Download as pdf or txt
Download as pdf or txt
You are on page 1of 17

Department of Construction Engineering Advanced Soil Mechanics

Chaoyang University of Technology -- General KT Theory --

UNIT 3
APPLICATION OF TERZAGHI’S THEORY
OF ONE DIMENSIONAL CONSOLIDATION
TO PROBLEMS INVOLVING VARIOUS STRESS SURFACES
AND DRAINAGE CONDITIONS,
TIME DEPENDENT LOADING,
AND LAYERED SYSTEMS

Prepared by Dr. Roy E. Olson on Spring 1989


Modified by Jiunnren Lai on Fall 2002

______

3.1 Introduction
Previously, Terzaghi's theory of one dimensional consolidation was applied to the case of
instantaneous loading, uniform distribution of initial excess pore water pressure with depth
(rectangular stress surface) and freely draining boundaries. Although this solution is the one
of most practical usefulness there are problems in foundation engineering where other
solutions should be used. As a first step in generalizing previous solutions, this set of notes
will contain solutions for cases in which the stress surface is not rectangular, and where the
boundaries may not be freely draining, the loading may be time dependent, the soil is
stratified, and a few other cases.

Several cases in which the earlier solutions do not apply will be presented to illustrate the
types of problems that may be investigated using the solutions given in this chapter. In all
cases the magnitude and rate of settlement are to be determined.

Case 1: At various, and unknown, times in the past, fill was placed over the surface of a
deposit of soft clay. It is now desired to add a final layer of compacted fill and to
construct industrial buildings on the site. Piezometers installed in the clay layer
indicated that excess pore water pressures remain in the clay from earlier loadings.
Therefore these remaining excess pore water pressures, which are not uniform
through the clay, have to be added to the uniform excess pore water pressure induced
by the new fill to obtain the total initial excess pore water pressure to be used in the
settlement analysis (Fig. 3.1a).

Case 2a: A compressible clay layer has been subjected to steady state seepage from an
underlying sand layer that is under artesian pressure. Then the pressure in the sand
layer is suddenly reduced to the static value (Fig. 3.1b).

Case 2b: A layer of soft organic silt is overlain and underlain by sand. As part of
construction work at the site, deep well pumps are used to dewater the lower sand
without influencing the water level in the upper sand (Fig. 3.1c).

33 2004/9/15
Department of Construction Engineering Advanced Soil Mechanics
Chaoyang University of Technology -- General KT Theory --

Fig. 3.1 Non-Rectangular Stress Surfaces Encountered in Engineering Practice

Case 3: An approach embankment for a large bridge is constructed over a deep deposit of
clay. A period of several years is involved in the placement of the fill. Under the
center of the embankment the consolidation is essentially one dimensional but the
loading is time dependent.

Case 4: A large grain-elevator structure with a mat foundation is to be constructed at a site


where there is a desiccated crust overlying a soft clay overlying a freely draining sand.
The thickness and coefficient of permeability of the crust are such that a significant
amount of pore water escaping from the soft clay will pass through the crust. In this
case, one boundary of the compressible soft clay is freely draining but the other is
neither freely draining nor impervious.

Solutions for these and a variety of other similar consolidation problems will be presented in
these notes.

34 2004/9/15
Department of Construction Engineering Advanced Soil Mechanics
Chaoyang University of Technology -- General KT Theory --

3.2 Compressible Layer of Thickness 2H Located Between Two Freely


Draining Boundaries with Initial Excess Pore Water Pressure a
Function of Depth
Because the boundary conditions are the same as those used previously, the analysis may
proceed directly from the following equation from the previous set of notes (valid for double
drainage):

∞ 
nπz  nπz
2H
1 1
u = ∑ ( ) ∫ u ( z,0) sin( )dz  sin( ) exp( − n 2π 2T ) (3.1)
n =1  H 0 2H  2H 4

where u(z,0), represents the initial distribution of excess pore water pressure, i.e., the stress
surface. Since only integration of simple functions is involved in finding the solutions, only
the final solutions will be presented. Equations will be presented for the isochrones and for
the average degree of consolidation. In all cases the average degree of consolidation is
defined in such a way that:

S = U Su (3.2)

3.2.1 Triangular Stress Surface

The term "triangular stress surface" is used to designate a stress surface of triangular shape
that has its base at one boundary and its apex at the other (Cases 2a and 2b). To avoid
unnecessary complications in the analysis, the depth coordinate, z, will always be measured
into the compressible layer from the apex of the triangular stress surface. The initial pore
water pressure at the base of the triangular stress surface will be designated as ub. The
solutions for the isochrones and average degree of consolidation are:

2u b nπ z 1
u= ∑
n =1, 2 , 3K nπ
( −1) n +1 sin(
2H
) exp( − n 2π 2T )
4
(3.3)

and:

2
U =1− ∑
m = 0 ,1, 2 ,K M
2
exp( − M 2T ) (3.4)

1 1 ct
in which M = π ( 2m + 1) = nπ and T = v 2 . The isochrones are plotted in Fig. 3.2.
2 2 H
The T-U relationship is exactly the same as that obtained in the case of a rectangular stress
surface.

3.2.2 Trapezoidal Stress Surface

The stress surface has a trapezoidal shape in cases where the initial excess pore water
pressure increases linearly from a finite value at one boundary to a greater value at the other
boundary. In practice this case will arise if a uniform distribution of initial excess pore
pressure is superimposed on the conditions described in Cases 2a and 2b. Hence the

35 2004/9/15
Department of Construction Engineering Advanced Soil Mechanics
Chaoyang University of Technology -- General KT Theory --

0.0

Elevation Relative to the Clay Surface (z/2H)


-0.2

-0.4
70%
95%
60%
-0.6
90% 50%
40%
80%
30%
-0.8
20%
0%
10%
-1.0
0.0 0.2 0.4 0.6 0.8 1.0
Dimensionless Excess Pore Water Pressure

Fig. 2 Isochrones for Triangular Stress Surface

solutions for isochrones are obtained by dividing the trapezoidal stress surface into its
rectangular and triangular components and obtaining two separate solutions which are then
added to obtain the isochrones. The T-U relationship is the same as for the rectangular
stress surface.

3.2.3 Sinusoidal Stress Surface

As consolidation of a homogeneous layer of compressible soil continues, the pore pressure


isochrones tend toward a sinusoidal shape. Thus, if an areal fill is applied at some time, or
times, in the past, there may be a residual sinusoidal distribution of excess pore water
pressure at the time you become involved in the project. In the case in which both
boundaries are freely draining, the excess pore water pressures are zero at the two boundaries
and increase sinusoidally to a value of us at the mid-depth of the compressible layer, Case 1.

When this distribution of excess pore water pressure is inserted into Eq. 3.1, all terms
disappear except the first one and the excess pore water pressure and average
degree of consolidation are given by:

πz 1
u = us sin( ) exp( − π 2T ) (3.5)
2H 4
1
and: U = 1 − exp( − π 2T ) (3.6)
4
The T-U curve is plotted in Fig. 3.3 where it may be compared with the T-U curve for a
linear stress surface. Values of T are presented in Table 3.1.

36 2004/9/15
Department of Construction Engineering Advanced Soil Mechanics
Chaoyang University of Technology -- General KT Theory --

-10

-20

Degree of Consolidation (%) -30


sinusoidal
-40

-50

-60

-70 linear

-80

-90

-100
.01 .1 1 10
Time Factor

Fig. 3 T-U Curves for Linear and Sinusoidal


Stress Surfaces

Table 3.1 T-U Relationships for the Two Most Useful Stress Surfaces

Average Degree of Time Factor for Time Factor for Sinusoidal


Consolidation, % Trapezoidal Stress Surface Stress Surface
5 0.0020 0.0208
10 0.0078 0.0427
20 0.0314 0.0904
30 0.0707 0.1446
40 0.1257 0.2070
50 0.1967 0.2809
60 0.2864 0.3714
70 0.4028 0.4880
80 0.5672 0.6523
90 0.8481 0.9332
95 1.1290 1.2141
99 1.7813 1.8664

3.2.4 T-U Relationship for Composite Stress Surfaces

As noted earlier, settlement analyses for cases with complex stress surfaces, such as
trapezoidal, can be simplified by dividing the stress surface into its simple components, such
as a rectangle and triangle, obtaining separate solutions for each component stress surface,

37 2004/9/15
Department of Construction Engineering Advanced Soil Mechanics
Chaoyang University of Technology -- General KT Theory --

and then adding the solutions. For practical purposes the T-U relationship is of main interest.
It is convenient to develop an equation for the calculation of a composite T-U relationship in
terms of the separate solutions for each of the component stress surfaces. The average
degree of consolidation is defined using the equation:
2H 2H

∫ u dz − ∫ u dz
i

U= 0
2H
0
(3.7)
∫ u dz
0
i

where i designates the initial condition. Equation 7 can be rewritten in terms of component
stress surfaces as:

N 2H 2H

∑ ∫ u dz − ∫ u dz
j =1 0
ij j

U= 2H
0
(3.8)
∫ u dz
0
i

where j designates the j-th component of the stress surface. Equation 8 can be simplified to:
2H

N ∫ u dz
ij

U = ∑U j 0
2H
(3.9)
∫ u dz
j =1
i
0

where Uj denotes the average degree of consolidation for the j-th component stress surface.
The composite average degree of consolidation is thus a weighted average. The most
common example of the formation of a composite stress surface in engineering practice was
given as Case 1 at the beginning of this chapter.

3.2.5 Consolidation Resulting from Seepage Pressures

The clay layer shown in Fig. 1d is initially in equilibrium with the water table at the surface
of the overlying sand layer. The site under construction is located in a valley.
Downstream of this site a dam is constructed and the reservoir is filled with water. For
simplicity it is assumed that the time needed to fill the reservoir is very small compared with
the time needed for the clay to consolidate. If the water surface at the site is raised by a
distance L, then the pore water pressures throughout the sand and clay layers is raised by an
amount Lγw. However, the lower sand has a distant drainage outlet such that its pore water
pressures are maintained at their original value. The higher pore water pressures in the clay
layer than cause water to flow from the clay into the lower sand and consolidation occurs.

The resulting settlement of the surface is easily calculated by dividing the clay layer into sub-
layers of suitable thicknesses, calculating the initial and final values of effective stress at their
mid-depths, determining the change of thickness of each using the procedures outlined in
Chapter 1, and summing to obtain the settlement.

38 2004/9/15
Department of Construction Engineering Advanced Soil Mechanics
Chaoyang University of Technology -- General KT Theory --

The remaining problem is to calculate the T-U relationship and the pore water pressure
isochrones. The excess pore water pressures, u, are here defined as those in excess of the
initial pore water pressures, before the water surface was raised. The thickness of the
compressible layer is 2H. The boundary conditions are:

1. u (0, t ) = ub
2. u (2 H , t ) = 0

and the initial condition is:

3. u ( z,0) = ub

The excess pore water pressure for this problem consists of two parts only one of which
dissipates:

u = u b (1 − z / 2 H ) + u ∆ (3.10)

where u ∆ is the part that dissipates. Equation 3.10 satisfies the differential equation of
one-dimensional consolidation. The analysis proceeds, as in the previous set of notes, by
substituting F(z)G(t) for u∆ and applying the boundary and initial conditions. The solution
for the isochrones is:

z 2u nπz 1
u = ub (1 − ) + ∑ b ( −1) n+1 sin( ) exp( − n 2π 2T ) (3.11)
2H nπ 2H 4

The terms within the summation sign are identical to those for the full triangular stress
surface with double drainage (Eq. 3). Thus, the pore water pressure isochrones are obtained
directly from previous analysis (Fig. 2).

Since all terms involving time factor are the same as for the triangular stress surface with
double drainage, the T-U relationship is also identical and the time-settlement curve is easily
obtained.

3.3 Time Dependant Loading


3.3.1 Introduction

Previous analyses have applied to cases in which the loading time was so small, compared to
the times required to dissipate the excess pore water pressures, that the loading could be
assumed instantaneous. In some problems of practical interest this assumption cannot be
made (Case 3). For such problems the usual practice is to divide the load-time diagram into
a suitable number of subdivisions, assume that the load applied during each subdivision is
applied instantaneously, calculate the settlement-time curve for each such loading, and then
add these settlement-time curves to obtain an estimate of the actual time rate of settlement.

Although the foregoing procedure yields solutions of satisfactory accuracy, it is often simpler
to utilize solutions that have been developed by dividing the load-time diagram into
differential elements and integrating (Terzaghi and Frohlich, 1936; Olson, 1977). Although

39 2004/9/15
Department of Construction Engineering Advanced Soil Mechanics
Chaoyang University of Technology -- General KT Theory --

one-dimensional solutions for time-dependent loading can be obtained without great


difficulty for a variety of stress surfaces and shapes of load-time curves, it appears that the
case of most practical interest is a rectangular stress surface, single or double drainage, and a
loading that increases (or decreases) linearly with time up to the end of construction and then
remains constant; this will be the only case considered in this section. Other solutions will
be found in Terzaghi and Frohlich (1936).

3.3.2 Consolidation of a Doubly Drained Clay Layer Subjected to a Rectangular Stress


Surface Produced by a Uniform Surface Pressure that Increases Linearly with Time

A clay layer of thickness 2H is enclosed between two freely draining boundaries. The
applied pressures are uniform across the horizontal surface so that the problem is one
dimensional and the stress surface is rectangular. The surface pressure is increased at a
uniform rate for a period of time tc to a value of qc which is then maintained constant
indefinitely. The loading diagram is shown in Fig. 3.4. It is assumed that every differential

dq
qc
dt(i)

t(i) tc t

Fig. 3.4 Time-Dependant Loading Diagram

increment of loading causes a uniform increment of excess pore water pressure through the
clay layer. Thus:

qc
dui = dq = dti (3.12)
tc

As a matter of convenience, a dimensionless time, ν, defined by the following equation:


t
ν= t (3.13)
c

is used.

It is assumed that the properties of the clay remain constant during loading and consolidation
so that the principle of superposition can be applied. The excess pore water pressure, du,

40 2004/9/15
Department of Construction Engineering Advanced Soil Mechanics
Chaoyang University of Technology -- General KT Theory --

remaining at some depth z in the layer at some time t-ti after application of a differential
excess pore water pressure, dui, at time ti, is:

2dui Mz  − M 2 c v ( t − ti ) 
du = ∑ sin( ) exp   (3.14)
m =0 M H  H2 

To obtain the excess pore water pressures remaining from all previous differential loadings,
dui, is replaced by (qc/tc)dti and Eq. 3.14 is integrated. The integration is performed
separately for n < 1 and ν > 1.

For ν < 1, Eq. 3.12 is written:

u t ∞
2 qc Mz  − M 2 cv ( t − t i ) 
∫0 du = ∫0 m∑=0 Mtc sin( H ) exp  H 2 dti

(3.15)

Term-by-term integration yields:

[ ]

2qc Mz
u =∑ 3
sin( ) 1 − exp( − M 2T ) (3.16)
m = 0 M Tc H

1 cv t c cv t
in which M = 2 π (2m+1), Tc = 2 and T = 2 .
H H

The average degree of consolidation, U, is defined as:


2H
2 Hνqc − ∫ u dz 2H
1
U=
2 Hqc
0
=ν −
2 Hqc ∫ u dz
0
(3.17)

This particular definition was chosen so that the settlement of the surface of the compressible
layer would equal the average degree of consolidation times the ultimate settlement under the
load qc. Equation 16 is inserted into Eq 3.17 to obtain:

U=
T  2 ∞ 1
[
1 − ∑ 4 1 − exp( − M T ) 
2
] (3.18)
Tc  T m =0 M 

For ν ≥ 1, a similar analysis leads to the following equations:

[ ]

2qc Mz
u =∑ 3
exp( M 2Tc ) − 1 sin( ) exp( − M 2T ) (3.19)
m = 0 M Tc H

and:

∑ M [exp( M ]

2 1
U = 1− 4
2
Tc ) − 1 exp( − M 2T ) (3.20)
Tc m =0

41 2004/9/15
Department of Construction Engineering Advanced Soil Mechanics
Chaoyang University of Technology -- General KT Theory --

The T-U relationship are plotted in Fig. 3.5 for a wide range of values of the construction
time factor, Tc. It is apparent that time-dependent loading should be taken into account if
predictions of the time-settlement curve are needed before times equal to about 5 tc.

Fig. 3.5 T-U Relationships for Various Values of the Construction Time Factor

42 2004/9/15
Department of Construction Engineering Advanced Soil Mechanics
Chaoyang University of Technology -- General KT Theory --

3.3.3 Discontinuous Loading

A common problem in practice is that of calculating the time-settlement curve for an


embankment that is constructed at a discontinuous rate. For example, the loading may
increase linearly during the first construction season, remain constant during the winter,
increase linearly during the second season, and then remain constant. The solutions are
obtained, for such problems, by applying the foregoing equations to each stage of time-
dependent loading separately and adding the settlements to obtain the total time-settlement
curve. A composite T-U curve can be calculated using Eq. 9 with the weighting factor
defined as the stress applied by any one ramp load divided by the applied stress used in
calculating the ultimate settlement.

3.4 Partially Draining Boundaries


34.1 Introduction

In some problems of practical interest the main source of settlement is a single more-or-less
homogeneous layer which is bounded above and/or below by layers that contribute little to
total settlement but which have coefficients of permeability and thicknesses such that they
can be considered neither freely draining nor impervious. In real cases, of this type, the
properties of the partially draining boundaries are likely to be too ill-defined to make an
analysis useful. However, the analysis is relatively simple and will be presented for
possible use in answering "what if" questions.

The idealized case to be considered, Fig. 3.6, consists of a compressible


layer of thickness L bounded by two incompressible layers of thickness H1 and H2 and
coefficients of permeability k1 and k2, respectively. The incompressible layers are, in turn,
bounded by freely draining layers. Consolidation is assumed to be one dimensional. A
solution will be developed for a generalized stress surface with instantaneous loading.
Equations will be presented for the special case of a rectangular stress surface. These

Freely draining boundary

Incompressible Layer No.2 H2

Compressible Layer L
z

Incompressible Layer No.1 H1

Freely draining boundary

Fig. 3.6 Soil Profile Used for Analysis of a Single Compressible


Layer with Partially Draining Boundaries

43 2004/9/15
Department of Construction Engineering Advanced Soil Mechanics
Chaoyang University of Technology -- General KT Theory --

equations will then be integrated to account for time dependent loading. Solutions for
nonrectangular stress surfaces and diverse load-time relationships can be obtained from the
solutions presented.

3.4.2 Instantaneous Loading

Since the assumptions regarding the consolidation behavior of the compressible layer are the
same as those made previously, the analysis may begin directly from:

u = [C4 cos( Az ) + C5 sin( Az )] exp( − A2 cv t ) (3.21)

The boundary conditions are:

∂u (0, t ) R2
= u (0, t ) (3.22a)
∂z L

∂u ( L, t ) R1
= u ( L, t ) (3.22b)
∂z L

and the initial condition is:

u ( z,0) = f ( z ) (3.22c)

in which:
k1 L k2 L
R1 = and R2 = (3.23)
kH 1 kH 2

The dimensionless parameters R1 and R2 define the degree of perviousness of the


incompressible layers; a value of zero indicates an impervious boundary whereas a value of
infinity indicates free drainage. Use of these parameters was first suggested by Hamilton
Gray (1945).

Application of the boundary and initial conditions using the generalized Fourier series
(Appendix 3-A) leads to:

1 ∞
L
u= ∑ exp( −α n2T )Dn Z n ∫ Z n f ( z )dz
L n =1
(3.24)
0

in which αn represents successive positive roots, other than zero, of:

α n ( R1 + R2 )
tan α n = (3.25)
α n2 − R1R2
and:

2(α n2 + R12 )
Dn = (3.26)
(α n2 + R12 )(α n2 + R22 ) + (α n2 + R1 R2 )( R1 + R2 )

44 2004/9/15
Department of Construction Engineering Advanced Soil Mechanics
Chaoyang University of Technology -- General KT Theory --

z z
Z n = α n cos(α n ) + R2 sin(α n ) (3.27)
L L

The simplest, and certainly most useful, solution is for the case of a constant initial excess
pore water pressure, i.e., f(z) = ui . Integration of Eq. 3.24 for this case yields the following
solution:

u = ui ∑ Dn En Z n exp( −α n2T ) (3.28)
n =1

in which:

1
En = [α n sin(α n ) − R2 cos(α n ) + R2 ] (29)
αn

It is easily demonstrated that the solutions presented by Hamilton Gray (1945) and Bishop
and Gibson (1964) are special cases of Equation 28.

The average degree of consolidation is defined so as to preserve the validity of Eq. 3.2.
Thus Eq. 3.28 is integrated to obtain:

U = 1 − ∑ Dn En2 exp( −α n2T ) (30)
n =1

Representative T-U curves for R1 and R2 ranging from 0.1 to 100 are presented in Fig. 3.7.
For practical purposes the incompressible drainage layers may be considered impervious for
values of R less than 0.1 and freely draining for R in excess of 100. The curves in Fig. 3.7
make it possible to calculate time-settlement curves for most practical problems in which
there is a single compressible layer.

3.4.3 Time Dependent Loading

The solutions for the isochrones and T-U curves for instantaneous loading can easily be
expanded to include any desired variation of applied surface load with time by applying the
methods used earlier in this chapter. Again, it is assumed that the load-time curve of most
interest is given in Fig. 3.4. The solutions are as follows:

For ν ≤ 1

q Dn E n Z n
u= c
Tc

n =1 α 2
[1 − exp( −α n2 T )] (3.31)
n


1 Dn En2
U =ν −
Tc

n =1 α 2
[1 − exp( −α n2T )] (3.32)
n

For ν ≥ 1:

45 2004/9/15
Department of Construction Engineering Advanced Soil Mechanics
Chaoyang University of Technology -- General KT Theory --

Fig. 3.7 T-U Relationships for a Compressible Layer with Partially Drainage Boundaries

46 2004/9/15
Department of Construction Engineering Advanced Soil Mechanics
Chaoyang University of Technology -- General KT Theory --


qc Dn E n Z n
u=
Tc

n =1 α 2
[exp(α n2Tc ) − 1] exp( −α n2T ) (3.33)
n

1 ∞
Dn E n2
U = 1−
Tc

n =1 α 2
[exp(α n2Tc ) − 1] exp( −α n2T ) (3.34)
n

In Eqs. 3.32 and 3.34, U was defined so as to maintain the validity of Eq. 3.2 for calculating
the time-settlement curve.

The addition of the variable Tc precludes convenient presentation of numerical results.


However, solutions can easily be obtained using a digital computer.

3.5 Other Cases Including Various Initial and Boundary Conditions


Equation 3.24 can be used to obtain solutions for one dimensional consolidation of a
compressible layer with a variety of initial and boundary conditions. Solutions are easily
obtained for any distribution of initial excess pore water pressure that can be approximated
by an integrable function, for any degree of boundary drainage, and for any variation of
applied load with time.

3.5.1 Two-Layer Systems

Solutions for a system of two contiguous layers with freely draining or impervious external
surfaces was presented by Gray (1945). The solution is obtained by applying Terzaghi's
differential equation within each layer and, at the interface between the two layers, requiring
that there be a single excess pore water pressure and that the flow rates in the two layers be
equal:

∂u ∂u
ku ( ) = kl ( ) (3.35)
∂zu ∂zl

For the case of double drainage the T-U relationships are:

sin ξα n (α sin ξα n + sin α n )


U1 = 1 − 2∑ (1 − cosα n ) exp( −α n2T ) (3.36a)
α n (α sin ξα n + ξ sin α n )
2 2 2

2 sin α n (α sin ξα n + sin α n )


U 2 = 1− ∑
ξ α
(1 − cos α n ) exp( −α n2T ) (3.36b)
n (α sin ξα n + ξ sin α n )
2 2

Where αn represents successive positive roots of:

α cos αn sin ξαn + sin αn cos ξαn = 0 (3.36c)

47 2004/9/15
Department of Construction Engineering Advanced Soil Mechanics
Chaoyang University of Technology -- General KT Theory --

and:
H2 c v1
ξ= (3.36d)
H1 cv 2

k1
k
α= 2 (3.36e)
c v1
cv 2

c v 1t
T= (3.36f)
H 12

and cv, k, and H are the coefficients of consolidation and permeability, and the total layer
thickness, respectively, and subscripts 1 and 2 denote the two layers. Either layer (1 or 2)
can be on top. The time factor, T, is defined using the properties of layer 1, even in the
equation for layer 2.

For the case of single drainage, layer 1 must be next to the impervious boundary. The
solutions are:

sin α n cosα n sin ξα n


U1 = 1 − 2∑ exp( −α n2T ) (3.37a)
(α )(α sin ξα n + ξ cos α n )
2
n
2 2

2cos2 α n (1 − cos ξα n )
U2 = 1 − ∑ 2 exp( −α n2T ) (3.37b)
ξ (α n )(α sin ξα n + ξ cos α n )
2 2

where αn represents successive positive roots of:

α sin α n sin ξα n − cosα n cosξα n = 0 (3.37c)

and Eqs. 3.36d-3.36f remain valid.

It is convenient to define a composite U such that Eq. 3.2 remains valid, but with:

Su = ∆Sul + ∆Su2 (3.38)

and ∆Sul and ∆Su2 are the ultimate compressions of the two layers. It may be shown that:

(U1 + ζU2)
U= (3.39)
(1+ζ)

where:
∆Su2 H2 cvl k2
ζ= = (H )(c )(k ) (3.40)
∆Sul 1 v2 1

48 2004/9/15
Department of Construction Engineering Advanced Soil Mechanics
Chaoyang University of Technology -- General KT Theory --

(Gray used a different definition of U which invalidated Eq. 3.2).

Gray's equations have found no known application in engineering practice because they are
too difficult to solve in a specific case and because the number of variables is too large to
allow convenient presentation of general solutions. The equations can be (and have been)
solved using a digital computer but the effort is excessive considering the remaining
approximations (constant properties, small strains, instantaneous loading . . .).

Efforts have been made to approximate layered systems using an equivalent homogeneous
soil, e.g., Davis and Lee (1969). Such solutions are not acceptably accurate and serve little
purpose in an age of numerical methods. Such methods will not be discussed.

It may be noted that Gray's equations can easily be extended to cover time dependent loading
problems following the methods discussed earlier.

There appear to be no published closed-form solutions for systems composed of three or


more layers but a "solution" left in somewhat more general terms, was developed by
Schiffman and Stein (1970) for multilayered systems.

3.5.2 Other Cases

Analyses have been performed for a variety of other cases involving one-dimensional
primary consolidation. However, the resulting equations have generally been such that
closed form solutions could not be obtained. The authors thus formulated numerical
solutions (see later notes).

3.6 References
Bishop, A. W. and R. E. Gibson (1964), "The Influence of the Provisions for Boundary
Drainage on Strength and Consolidation Character-istics of Soil Measured in the
Triaxial Apparatus," pp. 435-451, ASTM STP 361.

Davis, E. H. and I. K. Lee (1969), "One Dimensional Consolidation of Layered Soils," Proc.
Seventh Intern. Conf. on Soil Mech. and Found. Engr., Mexico City, Vol. 2, pp. 65-72.

Glick, G. W. (1945), discussion, Transactions, ASCE, Vol. 110, pp. 1351-1352.

Gray, Hamilton (1945), "Simultaneous Consolidation of Contiguous Layers of Unlike


Compressible Soils," Transactions, ASCE, Vol. 110, pp. 1327-1344.

Olson, R. E. (1977), "Consolidation Under Time Dependent Loading," Jour., Geot. Engr.
Div., ASCE, Vol. 103, No. 1, pp. 55-60.

Schiffman, R. L. and J. R. Stein (1970), "One-Dimensional Consolidation of Layered


Systems," Jour., Soil Mech. and Found. Div., ASCE, Vol. 96, No. SM4, pp. 1499-1504.

Terzaghi, K. T. and O. K. Frohlich (1936), Theorie der Setzung von Tonschichten, Franz
Deuticke, Leipzig, 166 pp.

49 2004/9/15

You might also like