Download as pdf or txt
Download as pdf or txt
You are on page 1of 20

CONSOLIDATION AFTER UNDRAINED PIEZOCONE

PENETRATION. I: PREDICTION

By Jacques-Noel Levadoux, 1 A. M. ASCE


and Mohsen M. Baligh, 2 M. ASCE
Downloaded from ascelibrary.org by New York University on 02/13/15. Copyright ASCE. For personal use only; all rights reserved.

ABSTRACT: Piezocone measurements during steady cone penetration provide


valuable soil-profiling and soil-identification data. The potential of the piezo-
cone as a soil-exploration tool can be significantly enahnced if the pore-pressure
decay that takes place after interrupting penetration can be interpreted to es-
timate the consolidation and/or permeability of the soil. This paper provides
the necessary theoretical background to develop a rational interpretation method
and presents linear consolidation solutions to be used in analyzing dissipation
data in clays. Results show the two-dimensional aspects of the problem and
evaluate the effects of clay anisotropy, cone angle, porous sensing element lo-
cation, linear coupling, mesh size, and resolution as well as sensitivity of pre-
dictions to soil variability and measurement errors.

INTRODUCTION

The development of piezometer probes with rapid response enabled


meaningful measurements of pore pressures, u, to be obtained during
cone penetration testing (Janbu and Senneset, 1974; Wissa, et al., 1975;
Torstensson, 1975). These measurements together with cone resistance,
Cjc, data from Dutch cone penetrometers, were effectively utilized by Bal-
igh, et a l , (1978, 1980a, 1980b) and others for detailed soil profiling and
identification purposes. In view of the importance and potential of these
measurements a n e w device, the piezocone (Baligh, et al., 1981), capable
of measuring qc and u simultaneously was conceived, designed, and
christened at the Massachusetts Institute of Technology (MIT) in a joint
research effort with Geotechniques International, Inc. Currently, piezo-
cones are produced worldwide a n d are extensively utilized for soil pro-
filing and identification, especially in offshore applications.
The value of the piezocone in soil exploration can be significantly en-
hanced if the dissipation data obtained at the end of steady penetration,
when the soil consolidates, can be used to estimate the consolidation
a n d / o r permeability characteristics of soils. This paper provides the the-
oretical background and solutions n e e d e d to rationally interpret dissi-
pation data in clays. A companion article, Baligh and Levadoux (1986),
evaluates the prediction method in a Boston blue clay deposit by means
of comparisons with laboratory measurements a n d foundation perfor-
mance. Evaluations in other deposits are described by Baligh a n d Lev-
adoux, (1980), Wissa, et al. (1983), a n d Jamiolkowski, et al. (1985).
:
Mgr., Dames & Moore Consulting Engrs., 30 Avenue de l'Amiral Lemmonier,
78160 Marly le Roi, France.
2
Prof., Dept. of Civ. Engrg., MIT, Cambridge, MA 02139.
Note.—Discussion open until December 1, 1986. Separate discussions should
be submitted for the individual papers in this symposium. To extend the closing
date one month, a written request must be filed with the ASCE Manager of Jour-
nals. The manuscript for this paper was submitted for review and possible pub-
lication on November 1, 1985. This paper is part of the Journal of Geotechnical
Engineering, Vol. 112, No. 7, July, 1986. ©ASCE, ISSN 0733-9410/86/0007-0707/
$01.00. Paper No. 20784.

707

J. Geotech. Engrg. 1986.112:707-726.


TYPICAL DISSIPATION RECORDS

Fig. 1 shows typical excess pore pressure dissipation records at the tip
of an 18° cone at depths 20, 40, 50 and 68 ft in a deposit of Boston blue
clay (BBC) described elsewhere (Baligh, et a l , 1980a). The measured pore
pressure, u, after a time, t, from interrupting steady penetration de-
creases from an initial value M, to a final (hydrostatic) value u0.
Downloaded from ascelibrary.org by New York University on 02/13/15. Copyright ASCE. For personal use only; all rights reserved.

At any depth, the initial pore pressure, w,, for dissipation studies is
the steady penetration pressure at that depth. The latter can vary sig-
nificantly due to inherent soil variability. Furthermore, allowing for full
dissipation of excess pore pressures caused by cone penetration can cause
significant delays in the exploration program and, therefore, the final
pore pressure u0 is more often estimated than measured.
In dissipation studies, it is more meaningful to plot the decay of the
excess pore pressure Aw (= u — u0) especially when normalized by its
initial value AM, (= w, - u0). Using the same test results presented in Fig.
1, a plot of the normalized excess pore pressure, u (= AW/AM,), is shown
in Fig. 2. During consolidation, u decreases from unity to zero and pro-
vides a good measure of the degree of consolidation (= 1 — u). Fig. 2
clearly indicates that dissipation of excess pore pressures becomes much
slower with depth in the BBC deposit. At a depth of 20 ft most of the
dissipation (u = 0.2) is completed in 6 sec whereas at a depth of 68 ft
only 30% consolidation (u = 0.7) takes place after 100 sec.
In investigating dissipation results plotted as u versus log t, it is in-
teresting to note that, according to linear consolidation analyses based
on Terzaghi's (uncoupled) theory, two soils with the same normalized
distribution of initial excess pore pressures but with different values of
the coefficient of consolidation should have parallel dissipation curves.
Furthermore, the amount of horizontal shift required to reach one dis-
sipation curve from another represents the ratio between the coefficients
of consolidation for the two soils. This simple rule is quite useful in

Q I 1 1 1 1—I I I I I 1 1 1 I—I I I I I 1 1 I 1 — I I I )
10 100 1000
T I M E , sec

FIG. 1.—Typical Dissipation Records after Interrupting Steady Cone Penetration


in Clay

708

J. Geotech. Engrg. 1986.112:707-726.


Downloaded from ascelibrary.org by New York University on 02/13/15. Copyright ASCE. For personal use only; all rights reserved.

10 100 1000
T I M E , ice

FIG. 2.—Typical Normalized Dissipation Curves

assessing the importance of measurement scatter, evaluating solutions


and predictions, etc. When applied to Fig. 2 this rule indicates that:

1. The dissipation curve at a depth of 50 ft can be obtained by shifting


the curve at 68 ft by a horizontal distance to the left approximately equal
to 2. This suggests that linear (uncoupled) analyses are applicable at these
depths using the same distribution of normalized initial excess pore
pressures. Furthermore, the coefficient of consolidation at 50 ft is roughly
twice the value at 68 ft.
2. The dissipation curve at a depth of 20 ft cannot be obtained by
horizontally shifting the cuve at 68 ft. This indicates that either the initial
distribution of excess pore pressures is different or that linear (uncou-
pled) analyses are not applicable to the soil at shallow depths in this
deposit.

RATIONAL INTERPRETATION

A rational method for interpreting the pore pressure dissipation that


takes place after interrupting steady cone penetration should answer three
basic questions that are encountered in the analysis of any other labo-
ratory or in situ test: (1) What is (are) the measured soil parameter(s)?
e.g., vertical or horizontal coefficient of consolidation, permeability, or
both; (2) what is the magnitude of this parameter? and (3) what is the
range of practical applications to which it applies? The latter question
arises because of the complicated behavior of soils that usually prevents
the use of one parametric magnitude for all applications, e.g., the ap-
propriate undrained shear strength of a clay for a bearing capacity prob-
lem is, in general, different from that controlling the earth pressures on
a retaining wall.
Additional important questions that need special consideration in-
clude: (4) How long should dissipation be allowed? i.e., what degree of
consolidation is required to permit satisfactory interpretation? (5) what
709

J. Geotech. Engrg. 1986.112:707-726.


cone angle provides the most reliable results? (6) where to measure the
pore pressures, on the cone or the shaft behind it? and many others.
Theoretically, dissipation records are difficult to interpret because of
two major factors: (1) The uncertainties associated with the initial excess
pore pressure distribution in the soil caused by steady cone penetration;
and (2) the complicated aspects of soil behavior as related to soil con-
solidation around and behind cones, e.g., nonlinearities, anisotropy,
Downloaded from ascelibrary.org by New York University on 02/13/15. Copyright ASCE. For personal use only; all rights reserved.

nonhomogeneities caused by remolding due to penetration, and possi-


bly time-dependent (creep) behavior. Additional difficulties in analyses
are cased by: (1) The two-dimensional nature of consolidation around a
cone such that stresses and deformations depend on the radial as well
as the vertical location of a soil element; (2) the singularities and high
gradients near the tip of the cone; and (3) the coupling between total
stresses and pore pressures during consolidation.
Finally, note that analyses required to develop an interpretation method
are conceptually different from those needed for predicting the response
in a given clay under particular piezocone testing conditions. A useful
interpretation method should provide sufficient generality and hence
applicability to a wide range of soils. An accurate consolidation analysis
that faithfully incorporates soil nonlinearities might provide excellent
dissipation predictions in a particular soil but cannot be expected to ap-
ply to other deposits because of the significant differences in the non-
linear behavior of various soils and the large number of parameters re-
quired to describe behavior. Linear analyses provide valuable nor-
malizations and hence can be applied to a wide range of soils. There-
fore only linear analyses are presented in this paper in order to utilize
the generalizations offered by linear theories. Nonlinear consolidation is
briefly discussed in Appendix I. Simple considerations of undrained clay
behavior and results of nonlinear consolidation analyses around cylin-
drical shafts conducted by Kavvadas and Baligh (1982) indicate that the
coefficient of consolidation backfigured from dissipation data is probably
closer to the overconsolidated coefficient c(OC), of the clay than its nor-
mally consolidated value.

EXISTING METHODS

Few attempts have been made to interpret records of dissipation around


cones because on one hand conical piezometers with a sufficiently rapid
response have only recently been developed and because of the diffi-
culties in estimating the initial excess pore pressure distribution. Em-
pirical methods (e.g., Jones and Van Zyl, 1981; Tavenas, et al., 1982)
will not be discussed herein. On the other hand, the rational and com-
prehensive method presented by Torstensson (1977), which was slightly
modified by Battaglio, et al. (1981), is reviewed in some detail to provide
some insight and to show the complexities of the problem.
Torstensson (1977) suggests that the pore pressures in the soil caused
by steady cone penetration can be estimated by one-dimensional (radial)
solutions corresponding to cyclindrical and spherical cavities (Soderberg,
1962; Ladanyi, 1963). He assumes the soil to be isotropic, initially sub-
jected to an isotropic state of stress, to behave as an elastic-perfectly
plastic material during cavity expansion and utilizes linear uncoupled
710

J. Geotech. Engrg. 1986.112:707-726.


one-dimensional finite difference consolidation analyses to estimate soil
consolidation rates. In order to determine the coefficient on consolida-
tion, c, Torstensson proposes the matching of predictions and measure-
ments at 50% consolidation (w = 0.5) and hence uses the expression

150 '
Downloaded from ascelibrary.org by New York University on 02/13/15. Copyright ASCE. For personal use only; all rights reserved.

in which T50 is the time factor at u = 0.5 predicted by the theory as a


function of E„/s„ and the type of cavity (cylindrical or spherical); E„ and
s„ are "equivalent" undrained Young's modulus and shear strength of
the clay, respectively; i50 is the measured time to achieve 50% consoli-
dation; and R is an "equivalent" cavity radius.
From a theoretical standpoint, Torstensson's interpretation method does
not provide an answer to most of the important questions raised earlier.
In particular, the method fails to provide: (1) The necessary insight into
cone penetration mechanisms that are essential in understanding and
hopefully accounting for the complications caused by the actual behavior
of soils, e.g., nonlinearities during consolidation, soil straining (remold-
ing), consolidation under variable shearing, creep effects, etc.; (2) the
"type" of the estimated coefficient of consolidation c, i.e., does the com-
puted value apply for horizontal versus vertical consolidation and to the
overconsolidated versus virgin condition; (3) guidelines to select an ad-
equate cone angle and whether to locate the porous stone on the cone
or the shaft behind it; (4) a rational justification for selecting the initial
excess pore pressure distribution corresponding to cavity expansion in
an elastic-perfectly plastic material; and (5) an acceptable argument to
select 50% consolidation to estimate c since any dissipation curve drawn
arbitrarily (fi versus log T) will predict some versus value of c for a given
degree of consolidation.
In practical applications, the determination of c according to Torstens-
son requires estimates of the equivalent radius, R; the type of cavity
(cylindrical or spherical); and the rigidity index, E„/s„. The selection of
adequate values for these parameters is by no means simple, especially
in view of the obvious discrepancies between the two-dimensional pen-
etration process and the one-dimensional cavity expansion solutions.

INITIAL EXCESS PORE PRESSURES

Steady cone penetration in saturated clays causes undrained shearing


and develops excess pore water pressures in the soil. When penetration
is interrupted, the initial excess pore pressures governing subsequent
dissipation are determined by the pore pressure distribution in the soil
during penetration. Baligh and Levadoux (1980) discuss experimental
difficulties associated with determining the initial pore pressure distri-
bution in the soil by means of measurements. They conclude that, due
to the limited reliability of existing data and the importance of the initial
excess pore pressure distribution on dissipation rates, methods of inter-
pretation solely based on pore pressure measurements during penetra-
tion (Acar, et aL, 1982) involve significant uncertainties.

Importance of Initial Excess Pore Pressures


Levadoux and Baligh (1980) conducted an extensive parametric study
711

J. Geotech. Engrg. 1986.112:707-726.


to determine the importance of the initial distribution of excess pore
pressures on dissipation using closed form and numerical linear uncou-
pled solutions for one-dimensional consolidation problems involving im-
pervious cylindrical and spherical cavities. The results of their study in-
dicate that:
Downloaded from ascelibrary.org by New York University on 02/13/15. Copyright ASCE. For personal use only; all rights reserved.

1. Linear solutions are not affected by the absolute value (magnitude)


of the excess pore pressure but dissipation curves (plotted as u versus
log T) are very sensitive to the initial distribution of the normalized ex-
cess pore pressures as characterized by: the radius, XR, of the soil sub-
jected to excess pore pressures compared to the cavity radius, R; and
the spatial variation in the soil of the excess pore pressure at any radius
AUj(r) normalized by the shaft excess pore pressure Aut(R).
2. Analyses performed for a given initial distribution (X and spatial
variation) assuming spherical symmetry, lead to slightly faster dissipa-
tion than cylindrical symmetry (a factor of 1.5 to 2 in the backfigured
coefficient of consolidation at 50% dissipation). At a first glance, this
appears to contradict results presented by Torstensson showing that
spherical solutions predict much faster dissipation rates than cylindrical
solutions (a factor of 5, say). The difference arises because Torstensson
predicts the initial distributions by means of cavity expansion theories
that lead, for a given soil, to a much smaller value of X.
3. Dissipation is mainly controlled by the soil properties within the
radius \R and is little affected by the consolidation characteristics of the
outer soil. Furthermore, the soil in the vicinity of the cavity is predom-
inantly subjected to a decrease in volume (compression or recompres-
sion) during dissipation whereas the soil outside (r/R & X/2 to X de-
pending on the spatial variation) is subjected to swelling during most of
the dissipation of interest.

Strain Path Predictions


Undrained deep cone penetration in saturated clays is essentially a
"strain-controlled" problem where strains and deformations are not se-
riously affected by the shearing behavior of the soil. For this type of
problem, Baligh (1975, 1985) proposes a method of solution called the
strain path method (SPM). Based on the SPM, Levadoux and Baligh (1980)
determine the normalized excess pore pressures needed for linear con-
solidation analyses due to steady penetration of 18° and 60° cones in the
following manner:

1. Neglecting the shearing resistance of the soil, predict soil velocities


and strain rates using potential theory (for ideal incompressible fluid
flow) and a suitable distribution of sources and sinks to simulate the
geometry of the cone.
2. Integrate strain rates along stream lines to determine the strain path
(history) of soil elements as they move past the cone.
3. Compute the deviatoric stresses and the shear-induced pore pres-
sures using a total stress soil model MIT-T1 developed for this purpose
(Levadoux and Baligh, 1980). MIT-T1 can account for the complicated
strain paths of various soil elements (including large strains and re-
versals of strains), initial and stress-induced anisotropy as well as strain
712

J. Geotech. Engrg. 1986.112:707-726.


softening of clays. Predictions obtained by Levadoux and Baligh (1980)
correspond to soil properties determined from laboratory tests on re-
sedimented normally consolidated Boston blue clay.
4. Determine the excess pore pressure distribution by satisfying equi-
librium in the radial direction.
Downloaded from ascelibrary.org by New York University on 02/13/15. Copyright ASCE. For personal use only; all rights reserved.

In spite of uncertainties due to the approximate nature of strain path


solutions (see point 1) comparisons of predicted normalized excess pore
pressures with extensive field measurements in BBC at different loca-
i tions on the cone a n d the shaft behind it show excellent agreement for
both the 18° a n d 60° cones, and, surprisingly, in the overconsolidated
BBC as well {OCR < 3) (Baligh and Levadoux, 1986). Further compari-
sons with pore pressure measurements obtained by others d u e to pile
installation in different clays also show a surprisingly close agreement
(Baligh and Levadoux, 1980). This suggests that the initial normalized
excess pore pressures determined for BBC and utilized in all subsequent
\ consolidation analyses are reasonably applicable to other clays.

RESULTS OF CONSOLIDATION ANALYSES

Uncoupled Linear Isotropic Solutions for 18° Cone


The analyses presented in this section were conducted with the finite
element program ADINAT (Bathe, 1977) with a fine m e s h (538 elements)
to provide a sufficient resolution commensurate with the high gradients
in this problem. The solid lines in Fig. 3 show the predicted contours
of the excess pore pressures (Au/a„ 0 ) during uncoupled consolidation
(i.e., according to the Terzaghi-Rendulic theory) a r o u n d an 18° cone in
a linear isotropic material for selected values of the time factor; T (= ctj
R1 where c = the coefficient of consolidation; t = the time; a n d R = the
shaft radius). The pore pressures at T = 0 are predicted from steady cone
penetration analyses using the SPM and the properties of normally con-
solidated BBC as discussed earlier.
Examination of the results in Fig. 3 indicates the following:

1. At early times (T < 1, say) the effect of consolidation is limited to


the immediate vicinity of the tip. For example, t h e contour of h.u/dvo =
0.1 remains virtually u n c h a n g e d at T = 1 compared to its initial location
(at T = 0).
2. As consolidation proceeds, the contours initially close to the conical
tip move closer to the tip where they eventually vanish. This indicates
that AM decreases near the cone. O n the other h a n d , the contours ini-
tially far from the tip first move outwards (see, e.g., contours of AM/O^O
= 0.01 in Fig. 3) a n d then, at later times (not s h o w n in Fig. 3) come back
toward the tip where they eventually vanish. This indicates that AH in
soil elements close to the tip decreases monotonically whereas elements
further away are first unloaded (i.e., Au increases) a n d then reloaded.
Such results are important in the interpretation of dissipation records
because soils exhibit a nonlinear behavior with a coefficient of consoli-
dation during loadng that is different from that during unloading.

The solid lines in Fig. 4 show plots of normalized excess pore pres-

713

J. Geotech. Engrg. 1986.112:707-726.


Downloaded from ascelibrary.org by New York University on 02/13/15. Copyright ASCE. For personal use only; all rights reserved.

FIG. 3.—Excess Pore Pressures during Uncoupled Consolidation around 18° Cone
(r = c„t/R2)

sure, u, versus time factor, T, at four selected locations (at the cone apex,
midcone, cone base, and on the shaft at 10R behind the cone apex) along
the tip and the shaft of an 18° piezometer probe during uncoupled con-
solidation in a linear isotropic material.
We note in Fig. 4 that: (1) Dissipation rates at various locations are
significantly different especially at early consolidation stages (u large);
(2) the fastest dissipation occurs near the tip, location 1, because of the
high gradients near this singular point, and the slowest on the shaft is
at location 4; (3) at locations 2 and 4, the times required to reach u =
0.8, 0.5, and 0.2 are roughly in the ratios of 10, 5, and 2.5, respectively.
This means that, if the two-dimensional nature of the problem is ne-
glected and a consistent interpretation method based on one-dimen-
sional analyses is used (e.g., Torstensson method), the ratio of the coef-
ficient of consolidation, c, backfigured from measurements at midheight
of the cone (location 2) will be 10, 5, and 2.5 times higher than from
shaft measurements (location 4) when matching is done at 20, 50, and
80% consolidation, respectively. Larger differences in c (up to 100 times)
can result if dissipations at the tip (location 1) and the shaft (location 4)

714

J. Geotech. Engrg. 1986.112:707-726.


Downloaded from ascelibrary.org by New York University on 02/13/15. Copyright ASCE. For personal use only; all rights reserved.

TIME FACTOR T = chl/R2

FIG. 4.—Dissipation Curves for 18° Cone (Uncoupled Linear Analysis)

are compared. Clearly such discrepancy in c due to porous stone location


would be unacceptable in any reasonable interpretation method; (4) the
pore pressure at point 3 on the shaft right behind the cone base increases
slightly at early times. This is due to the larger initial pore pressures
along the cone face as shown by the contours in Fig. 3; and (5) at T =
100, most of the consolidation (82 to 92%) has been achieved (u = 18 to
8%) and at T = 1,000, consolidation is practically completed (u — 1%,
say) at all locations.

Effect of Anisotropy
Clays typically have a greater horizontal permeability, kh, than vertical
permeability, kv, and the coefficient of consolidation in the horizontal
direction is generally higher than in the vertical direction (i.e., c,, > cv
in most cases). This section investigates the effect of anisotropy on the
uncoupled consolidation around an 18° piezocone probe.
The dotted lines in Fig. 3 show the contours of excess pore pressure
during uncoupled consolidation around an 18° cone in a linear cross-
anisotropic material having cv = 0.1 ch as obtained with ADINAT and
where T = cht/R2. Compared to the solid contours representing the re-
sults in case of an isotropic soil, Fig. 3 shows that the effect of decreasing
cv from ch to 0.1 ch is very limited especially at early times, provided that
the time factor is defined as T = cht/R2. This result is important in the
interpretation of dissipation records by indicating that ch controls dissi-
pation, especially at early stages of consolidation.
Fig. 4 presents another illustration of the effect of soil anisotropy on
consolidation during uncoupled consolidation in a linear soil. Again, the
results in Fig. 4 clearly indicate that ch governs the consolidation process.
A tenfold reduction of cv slightly delays the pore pressure dissipation at
715

J. Geotech. Engrg. 1986.112:707-726.


the locations of interest along the cone. For example, the normalized
time factors, Tso, required to reach 50% dissipation (i.e., u = 0.5) are
respectively increased by 20, 34, 36, and 24% at locations 1, 2, 3, and 4
when cv decreases from ch to 0.1 ch. These values represent the error in
the estimated ch if, in interpreting dissipation records, soil anisotropy is
neglected and results at 50% consolidation are used.
Downloaded from ascelibrary.org by New York University on 02/13/15. Copyright ASCE. For personal use only; all rights reserved.

Effect of Cone Angle


Analyses identical to those previously presented were conducted for
the consolidation around a piezometer probe with a 60° tip angle in an
isotropic soil. The predicted contours of excess pore pressure during the
uncoupled consolidation around a 60° cone in a linear isotropic material
are given in Fig. 5. Results in this figure show basically the same trends
previously discussed for the 18° cone. We note, however, that because
of the pushing rather than cutting penetration mechanism of blunt cones,
the contours around the 60° tip are more spherical in shape, especially
at early times.
Fig. 6 shows plots of the normalized excess pore pressure, u (= AM/

FIG. 5.—Exeess Pore Pressures during Uncoupled Consolidation around 60° Cone
in Linear Isotropic Material

716

J. Geotech. Engrg. 1986.112:707-726.


1—r-f i i n i | 1 — I I M i ii | I—i i i m i l I—i i i i im 1—i i i i in
Downloaded from ascelibrary.org by New York University on 02/13/15. Copyright ASCE. For personal use only; all rights reserved.

0.01 0.1 I 10 100 1000

TIME FACTOR T ~
R1

FIG. 6.—Dissipation Curves for 60° Cone According to Linear Isotropic Uncoupled
Solutions

AM,), versus time factor, T, at four selected locations along the tip and
the shaft of a 60° piezometer probe during uncoupled consolidation in
a linear isotropic material. We note in Fig. 6 that dissipation at locations
1 and 2 are virtually identical and, in fact, are very close to that of point
3 located at the cone base. Furthermore, dissipation at point 4 located
on the shaft at a distance WR above the cone apex, is very similar to
the case of the sharp 18° tip (Fig. 4).

Effect of Linear Coupling


The solutions presented earlier were obtained according to the Ter-
zaghi-Rendulic theory neglecting the coupling between total stresses and
pore pressures during consolidation. This theory is rigorously applicable
to a narrow range of problems including one-dimensional consolidation
in semi-infinite masses where the soil is assumed to be homogeneous,
linear, isotropic, and elastic (Sills, 1975).
This section investigates the effect of coupling during consolidation
around an 18° piezometer probe in a linear isotropic elastic soil skele-
ton by comparing results obtained with the finite element program
CONSOL [utilizing Biot's coupled theory and the formulation proposed
by Ghaboussi and Wilson (1971, 1973)] to those obtained by ADINAT
(utilizing the Terzaghi-Rendulic theory).
The finite element program CONSOL requires a much larger core stor-
age than ADINAT because of the larger number of unknowns and, hence,
consolidation analyses with CONSOL proved impossible to carry out
with the fine mesh utilized earlier. Consequently, a coarser mesh in-
volving 96 elements instead of 538 elements was used for the analyses
that follow. Details of the finite element meshes and analyses can be
717

J. Geotech. Engrg. 1986.112:707-726.


found elsewhere (Levadoux and Baligh, 1980).
The dotted contours in Fig. 7 show the decay of excess pore pressures
during coupled isotropic linear consolidation around an 18° cone when
the soil is saturated with an incompressible pore fluid as obtained with
CONSOL for a drained Poisson's ratio v = 0.25. The solid contours in
Fig. 7 denote the uncoupled analyses obtained by ADINAT using the
same coarse mesh and identical soil properties, time steps, etc. The re-
Downloaded from ascelibrary.org by New York University on 02/13/15. Copyright ASCE. For personal use only; all rights reserved.

sults in Fig. 7 indicate that, on a global scale, coupling effects are not
very significant and are limited to the immediate vicinity of the conical
tip.
Fig. 8 compares plots of noramlized excess pore pressures, u (= AM/
AH,-), versus time factor, T, at four locations along the tip and the shaft
of an 18° piezometer probe during uncoupled consolidation (solid lines)
and coupled consolidation (dashed lines) in a linear isotropic elastic ma-
terial. Both results were computed using the coarse mesh size.
Results in Fig. 8 indicate the following effects of linear coupling:

1. As expected, the pore pressure at location 4 is practically unaffected


by coupling because the one-dimensional cylindrical situation, for which
coupling has no effect (Sills, 1975), is approached.

SOLUTION W I T H A D I N A T ( L I N E A R - U N C O U P L E D )
SOLUTION WITH CONSOL I L I N E A R - C O U P L E D )

FIG. 7.—Effect of Coupling on Predicted Contours of Excess Pore Pressures dur-


ing Isotropic Consolidation around 18° Cone

718

J. Geotech. Engrg. 1986.112:707-726.


Downloaded from ascelibrary.org by New York University on 02/13/15. Copyright ASCE. For personal use only; all rights reserved.

I I 111 I I I M I il t I I I ml i 1 I I I 111

TIME FACTOR T = =r

FIG. 8.—Effect of Linear Coupling on Dissipation Curves for 18° Tip (Linear Iso-
tropic Analyses)

2. The coupled dissipation curve at location 3 exhibits Mandel-Cryer


effects at early times but comes very close to the uncoupled dissipation
curve when T > 1 (i.e., u < 0.92, approximately).
3. Coupling causes a faster dissipation at location 2. However, the
effect becomes small for u < 0.8 (about 20% at u = 0.5).
4. Coupling appears to be especially important at the apex of the cone,
location 1 at the tip of a perfect cone. However, numerical difficulties at
this singular point prevent definite conclusions to be reached.

Effect of Mesh Size


It is interesting to compare the effects of the mesh size on the uncou-
pled dissipation results obtained by means of ADINAT using a fine mesh
to those with the coarse mesh and shown by the solid lines in Figs. 4
and 8 respectively:

1. Results at location 4 are not affected by the mesh size (they differ
by less than 1%) for any time factor T.
2. Dissipation at locations 2 and 3 is very slightly affected at large
times (when u < 0.5) by the mesh size but is significantly affected during
early stages of consolidation.
3. Dissipation at the cone apex, location 1, is strongly affected by the
mesh size. For example, the time factor to reach 20% consolidation (u
= 0.8) is T = 0.065 for the fine grid and T = 0.38 for the coarse grid.
Therefore, results from the coarse mesh at 20% consolidation overesti-
mate the coefficient of consolidation by a factor of 5.85.

719

J. Geotech. Engrg. 1986.112:707-726.


This clearly demonstrates the need for a fine mesh to provide an ad-
equate resolution in the vicinity of the tip where large pore pressure
gradients occur especially in the early stages of consolidation.

Effects of Soil Variability Errors in Measurements, or Both


Dissipation results previously presented to be used for estimating the
Downloaded from ascelibrary.org by New York University on 02/13/15. Copyright ASCE. For personal use only; all rights reserved.

horizontal coefficient of consolidation, ch, are expressed in terms of the


normalized excess pore pressure u:
ku u - «0
u=— = (2)
Au,< Ui - M0
This section investigates the effect of uncertainties in the values of u0
and tii in order to select more reliable methods for estimating ch.
In practice, the static pore pressure, u0, should be measured at se-
lected depths by leaving the piezometer probe in the soil long enough
to achieve complete soil consolidation. This is, however, a time consum-
ing and expensive task that is often overlooked, leading to an error 8u0
in the estimated static pore pressure, u0, at any depth:
M0* = u0 + Sw0 (3)
where uQ is the "true" (or in situ) static pore pressure.
The apparent normalized excess pore pressure, u*, is thus given by:
H - ( K 0 + 8MO)
u* = (4)
u{ - (u0 + 8«0)
Fig. 9(a) shows the effects of the error 8M0 on the dissipation at midcone
of an 18° piezometer probe. The solid line represents the "true" dissi-
pation curve (identical to the solid curve 2 in Fig. 4) and the dashed
lines represent the "apparent" dissipation curves for 8M0/AM, = 20, 10,
-10, and —20% (AH, is the "true" initial excess pore pressure at time T
= 0). Clearly, the effect of 8«0 is more pronounced at late consolidation
stages.
On the other hand, errors in the penetration pore pressure, w,-, can
occur due to: (1) Inherent soil variability causing the recorded value
uf at the beginning of dissipation to be different from the relevant (av-
erage) value «,-; and (2) incomplete deairing of the piezometer or low
permeability of the clay causing a time lag in the measurements. The
measured penetration pore pressure, uf, thus becomes:
Uf = Ui + 8«; (5)
where M,- is the relevant penetration pore pressure.
The apparent normalized excess pore pressure, u*, is thus given by:
* u - u0
«* = (6)
(u{ + 8M,-) -U0

Fig. 9(b) shows the effect of 8M, on the dissipation at midcone of an 18°
piezometer probe. "Apparent" dissipation curves are given for 8M,/AM,
= 20, 10, -10, and -20%.
The results in Fig. 9 are valuable in providing guidelines for the pre-
720

J. Geotech. Engrg. 1986.112:707-726.


Downloaded from ascelibrary.org by New York University on 02/13/15. Copyright ASCE. For personal use only; all rights reserved.

0.01 1000

T I M E FACTOR

FIG. 9.—-Effect of Errors on Dissipation Curves at Midcone of 18° Piezometer Probe:


(a) Error in Static Pore Pressure; and (b) Error in Penetration Pore Pressure

721

J. Geotech. Engrg. 1986.112:707-726.


diction of the horizontal coefficient of consolidation, ch, from dissipation
measurements:

1. When u0 is reliably measured (i.e., 8u0 = 0), prediction of ch is best


achieved by matching measured and predicted dissipation curves at large
times where the effects of 8M, are least significant.
Downloaded from ascelibrary.org by New York University on 02/13/15. Copyright ASCE. For personal use only; all rights reserved.

2. When «,- is measured with a high level of confidence (i.e., 8w, = 0),
as indicated by the consistency and uniformity of measurements, pre-
diction of ch is reliably and economically achieved by early time match-
ing.
3. When errors in both u0 and «, are expected, prediction of ch at in-
termediate times (when u = 0.5, say) represents a resonable compro-
mise.

CONCLUSIONS

Estimates of in situ consolidation or permeability characteristics of clay


deposits can be achieved by adequately interpreting the excess pore
pressure dissipation that takes place after interrupting steady piezocone
penetration. Existing rational interpretation methods rely on one-dimen-
sional analyses and hence are at best confusing when applied to the two-
dimensional dissipation process around cones. On the other hand, a rig-
orous and complete analysis of piezocone dissipation data is a hope-
lessly difficult task because of the complicated behavior of soils, the dif-
ficulties in estimating the initial pore pressures prior to consolidation
(singularities, large strains, large gradients, etc.), as well as anisotropy,
coupling, disturbance, and nonlinear effects during consolidation.
This paper presents linear consolidation analyses aimed at achieving
a better understanding of the important factors affecting dissipation after
undrained piezocone penetration in clays and hence developing the
framework for a rational and reliable simplified interpretation method.
The initial distribution of excess pore pressures were predicted by the
strain path method (Baligh, 1985) using properties of resedimented nor-
mally consolidated Boston blue clay (BBC).
Results of linear uncoupled finite element analyses of isotropic con-
solidation around and behind a sharp 18° cone indicate that dissipation
rates at various locations are significantly different. The fastest dissi-
pation occurs at the tip where the gradients are very high near this sin-
gular point and the slowest dissipation takes place on the shaft behind
the cone. If the two-dimensional nature of the problem is neglected and
one-dimensional consolidation analyses are utilized to interpret dissi-
pation data, differences in the backfigured coefficient of consolidation
are typically on the order of 5 to 10 times but can reach 100 times de-
pending on the location of the porous stone where pore pressures are
measured and on the degree of consolidation achieved. Predictions also
indicate that, during consolidation, pore pressures in the soil located in
the vicinity of the cone decrease monotonically whereas, at some dis-
tance of the cone, excess pore pressures first increase and then vanish
with time.
Results of analyses conducted to evaluate the importance of various
factors follow.
722

J. Geotech. Engrg. 1986.112:707-726.


Anisotropy.—A tenfold decrease in the vertical coefficient of consol-
idation has minor effects on dissipation rates, especially at early con-
solidation stages. Practically, this means that dissipation is essentially
controlled by the horizontal coefficient of consolidation, ch, and that, by
neglecting soil anisotropy, the error in the estimated value of ch will be
small (less than 40% at 50% dissipation).
Cone Angle.—Dissipation around the blunt 60° cone exhibits the same
Downloaded from ascelibrary.org by New York University on 02/13/15. Copyright ASCE. For personal use only; all rights reserved.

trends as the sharp 18° cone with the exception that pore pressure con-
tours around the blunt cone are more spherical in shape and involve
milder gradients near the tip. As a result, dissipation rates around the
blunt cone are less sensitive to the exact location of the porous stone on
the cone face and less susceptible to computational errors due to insuf-
ficient resolution.
Coupling and Resolution.—Comparisons between coupled and un-
coupled linear consolidation solutions around the sharp 18° cone indi-
cate that, after 20% consolidation, coupling between total stresses and
pore pressures has a relatively minor effect (less than 50%) on dissipa-
tion rates except at the cone tip where the effect is very significant es-
pecially at early consolidation stages. Also, the very large gradients near
the tip of the sharp cone lead to important effects of the mesh size (or
the resolution in numerical computations) on dissipation rates at the tip.
Errors.—Uncertainties in the penetration pore pressures, ut, due to
soil variability or imperfect deairing of the probe, and errors in the static
pore pressures, u0, due to incomplete dissipation can seriously affect
the interpreted coefficient of consolidation ch. In cases where ut is mea-
sured reliably, acceptable values of ch can be achieved from early dissi-
pation data. On the other hand, when «, cannot be estimated reliably,
long dissipation records are required.

APPENDIX I.—NONLINEAR CONSOLIDATION

Linear analyses assume that the coefficient of consolidation, c, does


not change during consolidation. This requires the ratio of permeability
to compressibility of the soil to remain constant during consolidation. In
reality, the compressibility of the clay at any consolidation stage is a
function of its stress history, stress level, and stress increment, whereas
its permeability depends on its current void ratio.
When the clay is in a normally consolidated state, the changes in com-
pressibility and permeability are self compensating such that c remains
sensibly constant. On the other hand, if a clay is loaded to effective
stress levels exceeding the maximum past pressure, 6\,m, or is unloaded
near &mn, severe nonlinearities arise during consolidation. Therefore, lin-
ear solutions are reasonably applicable to normally consolidated states
or to overconsolidated states involving relatively small stress increments
away from avm (i.e., when the OCR is large). Added difficulties to the
rigorous interpretation of dissipation data are caused by the effects of
disturbance (or remolding) caused by penetration and the possible changes
in shear (deviatoric) stresses during consolidation. Disturbance increases
the compressibility of the clay in the overconsolidated range as ex-
pressed by the recompression ratio, RR, and decreases the compressi-
bility in the normally consolidated range as expressed, say, by the

723

J. Geotech. Engrg. 1986.112:707-726.


compression ratio, CR. The increase in shearing stresses during consol-
idation can result in a softer behavior.
Cone penetration causes large non-uniform shearing strains in the soil,
especially in the vicinity of the cone. Levadoux and Baligh (1980) show
that these strains produce a failure zone extending a distance at least
four to six times the shaft radius from the cone. In a normally consoli-
dated clay, undrained shearing develops positive pore pressures, Aws,
Downloaded from ascelibrary.org by New York University on 02/13/15. Copyright ASCE. For personal use only; all rights reserved.

and hence causes a negative mean effective stress change. With regards
to subsequent clay consolidation, this reduction in mean effective stress
at a constant void ratio can be viewed as an artificial overconsolidation
of the clay insofar as subsequent clay consolidation is concerned. Fur-
thermore, undrained shearing of an overconsolidated clay generates
negative A«s and thus the mean effective stress is increased. However,
this increase is generally small compared to the difference between o-„,„
and the overburden effective stress, vv0 (typically 15 ± 5% for OCR =
4). Therefore, simple considerations of urtdrained soil behavior suggest
that pore pressure dissipation during early stages of consolidation of the
soil in the vicinity of cones takes place in a recompression mode (as
opposed to virgin compression) for both normally consolidated and
overconsolidated clays.
Kavvadas and Baligh (1982) conducted a comprehensive theoretical
study of the effects of soil nonlinearities on the consolidation of BBC
around the shaft behind a 60° cone. Their results, supported by in situ
simultaneous measurements of total stresses and pore pressures on model
pile shafts (Morrison, 1984), indicate that:

1. In the early consolidation stages (u a 0.65 for a normally consoli-


dated clay OCR = 1, and u a 0.3 for a clay with OCR — 1.35) dissipation
takes place in an "elastic" regime with no soil yielding. The clay in the
vicinity of the shaft is deformed in a reloading mode and the outer soil
is rebounded. Dissipation takes place rapidly because it is governed by
the overconsolidated coefficient of consolidation, C(OC), and minor vol-
ume changes take place in the soil.
2. At later consolidation stages, soil yielding causes minor deviations
in pore pressure dissipation rates because of two self-compensating phe-
nomena: (a) A plastic (soft) zone develops close to the shaft that tends
to retard dissipation; and (b) the soil stiffness increases due to an in-
crease in effective stresses during consolidation.

APPENDIX II.—REFERENCES

Acar, Y. B., Tumay, M. T., and Chan, A. (1982). "Interpretation of the Dissi-
pation of Penetration Pore Pressures," Int. Sx/mp. on Num. Models in Geome-
chanics.
Baligh, M. M. (1975). "Theory of Deep Static Cone Penetration Resistance," Pub-
lication No. R75-56, Order No. 517, Dept. of Civil Engineering, Massachusetts
Institute of Technology, Cambridge, Mass., Sept., 133 pp.
Baligh, M. M., Vivatrat, V., and Ladd, C. C. (1978). "Exploration and Evaluation
of Engineering Properties for Foundation Design of Offshore Structures," Pub-
lication No. R78-40, Order No. 607, Dept. of Civil Engineering, Massachusetts
Institute of Technology, Cambridge, Mass., Dec, 268 pp.
Baligh, M. M., and Levadoux, J. M. (1980). "Pore Pressure Dissipation after Cone

724

J. Geotech. Engrg. 1986.112:707-726.


Penetration," Research Report R80-11, Order No. 662, Dept. of Civil Engineer-
ing, Massachusetts Institute of Technology, Cambrige, Mass., 367 pp.
Baligh, M. M., Vivatrat, V., and Ladd, C. C. (1980a). "Cone Penetration in Soil
Profiling," Journal of the Geotechnical Engineering Division, ASCE, 106(GT4), Apr.,
pp. 447-461.
Baligh, M. M., Azzouz, A. S., and Martin, R. T. (1980b). "Cone Penetration Tests
Offshore the Venezuelan Coast," Research Report R80-21, Order No. 680, Dept.
of Civil Engineering, Massachusetts Institute of Technology, Cambridge, Mass.,
Downloaded from ascelibrary.org by New York University on 02/13/15. Copyright ASCE. For personal use only; all rights reserved.

Aug., 163 pp.


Baligh, M. M., et al. (1981). "The Piezocone Penetrometer," Proceedings, ASCE
Conference on Cone Penetration Testing and Experience, Geotechnical Engi-
neering Division, St. Louis, Mo., pp. 247-263.
Baligh, M. M. (1985). "Strain Path Method," Journal of Geotechnical Engineering,
ASCE, 111(GT9), Sept., pp. 1108-1136.
Baligh, M. M., and Levadoux, J. N. (1986). "Consolidation after Undrained Pi-
ezocone Penetration. II: Interpretation," Journal of Geotechnical Engineering, ASCE,
112(GT7), July, pp. 727-745.
Bathe, K. J. (1977). "ADINAT: A Finite Element Program for Automatic Dynamic
Incremental Nonlinear Analysis of Temperatures," Report 82448-5, Mechanical
Engineering Dept., Massachusetts Institute of Technology, Cambridge, Mass.,
May (Rev. Dec, 1978).
Battaglio, M., et al. (1981). "Piezometer Probe Test in Cohesive Deposits," Pro-
ceedings, ASCE Conference on Cone Penetration Testing and Experience, St.
Louis, Mo., pp. 264-302.
Ghaboussi, J., and Wilson, E. L. (1971). "Flow of Compressible Fluid in Porous
Elastic Media," UC-SESM Report No. 71-72, Structural Engineering Laboraotry,
Univ. of California, Berkeley, Calif., July, 33 pp.
Ghaboussi, J., and Wilson, E. L. (1973). "Flow of Compressible Porous Elastic
Media," Int. Journal for Numerical Methods in Engrg., 5, pp. 419-422.
Jamiolkowski, M., et al. (1985). "New Developments in Field Laboratory Tested
of Soils," 11th International Conference of Soil Mech. & Fdn. Engrg., San Francisco,
Calif., Vol. 1, pp. 57-154.
Janbu, N., and Senneset, K. (1974). "Effective Stress Interpretation of In Situ
Static Penetration Tests," Proceedings of the European Symposium on Penetration
Testing, I, Stockholm.
Jones, G. A., and VanZyl, D. J. A. (1981). "The Piezometer Probe—A Useful
Investigation Tool," 10th ICSMFE, Stockholm, Vol. 2, pp. 489-496.
Kavvadas, M., and Baligh, M. M. (1982). "Nonlinear Consolidation Analyses
Around Pile Shafts," Proceedings, BOSS '82, Cambridge, Mass., Vol. 2, Aug.,
pp. 338-347.
Ladanyi, B. (1963). "Expansion of a Cavity in a Saturated Clay Medium," Journal
of the Soil Mechanics and Foundations Division, ASCE, 89(SM4), July, pp. 127-
161.
Levadoux, J. N., and Baligh, M. M. (1980). "Pore Pressures During Cone Pen-
etration in Clays," Research Report R80-15, Order 66, Dept. of Civil Engineering,
Massachusetts Institute of Technology, Cambridge, Mass., Apr., 310 pp.
Morrison, M. (1984). "In Situ Measurements on a Model Pile in Clay," thesis
presented to Massachusetts Institute of Technology, in 1984, in partial fulfill-
ment of the requirements for the degree of Doctor of Philosophy.
Sills, G. C. (1975). "Some Conditions under which Biot's Equations of Consoli-
dation Reduce to Terzaghi's Equation," Geotechnique, London, 25(1), pp. 129-
132.
Soderberg, L. O. (1962). "Consolidation Theory Applied to Foundation Pile Time
Effects," Geotechnique, London, Vol. 12, pp. 217-225.
Tavenas, F., Leroueil, S., and Roy, M. (1982). "The Piezocone Test in Clays: Use
and Limitations," ESOPT 2, Amsterdam, Vol. 2, pp. 889-894.
Torstensson, B. A. (1975). "Pore Pressure Sounding Instrument," Proceedings, ASCE
Specialty Conference on In Situ Measurement of Soil Properties, Raleigh, N.C.,
Vol. 2, pp. 48-54.

725

J. Geotech. Engrg. 1986.112:707-726.


Torstensson, B. A. (1977). "The Pore Pressure Probe," Nordiske Geotekniske Mote,
Oslo, Paper No. 34, pp. 34.1-34.15.
Wissa, A. E. Z., Martin, R. T., and Garlanger, J. E. (1975). "The Piezometer
Probe," Proceedings, ASCE Specialty Conference on In Situ Measurement of
Soil Properties, Raleigh, N.C., Vol. I, pp. 536-545.
Wissa, A. E. Z., Fuleihan, N. F., and Ingra, T. S. (1983). Evaluation of Phosphatic
Clay Disposal and Reclamation Methods. Vol. 4. Consolidation Behavior of Phosphatic
Downloaded from ascelibrary.org by New York University on 02/13/15. Copyright ASCE. For personal use only; all rights reserved.

Clays, Florida Institute of Phosphate Research, Project FIPR 80-02-002.

726

J. Geotech. Engrg. 1986.112:707-726.

You might also like