models for surface forces and adhesion and their measurement usin...Theoretical models for surface forces and adhesion and their measurement usin...
Theoretical models for surface forces and adhesion and their measurement usin...
 
3 Surface Tension and Its MeasurementSina Ebnesajjad3.1 In.docx
3 Surface Tension and Its MeasurementSina Ebnesajjad3.1 In.docx3 Surface Tension and Its MeasurementSina Ebnesajjad3.1 In.docx
3 Surface Tension and Its MeasurementSina Ebnesajjad3.1 In.docx
 
Shear Strength Behavior for Unsoaked and Soaked Multistage Triaxial Drained T...
Shear Strength Behavior for Unsoaked and Soaked Multistage Triaxial Drained T...Shear Strength Behavior for Unsoaked and Soaked Multistage Triaxial Drained T...
Shear Strength Behavior for Unsoaked and Soaked Multistage Triaxial Drained T...
 

Plus de Vcoi Vit

2eae0e86 e7c6-431d-aa08-b8f89ba71921 giaotrinhkinhtechinhtri31
2eae0e86 e7c6-431d-aa08-b8f89ba71921 giaotrinhkinhtechinhtri312eae0e86 e7c6-431d-aa08-b8f89ba71921 giaotrinhkinhtechinhtri31
2eae0e86 e7c6-431d-aa08-b8f89ba71921 giaotrinhkinhtechinhtri31Vcoi Vit
 
2eae0e86 e7c6-431d-aa08-b8f89ba71921 giaotrinhkinhtechinhtri14
2eae0e86 e7c6-431d-aa08-b8f89ba71921 giaotrinhkinhtechinhtri142eae0e86 e7c6-431d-aa08-b8f89ba71921 giaotrinhkinhtechinhtri14
2eae0e86 e7c6-431d-aa08-b8f89ba71921 giaotrinhkinhtechinhtri14Vcoi Vit
 
2eae0e86 e7c6-431d-aa08-b8f89ba71921 giaotrinhkinhtechinhtri13
2eae0e86 e7c6-431d-aa08-b8f89ba71921 giaotrinhkinhtechinhtri132eae0e86 e7c6-431d-aa08-b8f89ba71921 giaotrinhkinhtechinhtri13
2eae0e86 e7c6-431d-aa08-b8f89ba71921 giaotrinhkinhtechinhtri13Vcoi Vit
 
2eae0e86 e7c6-431d-aa08-b8f89ba71921 giaotrinhkinhtechinhtri05
2eae0e86 e7c6-431d-aa08-b8f89ba71921 giaotrinhkinhtechinhtri052eae0e86 e7c6-431d-aa08-b8f89ba71921 giaotrinhkinhtechinhtri05
2eae0e86 e7c6-431d-aa08-b8f89ba71921 giaotrinhkinhtechinhtri05Vcoi Vit
 
Giao trinh nuoi_trong_thuy_san46
Giao trinh nuoi_trong_thuy_san46Giao trinh nuoi_trong_thuy_san46
Giao trinh nuoi_trong_thuy_san46Vcoi Vit
 
Giao trinh nuoi_trong_thuy_san28
Giao trinh nuoi_trong_thuy_san28Giao trinh nuoi_trong_thuy_san28
Giao trinh nuoi_trong_thuy_san28Vcoi Vit
 
Giao trinh nuoi_trong_thuy_san23
Giao trinh nuoi_trong_thuy_san23Giao trinh nuoi_trong_thuy_san23
Giao trinh nuoi_trong_thuy_san23Vcoi Vit
 
Giao trinh nuoi_trong_thuy_san22
Giao trinh nuoi_trong_thuy_san22Giao trinh nuoi_trong_thuy_san22
Giao trinh nuoi_trong_thuy_san22Vcoi Vit
 
Giao trinh nuoi_trong_thuy_san21
Giao trinh nuoi_trong_thuy_san21Giao trinh nuoi_trong_thuy_san21
Giao trinh nuoi_trong_thuy_san21Vcoi Vit
 
Giao trinh nuoi_trong_thuy_san16
Giao trinh nuoi_trong_thuy_san16Giao trinh nuoi_trong_thuy_san16
Giao trinh nuoi_trong_thuy_san16Vcoi Vit
 
Giao trinh nuoi_trong_thuy_san05
Giao trinh nuoi_trong_thuy_san05Giao trinh nuoi_trong_thuy_san05
Giao trinh nuoi_trong_thuy_san05Vcoi Vit
 
Giao trinh nuoi_trong_thuy_san02
Giao trinh nuoi_trong_thuy_san02Giao trinh nuoi_trong_thuy_san02
Giao trinh nuoi_trong_thuy_san02Vcoi Vit
 
Giao trinh tong hop sv50
Giao trinh tong hop sv50Giao trinh tong hop sv50
Giao trinh tong hop sv50Vcoi Vit
 
Giao trinh tong hop sv49
Giao trinh tong hop sv49Giao trinh tong hop sv49
Giao trinh tong hop sv49Vcoi Vit
 
Giao trinh tong hop sv48
Giao trinh tong hop sv48Giao trinh tong hop sv48
Giao trinh tong hop sv48Vcoi Vit
 
Giao trinh tong hop sv47
Giao trinh tong hop sv47Giao trinh tong hop sv47
Giao trinh tong hop sv47Vcoi Vit
 
Giao trinh tong hop sv46
Giao trinh tong hop sv46Giao trinh tong hop sv46
Giao trinh tong hop sv46Vcoi Vit
 
Giao trinh tong hop sv45
Giao trinh tong hop sv45Giao trinh tong hop sv45
Giao trinh tong hop sv45Vcoi Vit
 
Giao trinh tong hop sv43
Giao trinh tong hop sv43Giao trinh tong hop sv43
Giao trinh tong hop sv43Vcoi Vit
 
Giao trinh tong hop sv41
Giao trinh tong hop sv41Giao trinh tong hop sv41
Giao trinh tong hop sv41Vcoi Vit
 

Plus de Vcoi Vit (20)

2eae0e86 e7c6-431d-aa08-b8f89ba71921 giaotrinhkinhtechinhtri31
2eae0e86 e7c6-431d-aa08-b8f89ba71921 giaotrinhkinhtechinhtri312eae0e86 e7c6-431d-aa08-b8f89ba71921 giaotrinhkinhtechinhtri31
2eae0e86 e7c6-431d-aa08-b8f89ba71921 giaotrinhkinhtechinhtri31
 
2eae0e86 e7c6-431d-aa08-b8f89ba71921 giaotrinhkinhtechinhtri14
2eae0e86 e7c6-431d-aa08-b8f89ba71921 giaotrinhkinhtechinhtri142eae0e86 e7c6-431d-aa08-b8f89ba71921 giaotrinhkinhtechinhtri14
2eae0e86 e7c6-431d-aa08-b8f89ba71921 giaotrinhkinhtechinhtri14
 
2eae0e86 e7c6-431d-aa08-b8f89ba71921 giaotrinhkinhtechinhtri13
2eae0e86 e7c6-431d-aa08-b8f89ba71921 giaotrinhkinhtechinhtri132eae0e86 e7c6-431d-aa08-b8f89ba71921 giaotrinhkinhtechinhtri13
2eae0e86 e7c6-431d-aa08-b8f89ba71921 giaotrinhkinhtechinhtri13
 
2eae0e86 e7c6-431d-aa08-b8f89ba71921 giaotrinhkinhtechinhtri05
2eae0e86 e7c6-431d-aa08-b8f89ba71921 giaotrinhkinhtechinhtri052eae0e86 e7c6-431d-aa08-b8f89ba71921 giaotrinhkinhtechinhtri05
2eae0e86 e7c6-431d-aa08-b8f89ba71921 giaotrinhkinhtechinhtri05
 
Giao trinh nuoi_trong_thuy_san46
Giao trinh nuoi_trong_thuy_san46Giao trinh nuoi_trong_thuy_san46
Giao trinh nuoi_trong_thuy_san46
 
Giao trinh nuoi_trong_thuy_san28
Giao trinh nuoi_trong_thuy_san28Giao trinh nuoi_trong_thuy_san28
Giao trinh nuoi_trong_thuy_san28
 
Giao trinh nuoi_trong_thuy_san23
Giao trinh nuoi_trong_thuy_san23Giao trinh nuoi_trong_thuy_san23
Giao trinh nuoi_trong_thuy_san23
 
Giao trinh nuoi_trong_thuy_san22
Giao trinh nuoi_trong_thuy_san22Giao trinh nuoi_trong_thuy_san22
Giao trinh nuoi_trong_thuy_san22
 
Giao trinh nuoi_trong_thuy_san21
Giao trinh nuoi_trong_thuy_san21Giao trinh nuoi_trong_thuy_san21
Giao trinh nuoi_trong_thuy_san21
 
Giao trinh nuoi_trong_thuy_san16
Giao trinh nuoi_trong_thuy_san16Giao trinh nuoi_trong_thuy_san16
Giao trinh nuoi_trong_thuy_san16
 
Giao trinh nuoi_trong_thuy_san05
Giao trinh nuoi_trong_thuy_san05Giao trinh nuoi_trong_thuy_san05
Giao trinh nuoi_trong_thuy_san05
 
Giao trinh nuoi_trong_thuy_san02
Giao trinh nuoi_trong_thuy_san02Giao trinh nuoi_trong_thuy_san02
Giao trinh nuoi_trong_thuy_san02
 
Giao trinh tong hop sv50
Giao trinh tong hop sv50Giao trinh tong hop sv50
Giao trinh tong hop sv50
 
Giao trinh tong hop sv49
Giao trinh tong hop sv49Giao trinh tong hop sv49
Giao trinh tong hop sv49
 
Giao trinh tong hop sv48
Giao trinh tong hop sv48Giao trinh tong hop sv48
Giao trinh tong hop sv48
 
Giao trinh tong hop sv47
Giao trinh tong hop sv47Giao trinh tong hop sv47
Giao trinh tong hop sv47
 
Giao trinh tong hop sv46
Giao trinh tong hop sv46Giao trinh tong hop sv46
Giao trinh tong hop sv46
 
Giao trinh tong hop sv45
Giao trinh tong hop sv45Giao trinh tong hop sv45
Giao trinh tong hop sv45
 
Giao trinh tong hop sv43
Giao trinh tong hop sv43Giao trinh tong hop sv43
Giao trinh tong hop sv43
 
Giao trinh tong hop sv41
Giao trinh tong hop sv41Giao trinh tong hop sv41
Giao trinh tong hop sv41
 

63027 07

  • 1. 173 CHAPTER 7 Copyrighted Material Effective, Intergranular, and Total Stress 7.1 INTRODUCTION The compressibility, deformation, and strength prop-erties of a soil mass depend on the effort required to distort or displace particles or groups of particles rel-ative 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 de-pend 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 ex-ample, by its water content, structure, density, or void ratio, reflects the influences of stresses applied in the past, and this further distinguishes soils from most other engineering materials, which, for practical pur-poses, 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. For-tunately, however, the stresses, the state, and the prop-erties are not independent, and the relationships between stress and volume change, stress and stiffness, and stress and strength can be expressed in terms of definable soil parameters such as compressibility and friction angle. In soils with properties that are influ-enced significantly by chemical and physicochemical forces of interaction, other parameters such as cohe-sion may be needed. Most problems involving volume change, deforma-tion, and strength require separate consideration of the stress that is carried by the grain assemblage and that carried by the fluid phases. This distinction is essential because an assemblage of grains in contact can resist both normal and shear stress, but the fluid and gas phases (usually water and air) can carry normal stress but not shear stress. Furthermore, whenever the total head in the fluid phases within the soil mass differs from that outside the soil mass, there will be fluid flow 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 fluid. 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 Engi-neering. 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 filled 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. Copyright © 2005 John Wiley Sons Retrieved from: www.knovel.com
  • 2. 174 7 EFFECTIVE, INTERGRANULAR, AND TOTAL STRESS 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 influence 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 meas-urable effects of a change of stress, such as compression, distortion and a change of shearing resistance are exclu-sively Copyrighted Material due to changes in the effective stresses 1, 2 and . Hence every investigation of the stability of a saturated 3 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 com-puted using the external forces and the body force due to weight of the soil–water mixture. A pore water pres-sure, denoted herein by u0, can be measured at a point remote from the interparticle zone. The actual pore wa-ter 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 is indeed equal to u cannot be ascertained i without more detailed examination of all the interpar- 1The 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 first three are the normal stresses and the latter three are the shear stresses. In this case, the effective stresses are defined as 11 11 u, 22 22 u, 33 33 u, , , and . 12 12 23 23 31 31 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 load-ing) (Fig. 7.1a)]. 2. Particle Level Forces These include particle weight force, buoyancy force when a particle is submerged under fluid, and hydrodynamic forces or seepage forces due to pore fluid 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). When external forces are applied, both normal and tangential forces develop at particle contacts. All par-ticles 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 in-terparticle skeletal forces in soils govern the macro-scopic stress–strain 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 photo-elastic model shown in Fig. 5.15. When the pores are filled with fluids, the weight of the fluids adds to the body force of the soil–fluids mixture. However, hydro-static pressure results from the fluid weight, and the uplift force due to buoyancy reduces the effective weight of a fluid-filled soil. This leads to smaller skel-etal forces for submerged soil compared to dry soil. Seepage forces that result from additional fluid pres-sures applied externally produce hydrodynamic forces on particles and alter the skeletal forces. 7.4 INTERPARTICLE FORCES Long-range particle interactions associated with elec-trical double layers and van der Waals forces are dis- Copyright © 2005 John Wiley Sons Retrieved from: www.knovel.com
  • 3. INTERPARTICLE FORCES 175 External Load Interparticle Forces Material Copyrighted Interparticle Forces (a) Body Force Buoyancy Force if Saturated Viscous Drag by Seepage Flow Seepage (b) Capillary Force or Cementation-reactive Force Electrical Forces (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 flocculation–deflocculation 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 impor-tant as well since they may influence the intergranular stresses and control the strength at interparticle con-tacts, 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 pres-sure 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 sufficiently great to prevent the interpenetra-tion of matter. At separation distances beyond the region of direct physical interference between adsorbed ions and hy-dration water molecules, double-layer interactions pro-vide 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 fluid. 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 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 sufficient to squeeze out all the water between parallel particle sur-faces in naturally occurring clays. Heat and/or high vacuum are needed to remove all the water from a fine-grained 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 pro-portion ( 1%) of the total soil cross-sectional area in most cases. The exact nature of an interparticle con-tact 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 dis-tance. 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 differ-ent potentials has been suggested as a cause. When the Copyright © 2005 John Wiley Sons Retrieved from: www.knovel.com
  • 4. 176 7 EFFECTIVE, INTERGRANULAR, AND TOTAL STRESS gap between parallel particle surfaces separated by dis-tance d at potentials V1 and V2 is conductive, there is an attractive force per unit area, or tensile strength, given by (Ingles, 1962) 4.4 106 (V V )2 F 1 2 N/m2 (7.1) d2 Copyrighted Material where F is the tensile strength, d is in micrometers, and V1 and V2 are in millivolts. This force is indepen-dent of particle size and becomes significant (greater than 7 kN/m2 or 1 psi) for separation distances less than 2.5 nm. Electromagnetic Attractions Electromagnetic at-tractions caused by frequency-dependent dipole inter-actions (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 specifically for fine-grained soils with various geomet-ric parameters such as particle length, thickness, ori-entation, and spacing. Primary Valence Bonding Chemical interactions between particles and between the particles and adja-cent liquid phase can only develop at short range. Co-valent 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 hydro-gen bonds, can develop at interparticle contacts with-out the presence of cementing agents is largely a matter of speculation. Very high contact stresses be-tween particles could squeeze out adsorbed water and cations and cause mineral surfaces to come close to-gether, perhaps providing opportunity for cold weld-ing. 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 (Chap-ter 11). Thus, interatomic bonding between particles seems possible. On the other hand, the absence of co-hesion in overconsolidated silts and sands argues against such pressure-induced bonding. Cementation Cementation may develop naturally from precipitation of calcite, silica, alumina, iron ox-ides, and possibly other inorganic or organic com-pounds. 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 ten-sile 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 ce-ment– particle interface. The following equation can be derived (Ingles, 1962) for the tensile strength T per unit area of soil cross section: 1 n Pk(7.2) T 1 e n Ai 1 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 Pk (7.3) T d2(1 e) For a random and isotropic assembly of rods of length l and diameter d Pk (7.4) T d(l d/2)(1 e) Bond strength P is evaluated in the following way (Fig. 7.2) for two cemented spheres of radius R. It may be shown that (R cos ) cosh R sin (7.5) so for known , can be computed. Then, for cement failure, P 2 (7.6) c where c is the tensile strength of the cement; for sphere failure, P ()2 (7.7) s where R sin , and s is the tensile strength of the sphere, and for failure at the interface Copyright © 2005 John Wiley Sons Retrieved from: www.knovel.com
  • 5. INTERPARTICLE FORCES 177 Copyrighted Material Figure 7.2 Contact zone failures for cemented spheres. sin P 2R2(1 cos ) (7.8) 1 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. Artificially ce-mented soils are often loaded after cementation has developed, whereas cementation develops during or af-ter overburden loading in natural soils. In the former case, the particles and cementation bonding are loaded together and contact forces can become negative de-pending on the tensile resistance of cementation bond-ing. The distribution and magnitude of skeletal forces are therefore influenced by both geometric arrange-ment 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 de-veloped before cementation coats the already loaded particles. In this case, it is possible that cementation creates extra forces at particle contacts. In some ce-mented natural materials, if the soil is unloaded from high overburden stress, elastic rebound may disrupt ce-mented 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 fluid 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 tem-porary 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 equa-tion, assuming a capillary tube as shown in Fig. 7.3a: 2 cos
  • 6. aw ˆP (7.9) c rp where is the capillary pressure at air entry, ˆc P is aw the air–water interfacial tension,
  • 7. is contact angle de-fined in Fig. 7.3, and rp is the tube radius. For pure water and air, aw depends on temperature, for exam-ple, 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 then air in- ˆP , c vades the pore.2 Since soil has pores with various sizes, the water in the largest pores is displaced first followed by smaller pores. This leads to a macroscopic model of the soil–water characteristic curve (or the capillary pressure–saturation relationship), as discussed in Sec-tion 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 infiltration when air in the pores cannot escape or the air boundary is completely blocked. Copyright © 2005 John Wiley Sons Retrieved from: www.knovel.com
  • 8. 178 7 EFFECTIVE, INTERGRANULAR, AND TOTAL STRESS Capillary Tube Representing a Pore 2rp dc Material Copyrighted θ ua uw ^ 2σaw cosθ Pc = ρw gdc = 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. 2r2 cos
  • 9. F r2 Pˆ aw (7.10) c c rp where rp is the size of the pore into which the air has entered. Since the fluid acts like a membrane with neg-ative 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 be-comes disconnected and remains in the form of me-nisci or liquid bridges at the interparticle contacts [termed the ‘‘pendular’’ regime by Bear (1972)]. The curved air–water interface produces a pore water ten-sion, which, in turn, generates interparticle compres-sive 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 sep-aration between the two particles, the radius of the liq-uid bridge, interfacial tension, and contact angle (Lian et al., 1993). Once the water phase becomes discontin-uous, evaporation and condensation are the primary mechanisms of water transfer. Hence, the humidity of the gas phase and the temperature affect the water va-por pressure at the surface of water menisci, which in turn influences the air pressure ua. 7.5 INTERGRANULAR PRESSURE Several different interparticle forces were described in the previous section. Quantitative expression of the in-teractions of all these forces in a soil is beyond the present state of knowledge. Nonetheless, their exis-tence bears directly on the magnitude of intergranular pressure and the relationship between intergranular pressure and effective stress as defined by u. A simplified 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 x–x is con-sidered. 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. Define area a as 3Note that only vertical forces at the contact are considered in this simplified analysis. It is evident, however, that applied boundary nor-mal and shear stresses each induce both normal and shear forces at interparticle contacts. These forces contribute both to the develop-ment of soil strength and resistance to compression and to the slip-ping 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. Copyright © 2005 John Wiley Sons Retrieved from: www.knovel.com
  • 10. INTERGRANULAR PRESSURE 179 Material (a) Copyrighted Soil Particles Continuous Water Film Negative pore pressure acts all around the particles Suction forces act only at particle contacts and the magnitude of the forces depends on the size of liquid bridges. (b) Liquid Bridges Soil Particles Interparticle Forces Pores of Radius rp Filled with Air Air Figure 7.4 Microscopic water–soil interaction in unsaturated soils: (a) funicular regime and (b) pendular regime. Figure 7.6 Forces acting on interparticle contact A. Figure 7.5 Surfaces through a soil mass. the average total cross-sectional area along a horizontal plane served by the contact. It equals the total hori-zontal area divided by the number of interparticle con-tacts along the wavy surface. The forces acting on area a in Fig. 7.6 are: 1. a, the force transmitted by the applied stress , which includes externally applied forces and body weight from the soil above. Copyright © 2005 John Wiley Sons Retrieved from: www.knovel.com
  • 11. 180 7 EFFECTIVE, INTERGRANULAR, AND TOTAL STRESS 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 long-range attractive stress A, that is, van der Waals and electrostatic attractions. 4. Aac, the force developed by the short-range at-tractive stress A, resulting from primary valence Material Copyrighted (chemical) bonding and cementation. 5. Cac, the intergranular contact reaction that is gen-erated by hydration and Born repulsion. Vertical equilibrium of forces requires that a Aa Aa ua Ca (7.11) c c Division of all terms by a converts the forces to stresses per unit area of cross section, ac (C A) u A (7.12) a 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 me-chanics usage. Designation of this term by gives i A u (7.13) i 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 con-dition that a a a i.e., a → 0 w a c The resulting equation is aw A u (u u ) (7.14) i a a w a In the absence of significant long-range attractions, this equation is similar to that proposed by Bishop (1960) for partially saturated soils u (u u ) (7.15) i a a w 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 sat-uration. Further discussion of the effective stress con-cept for unsaturated soils is given in Section 7.12. Limiting the discussion to saturated soils, two ques-tions arise: 1. How does the intergranular pressure relate to i the effective stress as defined for most analyses, that is, u? 2. How does the intergranular pressure relate to i the measured quantity, u0m , that is taken as the effective stress, recalling (Section 7.2) that pore pressure can only be measured at points out-side the true interparticle zone? Answers to these questions require a more detailed consideration of the meaning of fluid pressures in soils. 7.6 WATER PRESSURES AND POTENTIALS Pressures in the pore fluid of a soil can be expressed in several ways, and the total pressure may involve several contributions. In hydraulic engineering, prob-lems are analyzed using Bernoulli’s equation for the total heads and head losses associated with flow be-tween two points, that is, p v2 p v2 Z 1 1 Z 2 2 h (7.16) 1 2 1–2 2g 2g w w 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 flow velocities at points 1 and 2, w is the unit weight of water, g is the acceleration due to gravity, and h1–2 is the loss in head between points 1 and 2. The total head H (dimension L) is p v2 H Z (7.17) 2g w Flow results only from differences in total head; conversely, if the total heads at two points are the same, there can be no flow, even if Z1 Z2 and p1 p2. If there is no flow, there is no head loss and h1–2 0. The flow velocity through soils is low, and as a re-sult v2/2g → 0, and in most cases it may be neglected. Therefore, the relationship Copyright © 2005 John Wiley Sons Retrieved from: www.knovel.com
  • 12. WATER PRESSURE EQUILIBRIUM IN SOIL 181 p p Z 1 Z 2 h (7.18) 1 2 1–2 w w is the basis for evaluation of pore pressures and anal-ysis of seepage through soils and other porous media. Although the absence of velocity terms is a factor that seems to simplify the analysis of flows and pres-sures in soils, there are other considerations that tend Copyrighted Material to complicate the problem. These include: 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 wa-ter. 3. Compositional differences in the water from point-to-point and adsorptive force fields from particle surfaces. 4. Differences in interparticle forces and the energy state of the pore fluid from point to point owing to thermal, electrical, and chemical gradients. Such gradients can cause fluid flows, deforma-tions, and volume changes, as considered in more detail in Chapter 9. Some formalism in definition and terminology is necessary to avoid confusion. The status of water in a soil can be expressed in terms of the free energy rel-ative to free, pure water (Aitchison, et al., 1965). The free energy can be (and is) expressed in different ways, including 1. Potential (dimensions—L2T2: J/kg) 2. Head (dimensions—L: m, cm, ft) 3. Pressure (dimensions—ML1 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 definition of total po-tential is the work per unit quantity to transport re-versibly and isothermally an infinitesimal amount of pure water from a pool at a specified elevation at at-mospheric pressure to the point in soil water under consideration. The selection of the components of the total poten-tial (total head H, total pressure P) is somewhat arbitrary (Bolt and Miller, 1958); however, the follow-ing have gained acceptance for geotechnical work (Aitchison, et al., 1965): 1. Gravitational potential g (head Z, pressure pz) corresponds to elevation head in normal hydrau-lic usage. 2. Matrix or capillary potential m (head hm, pres-sure p) is the work per unit quantity of water to transport reversibly and isothermally an infinites-imal 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 consider-ation 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 pres-sures generated by capillarity menisci, and from water adsorption forces exerted by particle sur-faces. A piezometer measures the matrix poten-tial if it contains fluid of the same composition as the soil water. 3. Osmotic (or solute) potential s (head hs, pres-sure ps) is the work per unit quantity of water to transport reversibly and isothermally an infinites-imal quantity of water from a pool of pure water at a specified elevation and atmospheric pressure to a pool containing a solution identical in com-position 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 wa-ter, and it depends on the composition and ability of the soil particles to restrain the movement of adsorbed cations. The osmotic potential is nega-tive, that is, water tends to flow in the direction of increasing concentration. The total potential, head, and pressure then become (7.19) g m s H Z h h (7.20) m s P p p p (7.21) z s At equilibrium and no flow there can be no varia-tions 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 Copyright © 2005 John Wiley Sons Retrieved from: www.knovel.com
  • 13. 182 7 EFFECTIVE, INTERGRANULAR, AND TOTAL STRESS Material Copyrighted 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, H 0 h 0 h 0 m0 m0 It follows that P h u (7.22) 0 m0 w 0 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, H 0 h h 1 m1 s1 If no water is flowing, H1 H0, and h h h m1 s1 m0 Also, because p1 p0 u0 u h h (7.23) 0 m1w s1 w At point 2, which is between the same two clay par-ticles 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, u h h h h h 0 m2 s2 m1 s1 m0 w A similar analysis could be applied to any point in the system. If point 3 were midway between two clay par-ticles spaced the same distance apart as the particles on either side of point 1, then hs3 hs1, but Z3 0. Thus, u0 Z h h Z h h (7.24) 3 m3 s3 3 m3 s1 w A partially saturated system can also be analyzed, but the influences of curved air–water 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: 1. As the osmotic and gravitational components vary from point to point in a soil at equilibrium, Copyright © 2005 John Wiley Sons Retrieved from: www.knovel.com
  • 14. MEASUREMENT OF PORE PRESSURES IN SOILS 183 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 vi-sualize. Nonetheless, this underlies the whole concept of water flow by chemical osmosis. Copyrighted Material 2. The total potential, head, and pressure are meas-urable, and separation into components is possi-ble experimentally, although it is difficult. 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 par-ticles, u0 is seen to include contributions from osmotic pressures as well as matrix pressures. Since osmotic pressures are the cause of long-range repulsions due to double-layer interactions, measured pore water pressures may include con-tributions 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 lab-oratory use, whereas others are intended for use in the field. 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 pi-ezometer communicates with the soil through a porous stone or filter. Pressures are determined from the water level in a standpipe, by a manom-eter, by a pressure gauge, or by an electronic pressure transducer. A piezometer used to mea-sure pressures less than atmospheric is usually termed a tensiometer. 2. Gypsum Block, Porous Ceramic, and Filter Paper The electrical properties across a spe-cially 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 deter-mines the amount of moisture in the block (Whalley et al., 2001). The same principle can be applied by placing a dry filter paper on a soil specimen and allowing the soil moisture to ab-sorb into the paper. When the suction in the filter paper is equal to the suction in the soil, the two reach equilibrium, and the suction can be deter-mined by the water content of the filter paper. These techniques are used for measurement of pore pressures less than atmospheric. 3. 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 con-tent and pressure is used to establish the relation-ship between soil suction and water content. 4. Consolidation Tests The consolidation pressure on a sample at equilibrium is the soil water suc-tion. If the consolidation pressure were instanta-neously removed, then a negative water pressure or suction of the same magnitude would be needed to prevent water movement into the soil. 5. Vapor Pressure Methods The relationship be-tween relative humidity and water content is used to establish the relationship between suction and water content. 6. Osmotic Pressure Methods Soil samples are equilibrated with solutions of known osmotic pressure to give a relationship between water content and water suction. 7. Dielectric Sensors Such as Capacitance Probes and Time Domain Reflectometry 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 soil’s capacitance, which is related to dielec-tric constants of soil particle, water, and air. The capacitance is largely influenced by water con-tent, as the dielectric constant of water is large compared to the dielectric constants of soil particle and air. Time domain reflectrometry measures the travel time of a high-frequency, electromagnetic pulse. The presence of water in the soil slows down the speed of the electromag-netic wave by the change in the dielectric prop-erties. Volumetric water content can therefore be indirectly measured from the travel time mea-surement. 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 Copyright © 2005 John Wiley Sons Retrieved from: www.knovel.com
  • 15. 184 7 EFFECTIVE, INTERGRANULAR, AND TOTAL STRESS descriptions and comparisons of these and other meth-ods are given by Croney et al. (1952), Aitchison et al. (1965), Richards and Peter (1987), and Ridley et al. (2003). 7.9 EFFECTIVE AND INTERGRANULAR PRESSURE In Section 7.5, it was shown that the intergranular pres-sure Copyrighted Material is given by A u (7.25) i where u is the hydrostatic pressure between particles (or hm w in the terminology of Section 7.7). General-ized forms of Eq. (7.24) are u Z h h (7.26) 0 w mw sw and u h u Z h (7.27) m w 0 w sw Thus, Eq. (7.25) becomes, for the case of no elevation difference between a piezometer and the point in ques-tion (i.e., Z 0), A u h (7.28) i 0 s w 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 reflects 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 A u R (7.29) i 0 From Eq. (7.25), it was seen that the intergranular pressure was dependent on long-range interparticle at-tractions A as well as on the applied stress and the pore water pressure between particles u. Equation (7.29) indicates that if intergranular pressure is to i 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 par-ticles u u0 R includes the effects of long-range repulsions as required by the condition of constant to-tal potential for equilibrium. In the general case, therefore, the true intergranular pressure A u0 R and the conventionally i defined effective stress u0 differ by the net interparticle stress due to physicochemical contribu-tions, A R (7.30) i 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 signifi-cantly different magnitude, would the intergranular and effective stress be significantly different. Such a con-dition appears not to be common, although it might be of importance in a well-dispersed sodium montmoril-lonite, 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 equi-librium, with contributing forces parallel to each other, that is, the intergranular stress is the sum of the i skeletal forces (defined as u0) and the elec-trochemical stress (A R), as illustrated in Fig. 7.8a. This implies that the deformation induced by the elec-trochemical stress (A R) is equal to the deformation induced by the skeletal forces at contacts [i.e., a ‘‘par-allel’’ model as described by Hueckel (1992)]. The change in pore fluid chemistry at constant confinement () leads to changes in intergranular stresses (), re- i sulting 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): R A u (7.31) i 0 This is called the ‘‘series’’ model (Hueckel, 1992), and the model can be applicable for very fine 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 or ef- i fective stress changes the interparticle spacing, which may contribute to changes in strength properties upon shearing. 4A detailed analysis of effective stress in clays is presented by Chat-topadhyay (1972), which leads to similar conclusions, including Eq. (7.29). was termed the true effective stress and it governed the i volume change behavior of Na–montmorillonite. Copyright © 2005 John Wiley Sons Retrieved from: www.knovel.com
  • 16. ASSESSMENT OF TERZAGHI’S EQUATION 185 Electrochemical Force Electrochemical Force Skeletal Force σ = σ _ u0 Material Copyrighted Skeletal Force Electrochemical Force A _ R σi σi Deformation at the Contact σi = σ _ u0 + A _ R (a) Skeletal Force σi Skeletal Force σ = σ _ u0 Total Deformation at the Contact σi σi = σ _ u0 = A _ R (b) Particle Deformation by Skeletal Force Electrochemical Force A _ R Electrochemical Force Skeletal Force Electrochemical Force Skeletal Force 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 ex-ternal 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]. 7.10 ASSESSMENT OF TERZAGHI’S EQUATION The preceding equations and discussion do not confirm that Terzaghi’s 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 ap-plication in practice. Skempton (1960b) showed that the Terzaghi equation does not give the true effective stress but gives an excellent approximation for the case of saturated soils. Skempton proposed three possible relationships for effective stress in saturated soils: 1. The true intergranular pressure for the case when A R 0 (1 a )u (7.32) c 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 k tan (7.33) i where is an intrinsic friction angle and k is a true cohesion. The following relationships were derived: For shear strength, ac tan 1 u (7.34) tan where is the effective stress angle of shearing resistance. For volume change, Copyright © 2005 John Wiley Sons Retrieved from: www.knovel.com
  • 17. 186 7 EFFECTIVE, INTERGRANULAR, AND TOTAL STRESS Table 7.1 Compressibility Values for Soil, Rock, and Concrete Material Compressibilitya per kN/m2 106 C Cs Cs /C Quartzitic sandstone 0.059 0.027 0.46 Quincy granite (30 m deep) 0.076 0.019 0.25 Vermont marble 0.18 0.014 0.08 Concrete (approx.) 0.20 0.025 0.12 Dense sand 18 0.028 0.0015 Loose sand 92 0.028 0.0003 London clay (over cons.) 75 0.020 0.00025 Gosport clay (normally cons.) 600 0.020 0.00003 After Skempton (1960b). aCompressibilities at p 98 kN/m2; water Cw 0.49 106 per kN/m2. 1 Csu (7.35) C 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) Copyrighted Material 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 V C (7.37) V and also satisfied the Coulomb equation for drained shear strength d : c tan (7.38) d 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 Skempton’s 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 negli-gible, 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 gen-eral form (Lade and de Boer, 1997): u (7.39) where is the fraction of the pore pressure that gives the effective stress.5 Different expressions for pro-posed by several researchers are listed in Table 7.2. 5A more general expression has been proposed as ij ijij u, 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; Did-wania, 2002). 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 incre-ment u and that due to the change in confining pres-sure ( u) (or u). The effective stress increment is defined as the stress that produces the same volume change, CV V V CV ( u) 0 sks sku 0 C V u (7.40) u 0 where Vsks is the volume change of soil skeleton due to change in confining 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 confining pressure change, and Cu is the compressibility of the soil skeleton by pore pres-sure change. Rearranging Eq. (7.40) leads to Cu 1 u (7.41) C Lade and de Boer (1997) used this equation to de-rive an effective stress equation for granular materials under drained conditions. Consider a condition in which the total confining pressure is constant [ ( Copyright © 2005 John Wiley Sons Retrieved from: www.knovel.com
  • 18. ASSESSMENT OF TERZAGHI’S EQUATION 187 Table 7.2 Expressions for to Define Effective Stress Pore Pressure Fraction Note Reference 1 Terzaghi (1925b) n n porosity Biot (1955) 1 ac ac grain contact area per unit area of plane Skempton and Bishop (1954) 1 ac tan tan Equation (7.34) Skempton (1960b) 1 CEquation (7.35); for isotropic elastic s C deformation of a porous material; for solid rock with small interconnected pores and low porosity (Lade and de Boer, Material 1997) Copyrighted Biot and Willis (1957), Skempton (1960b), Nur and Byerlee (1971), Lade and de Boer (1997) 1 (1 n) Cs C Equation (7.43) Suklje (1969); Lade and de Boer (1997) After Lade and de Boer (1997). Figure 7.9 Variation of with stress for quartz sand and gypsum sand (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 vol-umetric compression of the solid grains ( Vgu). Hence, V C V u C (1 n)V u V or sku u 0 s 0 gu C C (1 n) u s (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) Cs u or C Cs 1 (1 n) (7.43) C 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 essen-tially one. Thus, Terzaghi’s effective stress equation, while not rigorously correct, is again shown to be an excellent approximation in almost all cases for satu-rated soils (i.e., solid grains and pore fluid are consid-ered to be incompressible compared to soil skeleton compressibility). 6An example of this condition is a soil under a seabed, in which the sea depth varies. This condition is often called the ‘‘unjacked con-dition.’’ 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 mix-ture theory, the total stress increment ( iso) is sepa-rated into partial stress increments for the solid phase ( s) and the fluid phase ( ƒ) (Oka, 1996). Consid-ering that the macroscopic volumetric strains by two phases are equal but of opposite sign for undrained Copyright © 2005 John Wiley Sons Retrieved from: www.knovel.com
  • 19. 188 7 EFFECTIVE, INTERGRANULAR, AND TOTAL STRESS θ Solid Surface Material θ Copyrighted Air Water (reference fluid) (a) Air Water (reference fluid) Solid surface (b) (c) Air Water Solid Figure 7.10 Wettability of two fluids (water and air) on a solid surface: (a) contact angle less than 90, (b) contact an-gle more than 90, and (c) unsaturated sand with water as the wetting fluid and air as the nonwetting fluid. conditions, Oka (1996) showed that the partial stresses are related to the total stress as follows: C Cs ƒ iso (C/n) (1 1/n)C C s l (7.44) [(1/n) 1]C (C /n) C s l s iso (C/n) (1 1/n)C C s 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 fluid. If the excess pore pressure generated by undrained isotropic loading is u, the partial stress increment for the fluid phase becomes (Oka, 1996) n u (7.45) ƒ Combining Eqs. (7.45) and (7.46), C Cs u (7.46) C C n(C C ) iso s l s The multiplier in the right-hand side of the above equation is in fact Bishop’s pore water pressure coef-ficient 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 Terzaghi’s effective stress equation is also applicable for undrained conditions for most soils. 7.11 WATER–AIR INTERACTIONS IN SOILS Wettability refers to the affinity of one fluid for a solid surface in the presence of a second or third fluid or gas. A measure of wettability is the contact angle, which was introduced in Eq. (7.9). Figure 7.10 illus-trates 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 fluid (air for Fig. 7.10a and water for 7.10b). The interface between the two fluids meets the solid surface at a contact angle
  • 20. . If the angle is less than 90, the reference fluid is referred to as the wetting fluid for a given solid surface. If the angle is greater than 90, the reference liquid is referred to as the nonwetting phase. The figure shows that water and 7A similar equation for B value has been proposed by Lade and de Boer (1997). air are the wetting and nonwetting fluid, respectively.8 The environmental SEM photos in Fig. 5.27 showed that water can be either wetting or nonwetting fluid depending soil mineralogy. The contact angle is a property related to interac-tions of solid and two fluids (water and air, in this case). as ws cos
  • 21. (7.47) aw 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 8Some 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 fluid with respect to NAPL and that NAPL is a wetting fluid with respect to air, implying that the wettability order is water NAPL air. Below the water table, water is the wetting fluid and NAPL is the nonwetting fluid. Copyright © 2005 John Wiley Sons Retrieved from: www.knovel.com
  • 22. WATER–AIR INTERACTIONS IN SOILS 189 106 105 104 Material 103 102 101 100 Copyrighted 7 5 3 2 0.0 0.1 0.2 0.3 0.4 0.5 0.6 Volumetric Water Content θw 10-1 1 4 6 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 Matric suction ua – uw (kPa) Figure 7.11 Soil–water characteristic curves for some Dutch soils (from Koorevaar et al., 1983; copied from Fredlund and Rahardjo, 1993). 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 sur-faces. It has been observed for movement of water rel-ative 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 sepa-rating two fluids. This pressure difference exists be-cause of the interfacial tension present in the fluid– fluid interface. Matric suction is a property that causes porous media to draw in the wetting fluid and repel the nonwetting fluid and is defined as the difference between the nonwetting fluid pressure and the wetting fluid pressure. For a two-phase system consisting of water and air, the matric suction is u u (7.48) n w where un is the pressure of the nonwetting fluid (air) and uw is the pressure of the wetting fluid (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 propor-tional to the interfacial tension and inversely propor-tional to the radius of curvature. It follows that higher air pressure is required for air to enter water-saturated fine-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 soil–water characteristic relationship in which the matric suction is plotted against the volu-metric 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 water-saturated soil. As the volumetric water content de-creases, as a result of drainage or evaporation, the matric suction increases. When water infiltrates into the soil (wetting or imbibition), the conditions reverse, with the volumetric water content increasing and ma-tric suction decreasing. Usually drainage and wetting 9The soil–water characteristic curve is referred to by a variety of names depending on different disciplines. They include moisture re-tention, soil–water retention, specific retention, and moisture char-acteristic. processes do not follow the same curve and the volu-metric water content versus matric suction curves ex-hibit 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-filled 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 points—the air entry pressure a and residual volu-metric water content
  • 23. r as defined in Fig. 7.12a. The entry pressure is the matric suction at which the air begins to enter the pores and the pores become inter-connected (Corey, 1994). At this point, the air per-meability becomes greater than zero. Corey (1994) also introduced the term ‘‘displacement pressure’’ (d in Fig. 7.12b) and defined it as the matric suction at which the first water desaturation occurs during a drainage cycle.10 The entry pressure is always slightly 10For the Dense NAPL–water two-phase system (often Dense NAPL is the nonwetting fluid and water is the wetting fluid), the displace-ment pressure may be important to examine the potential of DNAPL invading into a noncontaminated water-filled porous media. Copyright © 2005 John Wiley Sons Retrieved from: www.knovel.com
  • 24. 190 7 EFFECTIVE, INTERGRANULAR, AND TOTAL STRESS Hysteresis Scanning Curve Initial drainage Curve Water Content Material Copyrighted Scanning Curve Main Wetting Curve Main Drying Curve (a) θr ψa θr Residual Water Content ψa Air Entry Value Draining Wetting (b) ψd ψd Displacement pressure Suction Figure 7.12 Hysteresis of a soil–water 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 fine-grained than for coarse-grained soils because of their smaller pore sizes. Residual water content
  • 25. r is defined 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 de-scribed 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 defined here is a mathe-matical fitting parameter without a specific quantitative value. The shape of the soil–water characteristic curve de-pends on many factors, including the grain size distri-bution, 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 dis-tinct 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 de-creases and the middle part is not horizontal. Many algebraic formulas have been proposed to fit the mea-sured soil-water characteristic relations. The most pop-ular ones are (a) the Brooks–Corey (1966) equation:
  • 26. when (7.49) m d
  • 27. r 1/ when (7.50) d d
  • 28. m
  • 30. m is the volumetric water content at full saturation and is the curve-fitting parameter called the pore size distribution index and (b) the van Gen-uchten equation (1980): 1 /m 1m
  • 31. r 1 (7.51) 0
  • 32. m
  • 33. r where 0 and m are curve-fitting parameters. Various modifications have been proposed to these equations to include behaviors such as hysteresis, non-wetting fluid 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- Copyright © 2005 John Wiley Sons Retrieved from: www.knovel.com
  • 34. EFFECTIVE STRESS IN UNSATURATED SOILS 191 Limitations in Bishop’s equation were highlighted by Jennings and Burland (1962) in their experiments investigating the volume change characteristics of un-saturated compression curve of air-dry silt falls above that of saturated silt. Also, as shown in the figure, some air-dry samples were consolidated at four different pres-sures (200, 400, 800, and 1600 kPa) and then soaked. Material 0.84 0.80 0.76 0.72 0.68 Copyrighted 0.64 soils. Figure 7.14 shows that the oedometer Initially Soaked Test Air Dry (8 specimens) Soaked at Constant Void Ratio Soaked at Constant Applied Pressure 10 100 1000 Applied Pressure (kPa ) Figure 7.14 Oedometer compression curves of unsaturated silty soils (after Jennings and Burland, 1962 in Leroueil and Hight, 2002). _uw)b = Air Entry Value _uw)/(ua (ua _uw)b (ua trolled by an effective stress that is not the same as the total stress, the appropriate formulation for the effec-tive stress is less certain than for a fully saturated soil. As noted earlier, Bishop (1960) proposed Eq. (7.15) (assuming ): i u (u u ) (7.52) a a w 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 neg-ative pore pressures are known to increase strength and decrease compressibility. Using Eq. (7.52), the shear strength of unsaturated soil can be expressed as {( u ) (u u )}tan (7.53) a a w where is the effective friction angle of the soil. However, difficulties 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 to-gether into a single equivalent effective stress is un-certain (Toll, 1990). Degree of Saturation S (%) Void Ratio e 1. Compacted Boulder Clay 2. Compacted Shale 3. Breadhead silt 4. Silt 5. Silty clay 6. Sterrebeek silt 7. White clay (a) (b) Coefficient χ χ = (ua – uw) (ua – uw) – 0.55 Coefficient χ 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 Retrieved from: www.knovel.com
  • 35. 192 7 EFFECTIVE, INTERGRANULAR, AND TOTAL STRESS The void ratio decreased upon soaking and the final state was very close to the compression curve of the saturated silt. Additional tests in which constant vol-ume during soaking was maintained by adjusting the applied load were also done. Again, after equilibrium, the state of soaked samples was close to the compres-sion curve of the saturated silt. Soaking reduces the suction and, hence, Bishop’s effective stress decreases. This decrease in effective stress should be associated with an increase in void ratio. However, the experi-mental Copyrighted Material observations gave the opposite trend (i.e., a de-crease in void ratio is associated with irreversible compression). The presence of meniscus water lenses in the soil before wetting was stabilizing the soil struc-ture, which is not taken into account in Bishop’s equa-tion (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 ka-olin. 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 a( u ) b(u u ) (7.54) a a w Preconsolidation pressure 1.25 1.20 1.15 1.10 1.05 1.00 0.95 _ uw (kPa) 300 kPa 200 kPa 25 50 100 200 400 σ_ua (kPa) ua 100 kPa 0 kPa Curves are Averages of Several Tests Void Ratio e Figure 7.15 Isotropic compression tests of compacted kaolin (after Wheeler and Sivakumar, 1995 in Leroueil and Hight, 2002). in which a and b are material parameters that may also depend on degree of saturation and stress. For exam-ple, Fredlund et al. (1978) propose the following equa-tion: ( u )tan (u u )tan b (7.55) a a w where b is the angle defining the rate of increase in shear strength with respect to soil suction. An example of this parameter as a function of water content, fric-tion angle, and matric suction is given by Fredlund et al. (1995). Similarly, the change in void ratio e of an unsat-urated soil can be given by (Fredlund, 1985) a ( u ) a (u u ) (7.56) t a m a w where at is the coefficient of compressibility with re-spect to changes in ua and am is the coefficient of compressibility with respect to changes in capillary pressure. A similar equation, but with different coef-ficients, 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 modifications 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; Gallip-oli 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). Bishop’s parameter in Eq. (7.52) is a scalar quan-tity, 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 11A microstructural analysis by Li (2003) suggests the following ef-fective stress expression: u (u u ) ij ij a ij ij a w Copyright © 2005 John Wiley Sons Retrieved from: www.knovel.com
  • 36. QUESTIONS AND PROBLEMS 193 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 de-pendent 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 be-comes Copyrighted Material 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 dif-ferent from u0, as discussed in Section 7.9. The con-tribution of the long-range interparticle forces to mechanical behavior of unsaturated clayey soils re-mains to be fully evaluated. 7.13 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 ma-terials and not as continua, knowledge of these stresses and of the factors influencing them is a necessary pre-requisite to the understanding and quantification of compressibility, deformation, and strength in constitu-tive relationships for behavior. Various interparticle forces have been identified 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, influenced by the composition and ambient conditions. The relationships between soil structure and effective stress are devel-oped further in Chapter 8. Chemical, electrical, and thermal influences on effective pressures and fluid pressures in soils have not been considered in the de-velopments in this chapter. They may be significant, however, as regards soil structure stability fluid flow, volume change, and strength properties. They are an-alyzed 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. Inclu-sion of the effect of pore water suction and air or gas pressure on the mechanical behavior of unsaturated soils requires modification of the effective stress equa-tion used for saturated soils. Complications arise from the difficulty in the choice of stress variables and in treatment of contact-level forces (i.e., capillary forces 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 specific 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 un-der hydrostatic condition. The suction–volumetric 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 satu-ration 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 Bishop’s equation (7.52) with Sw. Com-ment on the result. 2. Repeat the calculations done in Question 1 with soil 5 in Fig. 7.11. The specific 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 ce-ment, sphere, and interface. The soil has a particle diameter of 0.2 mm and the void ratio is 0.7. As-sume 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 Gs wd3, where Gs 1–6 is the specific gravity (say 2.65) and w is the unit weight of water. b. Contact force by external load, N d2, where is the external confining 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). Copyright © 2005 John Wiley Sons Retrieved from: www.knovel.com
  • 37. 194 7 EFFECTIVE, INTERGRANULAR, AND TOTAL STRESS c. Long-range van der Waals attraction force, A ˚ AAd/(24t 2), where Ais the Hamaker constant hh (Section 6.12) and t is the separation between particles (Israelachvili, 1992, from Santamarina, 2003). Use A 1020 N-m and t 30 . h 5. Discuss why it is difficult to measure suction using a piezometer-type tensiometer for long-term moni-toring of pore pressures. Describe the advantages of Areas in 1965. Copyrighted Material other indirect measurement techniques such as po-rous ceramic and dielectric sensors. 6. For the following cases, compare the effective stresses calculated by the conventional Terzaghi’s equation and by the modified equation (7.39) with values presented in Fig. 7.8. Discuss the possible errors associated with effective stress estimation by Terzaghi’s 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. 7. Give a microscopic interpretation for why an un-saturated soil can collapse and decrease its volume upon wetting as shown in Fig. 7.14 even though the Bishop’s effective stress decreases. 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 poten-tial. 12 Discuss the values of each component of the above equation in the matrix pores and the intraag-gregate pores. 12 This was proposed by a Review Panel in the Symposium on Mois-ture Equilibrium and Moisture Changes in Soils Beneath Covered Copyright © 2005 John Wiley Sons Retrieved from: www.knovel.com