You are on page 1of 14

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/237442438

An approach to evaluation of field CPTU dissipation data in overconsolidated


fine-grained soils: Reply

Article  in  Canadian Geotechnical Journal · September 1999


DOI: 10.1139/cgj-36-2-369

CITATIONS READS

70 1,169

4 authors, including:

John Sully P. K. Robertson


Jocarsu Consultants Inc. 144 PUBLICATIONS   10,081 CITATIONS   
18 PUBLICATIONS   373 CITATIONS   
SEE PROFILE
SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Fraser River Delta Landslide Processes View project

Quantification of Soil Heterogeneity View project

All content following this page was uploaded by John Sully on 16 September 2014.

The user has requested enhancement of the downloaded file.


369

An approach to evaluation of field CPTU


dissipation data in overconsolidated fine-grained
soils
John P. Sully, Peter K. Robertson, Richard G. Campanella, and
David J. Woeller

Abstract: Dissipation of excess pore pressures during piezocone testing in firm to stiff overconsolidated fine-grained
soils provides data curves that cannot be interpreted using published theoretical solutions. Available solutions are based
on either cavity-expansion or strain-path methods, which have been developed for soft, normally consolidated soils and
do not adequately model the response in overconsolidated soils. During penetration in overconsolidated soils, large
pore-pressure gradients and unloading can exist as the soil moves past the singularity at the cone tip shoulder; these
gradients modify the initial pore-pressure distribution around the tip and along the friction sleeve and give rise to
nonstandard dissipation curves. The different types of response resulting from the modified pore-pressure distribution
are discussed and classified and a correction technique is proposed. In this way, the application of available theoretical
models can be used to evaluate in situ flow characteristics. Examples of the different types of pore-pressure response
are presented from various sites worldwide.

Key words: cone penetration test, in situ, clay, overconsolidation ratio, pore pressure, dissipation, consolidation.

Résumé : La dissipation de l’excès des pressions interstitielles au cours de l’essai au piézocône dans les sols fins
surconsolidés de consistance moyenne à ferme fournit des courbes de données qui ne peuvent pas être interprétées au
moyen des solutions théoriques publiées. Les solutions disponibles sont basées sur les méthodes d’expansion de cavité
ou de cheminement de déformation qui ont été développées pour les sols mous normalement consolidés et qui ne
modélisent pas adéquatement la réponse dans les sols surconsolidés. Au cours de la pénétration dans les sols
surconsolidés, de forts gradients de pression interstitielle et un déchargement peuvent se produire lorsque le sol se
déplace le long de la singularité à l’épaulement de la pointe conique; ces gradients modifient la distribution initiale de
la pression interstitielle autour de la pointe et le long du manchon de frottement et génèrent des courbes de dissipation
non standard. Les différents types de réaction résultant de la distribution modifiée des pressions interstitielles sont
discutés et classifiés, et une technique de correction est proposée. Ainsi, l’on peut appliquer les modèles théoriques
disponibles pour évaluer les caractéristiques d’écoulement in situ. L’on présente des exemples des différents types de
réaction de pression interstitielle provenant de divers sites à travers le monde.

Mots clés : essai de pénétration au cône, in situ, argile, rapport de surconsolidation, pression interstitielle, dissipation,
consolidation.

[Traduit par la Rédaction] Notes 381

Background lowing soil response parameters: tip resistance (qc), sleeve


friction ( fs), and penetration pore pressure (u). In addition,
Cone penetration testing with pore pressure measurement inclination and temperature may also be recorded depending
(CPTU or piezocone testing) has become a popular investi- on the type of piezocone being used (Campanella and Rob-
gation technique in geotechnical site investigation practice. ertson 1988). The penetration of the piezocone can be halted
The near-continuous data obtained during penetration at any depth and the variation with time of the measured pa-
(approx. every 5 cm) provide information related to the fol- rameters can be monitored. Of the above three quantities (qc,
fs, u), it is usually the variation of the pore pressure that is of
Received June 17, 1998. Accepted November 2, 1998. interest, as the results can be interpreted to provide estimates
of the in situ horizontal coefficient of consolidation, ch
J.P. Sully. MEG Consulting Ltd., 6211 Comstock Road,
Richmond, BC V7C 2X5, Canada. (Torstensson 1977). Rather than the total pore pressure, it is
P.K. Robertson. Department of Civil Engineering, University the change in the excess pore pressure (∆u) with time that is
of Alberta, Edmonton, AB T6G 2G7, Canada. required for the evaluation of ch, where ∆u is defined as
R.G. Campanella. Department of Civil Engineering,
University of British Columbia, Vancouver, BC V5T 2W4, [1] ∆u = ui – uo
Canada.
D.J. Woeller. ConeTec Investigations Ltd., 9113 Shaughnessy
Street, Vancouver, BC V6P 6R9, Canada. and where

Can. Geotech. J. 36: 369–381 (1999) © 1999 NRC Canada


370 Can. Geotech. J. Vol. 36, 1999

Fig. 1. Terminology used for indicating location of pore-pressure Fig. 2. Typical dissipation records for normally consolidated
measurement. fine-grained soils.

Penetrometer

PORE PRESSURE, u
shaft
D

u3
Friction
sleeve
Behind tip
Shoulder u2 TIME, t

u1 tip
Cone face

NORMALIZED EXCESS PORE PRESSURE, U


ui is the measured pore pressure at the depth of interest;
and
uo is the equilibrium in situ pore pressure at the depth of
interest.
Interpretation of dissipation records is generally based on
a normalized excess pore-pressure ratio, U, defined as
∆ u(t) [u (t) − uo ]
[2] U= =
∆ ui (ui − uo)
where
∆u(t) is the excess pore pressure at any time t after pen- Log TIME

etration is stopped;
∆ui is the initial excess pore pressure at t = 0, i.e., on by dislocation methods has also been proposed by Elsworth
stopping penetration; and (1993). Comparisons of the available solutions and results
u(t) is the total pore pressure at any time t. from field studies suggest that the cavity-expansion method
Hence, for standard dissipation records where the excess of Torstensson (1977) and the strain-path approaches of
pore pressure shows a monotonic decrease with time, U var- Levadoux (1980) and Teh (1987) all provide similar predic-
ies between unity (at t = 0) and zero when 100% dissipation tions of consolidation parameters from CPTU dissipation
of the excess pore pressure has occurred. data (Gillespie 1981; Kabir and Lutenegger 1990; Robertson
With the development of piezocone equipment, it is now et al. 1991). Robertson et al. (1991) have shown that these
possible to measure penetration pore pressures at one or methods, although developed for normally consolidated
more locations on the cone. Three specific locations will be soils, can be equally applied to overconsolidated soils. Fur-
discussed in this paper and designated according to the thermore, comparisons of field and laboratory data indicate
scheme illustrated in Fig. 1. Hence, the excess pore pressure that the trends in the measured (laboratory) and predicted
can also be subscripted according to where the measure- (CPTU) data are consistent provided the microfabric and na-
ments are obtained (Sully et al. 1988): ture of the soils being tested are taken into consideration
[3] ∆u1,2,3 = u1,2,3 – uo (Danziger 1990; Robertson et al. 1991). Limited published
data are available to verify the more recent dislocation
Because of the small distance between each of the three method.
measurement locations (0.1 m max. between u1 and u3), the However, the relevance of any of the above solutions de-
magnitude of u0 can be taken to be equal for each of the pends on many factors, the most important of which relates
three positions. to how well the initial pore-pressure distribution around the
Once the required dissipation data have been obtained cone compares with the theoretical idealization employed by
during the penetration testing, the excess pore-pressure vari- each of the models. The initial distribution around the probe
ation with time can be plotted if the equilibrium pore pres- may be such that the applicability of these methods may be
sure is known (or measured). In low-permeability soils, it is questioned or restricted only to normally consolidated soils.
usual practice to continue taking dissipation measurements A typical set of excess pore-pressure dissipations in soft,
until at least half the initial excess pore pressure has dissi- normally consolidated clay for the three above-mentioned
pated (U = 0.5). filter locations is presented in Fig. 2a. The corresponding
Interpretation of the dissipation results can be achieved normalized dissipation curves are shown in Fig. 2b. All
using either of the two main analytical approaches: cavity- three pore-pressure dissipation curves show a monotonic de-
expansion theory or the strain-path approach. Interpretation crease in the excess pore pressure with time and essentially

© 1999 NRC Canada


Notes 371

Fig. 3. Pore-pressure dissipation in stiff, moderately pressure changes occurs, and (or) (ii) redistribution of pore
overconsolidated silty clay (Strong Pit). pressure around the tip due to the large gradients that are
2000 generated in overconsolidated soils. (The Mandel–Cryer ef-
u1
STRONG PIT
fect is not considered to be of major importance for loca-
Depth : 6.65 m
OCR = 4
tions behind the tip for the overconsolidated soils examined
Penetration Pore Pressure (kPa)

here.)
A further possibility for the increase in pore pressure has
been suggested by Coop and Wroth (1989) as a result of the
u2
maximum penetration pore pressure being located at some
point away from the shaft of the piezocone. These points
1000
will be considered briefly below.
u3
Poor saturation and (or) poor response of the
measurement system
(u1) t=0 =2473 kPa
If the rise in pore pressure on halting penetration were due
(u2) t=0 =1111 kPa solely to saturation and (or) measurement problems, then the
(u3) t=0 =611 kPa
effects should be equally frequent in normally consolidated
0 and overconsolidated soils. This is not the case. Further-
1 10 100 1000 10000
more, data of this type have been reported by some of the
Log Time (s) main research centres around the world where over-
consolidated soils have been studied and experimental tech-
niques are well proven. It would thus appear that the
agree with the theoretical models of the dissipation curve. anomalous pore-pressure rise is not due solely to poor field
Under these conditions, the data can be interpreted accord- technique. However, it must also be borne in mind that in
ing to any of the available theories to estimate the in situ heavily overconsolidated soils pore pressures behind the
consolidation parameter, ch, which primarily governs the cone tip may become negative and in some instances give
rate of dissipation for CPTU tests (Baligh and Levadoux rise to cavitation of the measuring system (Powell et al.
1980). The rate of dissipation is highest on the face of the 1988). If cavitation occurs, the measurement system may be-
cone and reduces with distance behind the tip; these effects come desaturated and sluggish response will result, giving
are considered in the analysis by varying the time factor, T, rise to curves somewhat similar to those shown in Fig. 3.
according to location of the pore-pressure element and by
using the radius of the probe at the location of the pore- Redistribution of pore pressure
pressure measuring sensor in the numerical calculations. Due to the large gradient of pore pressures around the tip
Under certain circumstances, the pore pressures measured in overconsolidated and stiff, fine-grained soils (Robertson
behind the tip do not decrease immediately on stopping pen- et al. 1986; Davidson 1985), drainage from the tip (high
etration; rather, they show an initial increase over a definite pore pressure) to the zone behind the tip (lower pore pres-
period of time before finally beginning to dissipate. Where sure) occurs, the rate of which is determined primarily by
the pore-pressure measurement system is completely satu- the soil permeability and the magnitude of the gradient. The
rated, dissipation records of this type are characteristic to fil- effect of the flow around the tip on the pore pressures mea-
ter locations located behind the cone tip (u2 and u3) for sured on the shaft also varies according to the soil stiffness
penetration in overconsolidated soils. The interpretation of and strength, parameters which also determine the pore-
these dissipation records to obtain predictions of ch is the pressure gradient itself. The authors consider this to be the
subject of this paper. principal reason why pore pressures measured at locations
behind the tip in overconsolidated and stiff, fine-grained
Pore-pressure dissipation in soils show an initial increase when penetration is stopped
followed by dissipation of excess pore pressures. It is also
overconsolidated soil
interesting to note that a comparison of the u2 and u3 pore
A typical example of pore-pressure dissipation in a lightly pressures in Fig. 3 indicates that the u2 value reaches a peak
overconsolidated fine grained soil (overconsolidation ratio faster than the u3 measurement. This is logical if the driving
OCR = 4) is illustrated in Fig. 3 using results obtained at force for the initial increase is the pore-pressure gradient,
Strong Pit in the Lower Mainland of British Columbia since the gradient between the u1 and u2 locations is much
(Campanella et al. 1988). Similar types of dissipation record higher than that between the u2 and u3 locations (Robertson
in overconsolidated soils for locations behind the tip have et al. 1986; Sully et al. 1988; Whittle et al. 1991).
been reported by Tumay et al. (1981), Davidson (1985),
BRE/NGI (1985), Kabir and Lutenegger (1987), Gillespie et Maximum pore-pressure located away from the shaft
al. (1988), Coop and Wroth (1989), Lunne et al. (1986), Coop and Wroth (1989) have suggested that the maximum
Gomez and Escalante (1987), and Bond and Jardine (1991). penetration pore pressure in overconsolidated soils is located
The initial rise in pore pressures measured at locations be- at some distance away from the shaft of the penetrometer. If
hind the tip in overconsolidated soils may be explained by this were the case, soil strength and stiffness would control
(i) poor saturation and (or) poor response of the measure- not only the magnitude of the pore pressure behind the tip
ment system such that a time lag in response to pore- but also the distance of the maximum value from the shaft.

© 1999 NRC Canada


372 Can. Geotech. J. Vol. 36, 1999

Fig. 4. Normalized pore-pressure dissipation for Strong Pit clay. Fig. 4. None of the curves follow the theoretical dissipation
1.25 trends suggested by the available theories for normally con-
TYPE II solidated soils and hence cannot be evaluated, as is, to pro-
u3 vide information on the coefficient of consolidation of the
ui=1111 kPa
soil. The departure of the field curves from the theoretical
1.00
framework is considered to be a result of the mean normal
Normalized Excess Pore Pressure, U

ui=611 kPa
stress reduction that occurs as the soil passes around the
cone tip and along the shaft. None of the available theoreti-
ui=2473 kPa u2
0.75 cal approaches adequately considers the magnitude of the
stress reduction in stiff overconsolidated soils and the im-
STRONG PIT
Depth : 6.65 m portant effect it has on the initial pore-pressure distribution
OCR = 4
0.50
around a penetrating cone.

u1
u1 (face at mid-height)
Characteristic dissipation types
0.25 u2 (behind at mid-height) In normally consolidated soils, the pore-pressure dissipa-
u3 (behind tip)
tion curve for any filter location (Fig. 2) can be considered
as a type I response. A type I response implies a monotonic
decrease of the initial penetration excess pore pressure. In
0.00
overconsolidated soils, several different responses may be
1 10 100 1000 10000
obtained depending on soil characteristics and filter location.
Time (s) For the filter located on the cone tip (u1), the unloading-type
dissipation associated with overconsolidated soils can be
classified as a type II response (Fig. 5a). A type II response
Hence, in normally consolidated soils the maximum pore is similar to a type I response once the pore-pressure reduc-
pressure would be close to or on the shaft and pore-pressure tion due to unloading has occurred.
decrease would occur on stopping penetration. As the soil A type III response is assigned to the locations behind the
becomes more overconsolidated, the location of the maxi- tip (u2, u3) where the excess pore pressure is greater than hy-
mum pore pressure would move away from the shaft and drostatic (u(t) > uo), but increases on stopping penetration
progressively longer time delays would occur before the before dissipating (Fig. 5b). All of the above responses have
pore pressure reached its peak value after stopping penetra- been discussed earlier.
tion. This idea would imply, however, that at all locations on
the shaft, irrespective of the distance behind the tip, the In moderately to heavily overconsolidated soils, the pore
measured pore pressures would all attain peak values at the pressures measured at the location immediately behind the
same time. This is not the case, as shown in Fig. 3, and in tip may be less than hydrostatic, or even below zero. In this
other overconsolidated dissipation data referenced above. case, on halting penetration the pore pressure increases to fi-
General trends in published data also do not confirm the hy- nally arrive at the in situ equilibrium value. Two types of
pothesis of Coop and Wroth (1989). dissipation curve may result depending on the soil character-
istics. (1) The measured pore pressure may increase over
and above the in situ equilibrium value if the rate of pore-
Typical dissipation records in pressure redistribution is higher than the rate of dissipation.
overconsolidated soils After reaching some peak value, the pore pressure then de-
Considering the data presented in Fig. 3 it is apparent that creases until the equilibrium value is reached (Fig. 5c, type
only the u1 pore pressure shows initial dissipation on halting IV). The type IV curve is similar to the type III response, the
penetration. However, the unloading of the tip resistance difference being the degree of pore-pressure change or un-
causes a sudden decrease in the u1 pore pressure. This sud- loading that occurs. (2) If the rate of dissipation is faster
den decrease modifies the dissipation record such that nor- than the rate of redistribution, the pore-pressure dissipation
malization with the initial maximum penetration pore does not overshoot but directly arrives at the equilibrium
pressure (ui)1 of 2473 kPa gives rise to a nonstandard dissi- value (Fig. 5c, type V).
pation record compared with that suggested by the available The standard approach for interpreting dissipation records
theories. For the locations behind the tip, the pore pressure in normally consolidated soils cannot be applied to the re-
initially increases when penetration is stopped, before finally sponses in overconsolidated soils (with the exception of an
decreasing and arriving at the in situ equilibrium value. inverted type V), since dissipation does not follow the theo-
In conclusion, it appears that interpretation of pore-pressure retical response, that is, a monotonic reduction with time.
dissipation records in overconsolidated soils is complicated The theoretical framework for evaluating dissipation in soils
by unloading effects and redistribution at all three filter lo- showing responses similar to those defined by types II–IV is
cations considered here. not available at present. However, the anomalous curves can
Commonly, the dissipation results are presented in terms be corrected to permit interpretation using the available the-
of normalized curves, whereby the normalized pore pres- ories. In fact, Elsworth (1993) suggests curve correction
sure, U, at any time t, is given by eq. [2]. The normalized prior to applying the dislocation method for curves where
dissipation curves for the records in Fig. 3 are shown in pore pressure increases on stopping penetration.

© 1999 NRC Canada


Notes 373

Fig. 5. Idealized pore-pressure dissipation response in over- Fig. 6. Normalized dissipation – logarithm of time plot for data
consolidated fine-grained soil for different measurement locations. in Fig. 3 corrected for unloading (u1) and redistribution (u2 and
u3). The dissipation data have been normalized to the maximum
value in the file, and time is taken equal to zero at this
maximum value.
ui
1.25
(a)
Pore Pressure, u

Type II

A 1.00
uc u3

Normalized Excess Pore Pressure, U


tc 0.75
u0 STRONG PIT u2
Depth : 6.65 m
OCR = 4
u1
0.50
Time, t

u1 (face)
B
uc (b) 0.25 u2 (behind friction sleeve)
u3 (behind tip)
Type III
Pore Pressure, u

tc
0.00
ui 1 10 100 1000 10000

Time (s)

u0

Correction to dissipation curves for redistribution


effects
Two methods of data presentation can be utilized to cor-
rect the type II–IV dissipation curves so that the available
Time, t
dissipation theories can be used to estimate values of the co-
efficient of consolidation. One approach is based on a loga-
rithm of time plot and the other is based on a square-root of
B Type IV (c)
time representation, both similar to the routine methods
uc presently employed for laboratory consolidation data. Either
Pore Pressure, u

approach can be used separately or combined to provide a


tc check on the results obtained. The application of the meth-
ods is presented using the results in Fig. 4 as an example.
u0
The step-by-step correction procedure for both methods is
Type V outlined in the Appendix.
ui ui < u0 ∆ u = Negative Logarithm of time plot correction
The data in Fig. 4 have to be corrected according to the
location of the pore-pressure measurement, i.e., either on the
cone tip or behind the cone tip, since the unloading and (or)
Time, t
redistribution that occurs affects the three sets of pore pres-
sures in different ways.
On the tip, a sudden decrease in pore pressure occurs on
The type V response can be considered as an inverted dis- halting penetration. In Fig. 4 the normalized pore pressure
sipation and treated in the standard way (as dissipation of a 5 s after dissipation begins is already reduced by 25% due to
negative excess). This type of response may occur in soils the reduction in the bearing stress acting on the face of the
with an OCR of 4 or larger, although the type of response cone. For this location, the initial maximum pore pressure
also depends on soil type and structure. For example, in used for normalizing the dissipation record is taken as the
heavily overconsolidated Taranto Clay (OCR = 20–40) the peak value once the initial unload has occurred (uc), i.e., for
pore pressures behind the tip are positive, possibly due to this case the maximum value corresponds to the 5 s mea-
the cementation present (Battaglio et al. 1986), whereas in surement and this time (tc) is taken as the new zero time
fissured London Clay (OCR = 25–50) negative pore pres- point (5 s are subtracted from the time register throughout
sures are recorded (Powell and Uglow 1988). the record). The maximum pore pressure for the dissipation

© 1999 NRC Canada


374 Can. Geotech. J. Vol. 36, 1999

Fig. 7. Details of the square root of time method for evaluating Fig. 8. Measured dissipation data for stiff clay in square root of
pore-pressure dissipation data in overconsolidated soils according time plot.
to response type: (a) types III to IV, and (b) type II. 2500

(a) STRONG PIT


u Depth : 6.65 m
OCR = 4
2000
u1
uc (From Back-Extrapolation)

Pore Pressure (kPa)


Pore Pressure, u

1500
Straight Line Approximation to Initial u2
Post-Peak Dissipation
Measured Pore Pressure 1000
Pore Pressure Dissipation
u3
Rise due to
Dissipation
(Measured) ~ t5 0 500
ui
u0

0
Root Time, t
0 15 30 45 60 75
0. 5
Root Time (s)
(b)
u ui After Stopping Penetration
Fig. 9. Square root of time extrapolation for short-duration
uc From Back-Extrapolation dissipation records.
of Linear Segment
Pore Pressure, u

Pore Pressure Decrease = 50% dissipation


Due to Unloading

Measured Pore Pressure Field dissipation stopped here


Dissipation
Linear extrapolation
to obtain t50
~ t50

u0

Root Time, t

record is taken as the peak value which occurs, for this par-
ticular record, at 5 s. caused by redistribution of pore pressure, initially depicts a
For the behind-tip locations, the maximum pore pressure straight line which can be back-extrapolated to t = 0 in order
is taken as the peak value that occurs during the post- to obtain a modified uc for the corrected dissipation curve.
penetration increase, and the time at which this peak occurs This value is then used to produce the normalized dissipa-
is taken as the new zero time of the dissipation record, with tion curve. The correction technique is illustrated in Fig. 7a
all other times adjusted accordingly. for measurement locations behind the tip. The principle is
The data from Fig. 4, corrected in this way, are replotted the same for measurement locations on the tip except that in-
in Fig. 6 to show the new form of the normalized dissipation stead of an increasing pore pressure, the initial pore pressure
plot, adjusted to account for unloading and redistribution ef- suddenly drops as discussed previously (Fig. 7b).
fects. The method of correcting the data is considered to be The data from Fig. 4 have been plotted in the square root
theoretically acceptable, since it adjusts the dissipation file of time base and are presented in Fig. 8. Direct estimates of
so that a monotonic reduction of a maximum pore pressure t50 can be read directly from Fig. 8.
to the in situ equilibrium occurs thereby removing the anom- The additional advantage of the square root of time
alous effects discussed above. method is that the initial straight-line portion can be extrapo-
lated to 50% pore-pressure reduction if short dissipation pe-
Square root of time plot riods are used in the field and measured data to longer
It is also possible to adjust the dissipation data of Fig. 4 periods are not available (Fig. 9), or the initial linear slope
using a back-extrapolation technique on a square-root of in the normalized pore pressure – square root of time plot
time plot, similar to the Taylor method used for interpreting can be analyzed to provide estimates of ch using the theoreti-
t50 values (50% reduction of excess pore pressure) from lab- cal approach suggested by Teh (1987).
oratory one-dimensional incremental consolidation testing. The two correction methods described above will give rise
In the square root of time plot, the dissipation after the peak to slightly different normalized dissipation curves, since the

© 1999 NRC Canada


Notes 375

Table 1. Geotechnical characteristics for the University of British Columbia test sites considered.

Site Depth range (m) PI (%) OCR Range of σvo′ (kPa) Range of σvm′ (kPa)
Lower 232 Street 1–5 21–30 3–10 16–40 90–205
Strong Pit 1–9 11–20 2–15 16–180 350–500
200th Street 1–5 20 2–17 16–51 115–300
Note: PI, Plasticity Index; σvo′, vertical effective stress presently acting in the ground on the soil; σvm′, maximum past vertical effective pressure from
incremental oedometer tests.

Table 2. Comparison of field and laboratory coefficients of consolidation.


cv from oedometer (cm2/s)
2
Site Filter location (ch)OC from CPTU (cm /s) Overconsolidated Normally consolidated ch/cv (oedometer)
Lower 232 Street u1 0.002–0.005 0.006–0.1 0.0005–0.001 2–3
u2 0.005–0.016
Strong Pit u1 0.004–0.007 0.002–0.005 0.0006–0.001 1–2
u2 0.006–0.01
200th Street u1 0.014–0.047 0.05–0.18 0.001–0.03 1.5–2
u2 0.045–0.054

Fig. 10. Comparison of normalized dissipation curves after Evaluation of proposed procedures in
applying proposed logarithm of time (log time) and square root overconsolidated soils
of time (root time) corrections.
1.25 Basis of comparison
The two correction procedures have been applied to CPTU
STRONG PIT dissipation data from three University of British Columbia
Depth : 6.65 m
OCR = 4 research sites where overconsolidated soils are present in the
1.00
profile. These extrapolation techniques have also been veri-
Normalized Excess Pore Pressure, U

fied using data presented by some of the major piezocone re-


search centres worldwide (Robertson et al. 1991). The
results of comparisons between the log t correction and
0.75
laboratory-derived consolidation parameters are considered
below. The square root of time method is not presented here
Corrected maximum because the difference between the resulting times from the
pore pressures (kPa)
0.50 Location logt root t
two methods is only significant for short dissipation periods.
(u1)i 1828 1922 All the results presented here are for 50% dissipation of the
(u2)i 1250 1328
768 844
excess pore pressure and, as suggested by Fig. 10, the differ-
(u3)i
ence between the two values of t50 is insignificant when
0.25 taken in context of the overall magnitudes involved.
u3 with log time correction
u2 with log time correction The horizontal coefficient of consolidation can be evalu-
u2 with root time correction ated from the corrected CPTU data using any of the avail-
u3 with root time correction
able theories. For this study, the method proposed by Teh
0.00 (1987) has been used. The advantage of this method is that it
1 10 100 1000 10000 considers the effect of the rigidity index on the pore-
Time (s) pressure dissipation. The ch value is determined from
T *R2 I R0.5
[4] ch =
initial corrected uc values are, by definition, not the same. t 50
The resulting corrected dissipation data for the Strong Pit
site are compared in Fig. 10 (the square root of time plot has where
been reproduced in logarithm of time space for comparison T* is the Teh and Houlsby (1988) modified time factor;
purposes). While the dissipation curves for U less than 25% R is the cone radius at the measurement location; and
for both corrections may be different (as would be expected IR is the rigidity index of the soil.
from the different ui values) at U = 50%, the error between Results presented by Danziger (1990) suggest that the Teh
the predicted values is relatively small (5–10%). In essence, and Houlsby (1988) approach provides more consistent ch
the two correction techniques give similar values for t50. The estimates than the other available methods.
curves in Fig. 10 do, however, indicate the importance of the For comparison with the CPTU interpretation, consolida-
initial pore-pressure value at t = 0 on the normalized form of tion coefficients from incremental laboratory oedometer tests
the dissipation curve. are presented. While these laboratory-determined values of

© 1999 NRC Canada


376 Can. Geotech. J. Vol. 36, 1999

Fig. 11. Published dissipation data for overconsolidated fine-grained soils, illustrating the proposed classification scheme. (a) Type II
response in Haga clay. (b) Type III response in glacial till at Cowden.
1000

900
(a) Type II
Haga
800
Decrease Due Moderately Overconsolidated
Marine Clay
Pore Pressure, u (kPa)

to Unloading
700 u1 Pore Pressure
Depth: 5.2m
OCR = 6
600 BRE/NGI (1985)

500

400

300

200

100

0
0.01 0.1 1.0 10 100

Time (min)

1000

900
(b) Type III
Cowden
800 Glacial Till
u2 Pore Pressure
Depth: 17.2m
Pore Pressure, u (kPa)

700
OCR > 3
BRE/NGI (1985)
600

500

400

300

200

100

0
0.01 0.1 1.0 10 100 1000 10000

Time (min)

the coefficient of consolidation cv may not be wholly repre- dated data are available. Furthermore, one-dimensional oedo-
sentative of in situ conditions, the results do provide a basis meter tests have also been performed on directionally-cut
on which the relative CPTU magnitudes can be compared. samples to obtain estimates of ch. Hence, the ch/cv ratio can
Field data are available from several international research be estimated from the laboratory tests to correct the CPTU
sites where CPTU-derived values of ch and the vertical coef- ch values to cv for direct comparison with the laboratory data.
ficient of consolidation cv can be compared with the results
of large-scale field tests (Robertson et al. 1991). Geotechnical review of University of British Columbia
Laboratory cv values are determined from each loading sites considered
stage so that both overconsolidated and normally consoli- The general geotechnical characteristics of the University

© 1999 NRC Canada


Notes 377

Fig. 11 (continued). (c) Type III response in London Clay at Brent Cross. (d) Type III response in overconsolidated silty clay at
Richards Island.
1000

900 (c) Type III


Brent Cross
Unweathered Fissured Clay
800

Pore Pressure (kPa)


u2 Pore Pressure
Depth: 13.5m
700 OCR > 30
BRE/NGI (1985)
600

500

400

300

200

100 u0 = 120 kPa

0
0 0.1 1 10 100 1000

Time After Penetration Stopped (min)

10
(d) Type III
9 Richards Island
OC Silty Clay
8 u3 Pore Pressure
Depth: 3.87m
OCR = 8
Pore Pressure Ratio (u/u0)

7 Campanella et al (1986)

2
U0
1

0
0 40 80 120 160 200 240

Time (s)

of British Columbia test sites under consideration are pre- sured at the u3 location are prohibitively long and the use of
sented in Table 1. The soils at Lower 232 Street are soft, piezocone dissipation tests at this location is not considered
sensitive clay silts. At the other two sites the clay silts are by the authors to be of practical interest unless graphical ex-
nonsensitive, with undrained strengths up to 200 kPa. trapolation techniques can be employed. This is evident
from the data in Fig. 6 which highlight the problems associ-
Results of comparison for University of British ated with the execution and interpretation of dissipation data
Columbia sites at locations behind the tip, since the time periods required
Only data from pore-pressure locations u1 and u2 have are very long in low-permeability soils. The time for 50%
been used for the purposes of comparison. The times re- dissipation at the u1 location is 1100 s (18 min), which in-
quired for 50% dissipation of the excess pore 2 pressure mea- creases to 2300 s for u2 and approximately 10 000 s for

© 1999 NRC Canada


378 Can. Geotech. J. Vol. 36, 1999

Fig. 11 (concluded). (e) Type IV response at 200th Street. (f ) Type V response in stiff sandy silty clay.

Excess Pore Pressure (kPa)


(e) Type IV
200th Street
100 OC Silty Clay
u2 & u3 Pore Pressure
Depth: 2.95m U2
u0
OCR 10 U3
0 Unpublished UBC Data

-100
0 10 20 30 40
Time (s)

300
u0 280 kPa

250
u (kPa)

(f) Type V
200 Bahia El Tablazo
u2 Location
Sandy Silty Clay
150 OCR 4
Gomez and Escalante (1987)
100
0 50 100 150 200
Time (s)

u3 (the u3 record has been extrapolated for comparative pur- carried out in a variety of overconsolidated soils worldwide.
poses to obtain the 50% dissipation times). The data have been summarized by Robertson et al. (1991).
The coefficients of consolidation from CPTU (ch) and lab- Typical results highlighting the different types of pore-
oratory oedometer (cv) tests have been determined as de- pressure response described above are presented below.
scribed above. The obtained values are compared in Table 2. The results of a u1 dissipation in moderately overcon-
For the Lower 232 Street data an average Ir value of 50 has solidated Haga clay (BRE/NGI 1985) are presented in
been used to evaluate ch, whereas values of 100 and 200 Fig. 11a. The pore pressure of about 875 kPa generated dur-
have been used for 200th Street and Strong Pit, respectively ing penetration decreases immediately to 480 kPa after stop-
(the Ir = G/Su values have been obtained from in situ shear ping penetration. Thereafter, the dissipation occurs as the
wave velocity measurements and field vane shear tests, excess pore pressure close to the cone dissipates laterally
where G is the maximum shear modulus, and Su is the un- into the surrounding soil. This is similar to the type II re-
drained shear strength). sponse described earlier. The lightly overconsolidated gla-
The ch values from the two pore-pressure measurement cial till at Cowden (OCR = 3) provides an example of a type
locations (u1 and u2) are in very good agreement. Further- III dissipation at the u2 location (BRE/NGI 1985). The posi-
more, the CPTU values would suggest that the t50 dissipa- tive penetration pore pressure of 835 kPa increases after
tion provides an estimate of the ch corresponding to the stopping penetration until a maximum value of around
overconsolidated condition, i.e., the in situ condition of the 980 kPa is reached some 10 s later (Fig. 11b). Correcting the
soil. Considering the results presented by Baligh and dissipation curves for the maximum values and using the
Levadoux (1980) who show that, in normally consolidated new zero times, as described earlier, allow the modified dis-
soils, the t50 time corresponds to the (ch)NC, it would appear sipation curve to be interpreted to obtain the over-
that at degrees of dissipation of 50% the theory provides es- consolidated coefficient of consolidation. The data from
timates of the coefficient of consolidation relevant to the in Brent Cross (OCR > 30) are also characteristic of a type III
situ stress history condition, (ch)OC. In addition, it is appar- response for the u2 location (Fig. 11c).
ent from the types of dissipation curve considered that stress The Richards Island silty clay (OCR = 8) dissipation
history may be an important factor to be considered when results presented by Campanella et al. (1986) are typical of
interpreting CPTU dissipation data. As suggested by Kabir a type III response for the u3 location (Fig. 11d). For
and Lutenegger (1987), the dissipation curve may be useful this case the penetration pore pressure is only marginally
as a stress-history indicator. higher than the in situ equilibrium value, uo ; if it had been
lower than uo, then this would have been a type IV dissipa-
Evaluation of data from international sites tion response, similar to the result obtained at 200th Street in
Dissipation tests using the cone penetrometer have been the Lower Mainland of British Columbia (Fig. 11e) where
© 1999 NRC Canada
Notes 379

Fig. 12. Comparison of t50 from corrected dissipation data in overconsolidated soils with ch from laboratory consolidation tests using
dissipation interpretation nomograph presented by Robertson et al. (1991).

the penetration pore pressure at the start of the dissipation ing the ch–t50 nomograph. In Fig. 12 the CPTU dissipation
(u i = –90 kPa) is less than the equilibrium value (uo = data are compared with ch values obtained from laboratory
16 kPa) in this moderately overconsolidated clay (OCR = consolidation tests. Data have been reported in the literature
10). As the permeability of the soil increases, the type IV re- for a limited number of sites, where ch has also been deter-
sponse changes to a type V response, where the negative ex- mined from back-analysis of field performance. The pub-
cess generated during penetration dissipates to the lished values of ch obtained from field performance are
equilibrium value. This type of response has been recorded generally larger than laboratory values (Robertson et al.
in dilative silts and sandy silty clays (Fig. 11f ). 1991; Jones and Rust 1993) and provide better agreement
All the dissipation data for overconsolidated soils dis- with the theoretical range obtained from the Teh and
cussed above were presented by Robertson et al. (1991), Houlsby (1988) solution.
having been interpreted using the above correction tech-
Conclusions
niques. These data and other results for overconsolidated
soils presented by Robertson et al. are plotted in Fig. 12, us- This paper has presented a classification for the types of
© 1999 NRC Canada
380 Can. Geotech. J. Vol. 36, 1999

dissipation response recorded in overconsolidated, fine- Campanella, R.G., and Robertson, P.K. 1988. Current status of the
grained soils during piezocone penetration tests. The differ- piezocone test. In Proceedings of the 1st International Sympo-
ent dissipation responses have been explained in terms of sium on Penetration Testing, Orlando (ISOPT-1). A.A. Balkema,
the unloading and redistribution that occur as the soil under- Rotterdam The Netherlands, Vol. 1, pp. 93–116.
goes lateral displacement to permit advancement of the cone Campanella, R.G., Robertson, P.K., and Gillespie, D. 1986. Factors
tip and shaft. The singularity at the shoulder of the cone tip affecting the pore water pressure and its measurement around a
gives rise to the unloading, as the soil no longer undergoes penetrating cone. In Proceedings of the 39th Canadian
cavity-expansion type deformation, but rather shearing along Geotechnical Conference, Ottawa. pp. 291–299.
Campanella, R.G., Sully, J.P., and Robertson, P.K. 1988. Interpreta-
the shaft–soil interface. The soil stiffness and strength,
tion of piezocone soundings in clay — a case history. In Pro-
which can be considered to be dependent on the consolida-
ceedings, Penetration Testing in the UK 6–8 July 1988.
tion state of the soil, control the degree of unloading and re- Institution of Civil Engineers, London. pp. 203–208.
distribution, and consequently the extent of the modified Coop, M.R., and Wroth, C.P. 1989. Field studies of an instru-
pore-pressure response (compared to the type I response as- mented model pile in clay. Géotechnique, 39(4): 679–696.
sociated with normally consolidated soils). Also important in Danziger, F.A.B. 1990. Desenvolvamento de equipamento para
controlling the type of response is the location of the porous realizao de ensaio de piezocone: aplicao a argilas moles. Ph.D.
element used for measuring the pore-pressure variation with thesis, COPPE, Universidade Federal de Rio de Janeiro, Rio de
time. Janeiro, Brazil.
Theoretical models do not exist at present that adequately Davidson, J.L. 1985. Pore pressures generated during cone pene-
model the unloading and redistribution of pore pressure which tration testing in heavily overconsolidated clays. In Discussion
occur during penetration of a cone into stiff, overconsoli- to Session 2D: Field Instrumentation and Field Measurements,
dated, fine-grained soils. As a result, the above effects can- 11th International Conference on Soil Mechanics and Founda-
not be verified by numerical techniques and semi-empirical tion Engineering, San Francisco. Vol. 5, p. 2699.
interpretation is required. Similar empirical and semi-empirical Elsworth, D. 1993. Analysis of piezocone dissipation data using
approaches are used to interpret CPT data in granular soils dislocation methods. Journal of Geotechnical Engineering,
where stress rather than pore-pressure gradients occurs. How- ASCE, 119(10): 1601–1623.
ever, the consistent dissipation pore pressure response types Gillespie, D.G. 1981. The piezometer cone penetration test.
reported by major research centres worldwide indicate the M.A.Sc. thesis, Department of Civil Engineering, University of
validity of the reported behavior and the approach to inter- British Columbia, Vancouver, B.C.
pretation. Gillespie, D., Robertson, P.K., and Campanella, R.G. 1988. Con-
Two data-presentation methods have been discussed for solidation after undrained Piezocone penetration. I: Prediction:
correcting the different types of pore-pressure response in Discussion. Journal of Geotechnical Engineering, ASCE,
114(1): 126–128.
overconsolidated soils. In one method, the data in log t space
Gomez, N.R., and Escalante, B.N. 1987. Evaluacion de suelos de
are normalized to the maximum post-penetration value and
baja permeabilidad mediante el piezocono. Trabajo Especial de
the time at which this maximum value occurs. In the second Grado, Universidad Central de Venezuela, Caracas, Venezuela.
approach, a square root of time correction and extrapolation Jones, G., and Rust, E. 1993. Estimating coefficient of consolida-
are used to correct the field data (see the Appendix). Once tion from piezocone tests: Discussion. Canadian Geotechnical
either of the above modifications has been made, the data Journal, 30: 723–724.
can be interpreted using standard techniques available for Kabir, M.G., and Lutenegger, A.J. 1987. Consolidation after un-
normally consolidated soils. The Teh and Houlsby (1988) drained Piezocone penetration. I: Prediction: Discussion. Journal
method has been used in this paper to interpret the data us- of Geotechnical Engineering, ASCE, 114(1): 128–130.
ing the nomograph presented by Robertson et al. (1991). Kabir, M.G., and Lutenegger, A.J. 1990. In situ estimation of the
The ch values from CPTU generally fall into the range coefficient of consolidation of clays. Canadian Geotechnical
suggested by the laboratory data for the overconsolidated Journal, 27: 58–67.
state for the in situ tests at the wide range of sites consid- Levadoux, J.N. 1980. Pore pressures in clay due to cone penetra-
ered. The correction method proposed here appears to pro- tion. Ph.D. thesis, Massachusetts Institute of Technology, Cam-
vide consistent estimates of ch. bridge, Mass.
Lunne, T., Eidsmoen, T., Powell, J.J.M., and Quarterman, R.S.T.
1986. Piezocone testing in overconsolidated clays. Norwegian
References
Geotechnical Institute, Report NGI 52155-42.
Baligh, M.M., and Levadoux, J.L. 1980. Pore pressure dissipation Powell, J.J.M., and Uglow, I.M. 1988. Marchetti dilatometer test-
after cone penetration. Report No. MITSG 80-13, Massachusetts ing in U.K. soils. In Proceedings of the 1st International Sympo-
Institute of Technology, Cambridge, Mass. sium on Penetration Testing, Orlando (ISOPT-1). A.A Balkema,
Battaglio, M., Bruzzi, D., Jamiolkowski, M., and Lancellotta, R. Rotterdam, The Netherlands. Vol. 2, pp. 555–562.
1986. Interpretation of CPT’s and CPTU’s — undrained pene- Powell, J.J.M., Quarterman, R.S.T., and Lunne, T. 1988. Interpreta-
tration of saturated clays. In Proceedings of the 4th International tion and use of the piezocone test in U.K. clays. In Proceedings,
Geotechnical Seminar, Singapore 25–29 November. pp. 1–28. Penetration Testing in the UK. Institution of Civil Engineers,
Bond, A.J., and Jardine, R.J. 1991. Effects of installing displace- London. pp. 151–156.
ment piles in a high OCR clay. Géotechnique, 41(3): 341–363. Robertson, P.K., Campanella, R.G., Gillespie, D.G., and Greig, J.
BRE/NGI. 1985. Comparison of piezo cones in stiff 1986. Use of piezometer cone data. In Proceedings, In Situ ‘86,
overconsolidated clays. Building Research Establishment joint Blacksburg, Va. pp. 1263–1280.
report with the Norwegian Geotechnical Institute, BRE N(C) Robertson, P.K., Sully, J.P., Woeller, D.J., Lunne, T., Powell, J.J.,
55/85 and NGI 84223-1. and Gillespie, D.G. 1991. Estimating coefficient of consolida-

© 1999 NRC Canada


Notes 381

tion from piezocone tests. Canadian Geotechnical Journal, 29: structed Facilities Division, Department of Civil Engineering,
539–550. Massachusetts Institute of Technology, Cambridge, Mass.
Sully, J.P., Campanella, R.G., and Robertson, P.K. 1988.
Overconsolidation ratio of clays from penetration pore pres- Appendix: Classification used for the
sures. Journal of Geotechnical Engineering, ASCE, 114(2): 209–
216. overconsolidation state
Teh, C.I. 1987. An analytical study of the cone penetration test. The following classification is used here for description of
D.Phil. thesis, Oxford University, Oxford, U.K. the soil overconsolidation state, with the overconsolidation
Teh, C.I., and Houlsby, G.T. 1988. Analysis of the cone penetration ratio determined as OCR = σvm′/σvo′, where σvm′ is the maxi-
test by the strain path method. In Proceedings of the 6th Interna- mum vertical effective pressure the soil has experienced, and
tional Conference on Numerical Methods in Geomechanics,
σvo′ is the vertical effective stress presently acting in the
Innsbruck.
ground on the soil.
Torstensson, B.-A. 1977. The pore pressure probe. In Geotechnical
Meeting, Norwegian Geotechnical Society, Oslo, Norway. Paper
34, pp. 34.1–34.15.
Stress history Description
Tumay, M.T., Boggess, R.L., and Acar, Y. 1981. Subsurface inves-
tigations with piezo-cone penetrometer. In Proceedings of a Spe- OCR ≤ 1 Normally consolidated (NC)
cialty Conference, American Society of Civil Engineers, St. 1 < OCR ≤ 4 Lightly overconsolidated (LOC)
Louis. pp. 325–342. 4 < OCR ≤ 10 Moderately overconsolidated (MOC)
Whittle, A.J., Aubeny, C.P., Rafalovich, A., Ladd, C.C., and 10 < OCR ≤ 25 Heavily overconsolidated (HOC)
Baligh, M.M. 1991. Prediction and interpretation of in situ pen- 25 < OCR Very heavily overconsolidated (VHOC)
etration tests in cohesive soils. Research Report R91-01, Con-

© 1999 NRC Canada

View publication stats

You might also like