You are on page 1of 8

3rd International Symposium on Cone Penetration Testing, Las Vegas, Nevada, USA - 2014

On The Interpretation of Piezocone Dissipation Testing Data


F. M. Mantaras
Geoforma Engenharia Ltda, Joinville, Brazil

E. Odebrecht
Geoforma Engenharia Ltda State University of Santa Catarina - UDESC, Joinville, Brazil

F. Schnaid
Federal University of Rio Grande do Sul UFRGS, Porto Alegre, Brazil

ABSTRACT: Interpretation of piezocone dissipation testing data to obtain the best global value of the
coefficient of consolidation requires the time for 50% dissipation of excess pore pressure to be
calculated. In engineering practice, it implies on extracting a single point of the field dissipation curve
(t50) or adjusting the complete curve. This papers gives strainghtfoward recommendations on how to
curve fitting the field dissipation curve and use the first and second derivates to estimate the value of
t50. Major advantages of the proposed approach are: (a) for spreadsheet use, the derivate of the
dissipation response gives the corresponding time t50 and (b) the equilibrium in situ pore pressure is no
longer required to calculate the percentage of dissipation. In addition, the moment of interrupting dissipation tests in the field becomes undisputable.

1 INTRODUCTION
Coefficients of consolidation can be assessed in situ from observations of settlements under embankments
or directly from in situ test results, preferably from piezocone dissipation tests. Analytical and numerical
procedures have been developed to provide an estimate of the coefficient of consolidation ch from piezocone dissipation tests in which the decay of excess pore pressure with time is monitored. Methods rely either on one-dimensional cavity expansion (Torstensson, 1977; Randolph & Wroth, 1979) or twodimensional strain path method (Baligh & Levadoux, 1986; Teh & Houlsby, 1991; Burns & Mayne,
1998). Both monotonic and dilatory soil response can be modeled in this type of approach, being the Teh
& Houlsby method recommended for monotonic pore pressure dissipation response and the Burns &
Mayne method recommended for dilatory response.
The procedure for assessing ch from dissipation tests is based on a series of straightforward recommendations (Lunne et al, 1997):
1) Plot the early part of the dissipation (less than 10% dissipation) at an enlarged scale, either log or
square root time, and evaluate the initial pore pressure ui;
2) Define u0 from avaiable data on ground water level, piezometric readings or even data from
piezocone tests in adjacent sand layers;
3) Plot normalized excess pore pressure (U = (ut u0)/(ui u0) against time on a log and/or t scale,
where ut is the pore pressure at time t and u0 is the in situ equilibrium);
4) Determine the time for 50% dissipation (t50).
315

Dissipation test results are usually ploted as the variation of excess pore pressure and the logarithm of
time. Alternatively results can be normalized as a time factor T* (=[ch t /(r2 Ir0.5)]) and pore pressure ratio
U*2 (= u2/u2i). In these plots, dissipation tests should allow for 50% of dissipation of excess pore
pressure to provide sufficient data for interpetation. If dissipation is not sufficiently long to define t50, then
the slope of the straight line from the first part of the u versus t plot may be used to predic ch.
Albeit this set of recommendations is apparently simple to apply, reliable measurements of the equilibrium in situ pore pressure, important in many geotechnical design problems and essential to the interpretation of ch, are not always straightforward to obtain and requires considerable engineering judgment in
many applications. In clays, the pore water pressure generated in a CPTU test can be allowed to dissipate
to the equilibrium value to assess the in situ pore pressure but this can take many hours and makes the use
of the standard CPTU unattractive for this purposes. In addition hydrogeology is often complex and
piezometer measurements have to be interpreted to estimate average equilibirum pressures. Coastal and
nearshore tests are affected by tidal variations. The determination of phreatic surface location for tailings
and tailing-retention structures is influenced by pond location, anisotropic permeabiity of deposits,
boundary flow conditions, including the presence of permeable layers and drains, among other factors
(e.g. Vick, 1983).
Given the uncertainties in defining u0, the time for 50% dissipation of the measured maximum value
(t50) may be in error (and so is the estimated value of ch). To partially overcome this problem, Mayne
(2001) proposed a mathematical approach that instead of matching just a single point (usually 50%) of the
recorded dissipation, the complete curve is adjusted to obtain the best global value of the horizontal coefficient of consolidation, ch. The present paper extends this view by deriving the equation used to curve fitting the measured pore pressure data and, once this is accomplished, the slope of the derivate at the minimum point is zero and corresponds to t50. The time t50 can then be used in the interpretation of the value of
ch adopting standard methods.
2 EMPIRICAL DETERMINATION OF

The aproach is based on a method to back fit the field pore pressure measurements to evaluate the time for
50% dissipation of the measured maximum value (t50). The proposed approach lacks of a physical basis
and there is no intention in determining the best-fit parameters that define a model. The curve fitting is
used to generate a standard curve that can be useful in interpolating a set of pore pressure values to define
a curve that is smooth and comes close to the measured data. Once a function is selected and the curve fitting is completed, the function's first derivative is obtained and the minimum point of the function where
the slope is zero is identified. In the second derivate this point is zero. These points (zero slope in the first
derivate and zero in the second) correspond to t50.
In the proposed procedure a polynomial equation is used to curve fitting the data, with the actual mathematical expression (degree) defined by the minimum r2.

2.1 Teh & Houlsby (1988)


From the theoretical point of view, applicability of the approach can be illustrated using the strain path
method proposed by Teh & Houlsby (1991). The dissipation curves are not unique since the initial pore
pressure distribution is dependent on the value of Ir. Unifying results requires empirical normalization of
test data by means of a dimensionless time factor T*, expressed as:
316

(1)

where r is the probe radius, t the dissipation time (normally adopted as t50%), Ir the rigidity index (=
G/Su) and G the shear modulus. A constant Ir is used in the solution although in fact the value of the shear
modulus will depend on the shear strain amplitude, which is shown by strain path calculations to vary in a
complex manner around a 60o penetrometer.
Predicted dissipation curves shown in Figure 1 can be conveniently represented using approximate numerical algorithms, thus offering a means of implementing data on a spread-sheet. Mayne (2001) proposed a logarithm equation for matching the data. In the present work a polynomial equation was chosen
due to the fact that it is easy to derive and provides means of fitting both monotonic and dilatory soil response.

Figure 1 Teh & Houlsby solution (1988).

The equation proposed by Mayne (2001) to fit the data is:


(0.85 + 10), 0.08

(2)

In order to get sufficient amount of data to represent the initial part of the curve in a logarithmic time
scale the variable T* is replaced by 10 (x-3):
,

0.85 + 10 10()

0.08

(3)

The derivative of the function with respect to the variable x can now be expressed as:

4.5

10() ln(10)
()
.
[0.85
]

+ 10

(4)

Alternatively a polynomial expression can be used to facilitate the mathematical treatment of the differentiation. Expression (5) could be accurately represented by an xth degree polynomial as:
317

= + + + + +

(5)

Table 1 lists a suitable set of coefficients that best fit the Teh & Houlsbys equation.
Table 1: Polynomial coefficients for the solution proposed by Teh & Houlsby (1998).
a0
0.26384

a1
-0.34277

a2
0.11916

a3
0.03866

a4
-0.03729

a5
0.00329

a6
0.00860

a7
0.00076

a8
-0.00103

a9
-0.00021

Once the coefficients are determined, the first and the second derivate can be easily obtained as:

= + 2 + 3 + 4 + + ()

= 2 + 6 + 12 + + ( 1) ()

(6)

(7)

Equation 6 represents the slope of pore pressure dissipation in log scale and can be differentiated to obtain the inflection point of the U x T(log) curve (see Figure 2).

Figure 2 1st and 2ndderivates of Teh and Houlsby solution.

In the analysis the program is instructed to graph the curve to check the compatibility of the measured data
with the superimposed curve, as well to plot the first and second derivates to calculate t50. The minimum
value of the polynomial first derivate corresponds to a time factor T equal to 0.245 (the theoretical value
calculated by Teh & Houlsby, 1998).

318

2.2 Burns & Mayne (1998)


The mathematical solution proposed by Burns & Mayne (1998) is based on the cavity expansion-critical
state for the monotonic and dilatory response with regard to time. The excess pore water pressures, ut, at
any time (t) can be compared with the initial values during penetration, ui=u2-u0 and are represented as:
= ()[1 + 50 ] + ()[1 + 5000 ]

(8)

2
where: () =
3 2 ln()is the octahedral component during penetration and

() =
1 2 is the shear-induced component during penetration, being the effective
friction angle, M=(6sin)/(3-sin), the compression index and T* the modified time factor T*=[ch t
/(r2 Ir0.5)].

Figure 3 shows the normalized dissipation curves estimated for values of ' equal to 25o, = 0.8, Ir = 50
and OCRs ranging from 1 to 100 (Burns &Mayne 1991).

Figure 3 Solutions for Type 2 dilatory dissipation curves for various OCRs (after Burns & Mayne, 1998).

Pore pressures are represented according:


= ()[1 + 50 ] + ()[1 + 5000 ]

(9)

Equation 9 can by divided by the initial effective vertical stress (vo) and knowing that
= (2 3) ln(I), the first derivative of Equation 9 is:


50 2
1+ 2

=
5000

(1 + 50 )
(1 + 5000 )

(10)

Conversion of decimal scale T to logarithmic scale is done by substitution: T = 10x. For mathematical
convenience, T* is expressed as 10x with x ranging from -4 to +1. Equation 9 can then be re-written as:

319



50 2
1+ 2

=
5000

(1 + 50 10)
(1 + 5000 10)

(11)

The second derivative is then expressed as:

2
2
(10) 50
= 5000
10 (10)

(1
)
(1
+ 50 10
+ 50 10)



1
1

2
2
) () 5000
(10
+ 50000000
10 (10)
(1 + 5000 10 )
(1 + 5000 10)

(12)

Applications of the proposed methodology for OCR = 3 and OCR= 10 are shown in Figures 4 and 5,
respectively. In both cases and first and second derivatives correspond to the point of inflection of the
normalized dissipation curve and define the theoretical value of t50.

Figure 4 1st and 2ndderivatives of Burns and Mayne solution for OCR = 3.

Figure 5 1st and 2ndderivatives of Burns and Mayne solution for OCR = 10.

320

This simple analysis indicates that a mathematical approach used to adjust the complete dissipation curve,
followed by the determination of t50 from the first and second derivatives is a consistent approach to obtain the best global value of the coefficient of consolidation, ch.
3 CASE STUDY
Piezocone tests carried out according to ASTM D5778 (2000) standards are used to illustrate the proposed
approach. Figure 6 (a) presents a typical CPTU log in a soft clay deposit which plots the following data:
tip resistance (corrected for unequal cone ends qt, pore pressures using a type 2 pore pressure measurement (u2) and hydrostatic equilibrium pressure (u0), as well as the calculated Bq and OCR variation with
depth. The dissipation test carried out at 11m depth shows an initial dilatant response that gradually reduces towards the hydrostatic pressure of 110 kPa (Figure 6 (b)). The maximum excess pore pressure of
330kPa yields u50 of 221 kPa which is exactly the value given by the 1st and 2ndderivatives.
The second example discusses results of a site investigation carried out in bauxite tailing deposit from
northern Brazil. The material is saturated and recently deposited, undergoing consolidation. The equilibrium pore pressure is non-linear as indicated in Figure 7(a). The dissipation test at 8m shows a monotonic
reduction in pore pressure from a maximum value of 195 kPa to an equilibrium value estimated as 45 kPa
leading to u50 of 118 kPa (Figure 7 (b)). The 1st and 2ndderivatives are shown to give the same result directly from a spreadsheet calculation.

(a)

(b)

Figure 6 (a) Piezocone data from a soft clay deposit (b) Dissipation test interpretation at 11m depth.

4 CONCLUSIONS
This paper describes a procedure conceived to determine the value of 50% dissipation of excess pore
pressure from piezocone dissipation tests. It consists in curve fitting the field dissipation curve and use the
first and second derivates to estimate the value of t50 without any consideration regarding the equilibrium
in situ pore pressure.

321

(a)

(b)

Figure 7 (a) Piezocone results from bauxite tailings (b) Dissipation test interpretation at 8m depth.

The accuracy of the proposed solution depends mainly on how well the theoretical idealization describes the pore pressure distribution around the cone. Experience tends to suggest that in normally to
lightly overconsolidated soils the background theory (cavity expansion or stain path) is able to reproduce
essential features of initial pore pressure distribution and subsequent dissipation. In these cases, the proposed approach can be used to derive the consolidation data of soft soils.
5 REFERENCE
Baligh, M.M. & Levadox, J.N. 1986. Consolidation after undrained piezocone penetration. II: Interpretation. J. Soil
Mech. Found. Engng. Div., ASCE. 11(7), 727-745.
Burns, S.E. & Mayne, P.W. 1998. Monotonic and dilatory pore-pressure decay during piezocone tests in clay. Can.
Geotech. J., 35(6): 1063-1073.
Mayne, P.W. (2001) Stress-Strain-Strength-Flow Parameters from Enhanced In-Situ Tests, Proceedings, International Conference on In-Situ Measurement of Soil Properties and Case Histories, Bali, Indonesia, 2001, pp. 27
48.
Baligh, M. M. & Levadoux, J.N. 1986. Consolidation after undraned piezocone penetration. II: Interpretation J. Soil
Mech. Found. Engng. Div., ASCE, 112(7): 727-745.
Lunne, T.; Robertson, P.K. & Powell, J.J.M. (1997). Cone penetration testing in geotechnical practice, Blackie Academic & Professional, 312p.
Randolph, M.F., & Wroth, C.P. 1979. An analytical solution for the consolidation around a driven pile. International Journal for Numerical and Analytical Methods in Geomechanics, 3: 217- 229.
Teh, C.I. & Houlsby, G.T. 1991. An analytical study of the cone penetration test in clay. Gotechnique, 41(1): 1734.
Teh, C.I. & Houlsby, G.T. 1988. An Analysis of the Cone Penetration Test by the Stain Path Method. Numerical
Methods in Geomechanics, Vol, 1, Proc. 6th NUMOG, Innsbruck) Balkema, Rotterdam, 397-402.
Torstensson, B.A. 1977. Time-dependent effects in the field vane test. Int. Symp. Soft Clay, Bangkok, 387-397.
Vick, S.G. 1983. Planning, Design and analysis of Tailings Dams. John Wiley & Sons, Inc., 369p.

322

You might also like