You are on page 1of 15

Computation of Cavity Expansion Pressure and Penetration

Resistance in Sands
R. Salgado, M.ASCE1; and M. Prezzi, A.M.ASCE2
Downloaded from ascelibrary.org by UFRGS - Universidade Federal do Rio Grande do Sul on 01/09/20. Copyright ASCE. For personal use only; all rights reserved.

Abstract: A cavity expansion-based theory for calculation of cone penetration resistance qc in sand is presented. The theory includes a
completely new analysis to obtain cone resistance from cavity limit pressure. In order to more clearly link the proposed theory with the
classical cavity expansion theories, which were based on linear elastic, perfectly plastic soil response, linear equivalent values of Young’s
modulus, Poisson’s ratio and friction and dilatancy angles are given in charts as a function of relative density, stress state, and critical-state
friction angle. These linear-equivalent values may be used in the classical theories to obtain very good estimates of cavity pressure. A
much simpler way to estimate qc—based on direct reading from charts in terms of relative density, stress state, and critical-state friction
angle—is also proposed. Finally, a single equation obtained by regression of qc on relative density and stress state for a range of values
of critical-state friction angle is also proposed. Examples illustrate the different ways of calculating cone resistance and interpreting cone
penetration test results.
DOI: 10.1061/共ASCE兲1532-3641共2007兲7:4共251兲
CE Database subject headings: Computation; Cone penetration tests; Sand; In situ tests; Penetration resistance.

Introduction The cone penetration resistance analysis has evolved in vari-


ous ways since it was first presented in Salgado et al. 共1997b兲.
The cone penetration test 共CPT兲 is now widely used for geotech- Most notably, limit pressure is calculated in a much more effec-
nical site characterization and in situ determination of soil prop- tive way, and a new formulation for calculating cone resistance
erties. Among its advantages are simplicity, speed, and continuous from cavity limit pressure 共which considers the true interface fric-
profiling. Originally, the penetrometer was used to measure only tion angle between the cone and soil兲 has been developed and
tip resistance, defined as the vertical force acting on the tip of the implemented. These developments are presented in this paper.
penetrometer divided by the base area of the tip 共10 cm2 for the Moreover, the cavity expansion analysis on which the cone resis-
standard 35.7-mm-diameter penetrometer兲. Over the years, sen- tance analysis is largely based, is linked to the classical cavity
sors have been incorporated into the cone to measure the friction expansion analyses, all based on linear elasticity and perfect plas-
along a lateral sleeve, the arrival of a seismic shear wave 共which ticity, by the presentation of values of friction and dilatancy
allows determining shear wave velocities兲, pore pressure, and angles, Young’s modulus and Poisson’s ratio that, once plugged
other quantities 共Mitchell 1988兲. This versatility has further en- into those analyses, would produce substantially the same values
hanced the use of the CPT. In addition, an important advantage of of cavity expansion limit pressure as the present theory.
the CPT is that the penetration process is amenable to theoretical In practice, engineers often use simplified equations and
modeling. charts. Two additional methods of penetration resistance estima-
A general penetration resistance theory allowing calculation of tion, one relying on charts of cone resistance qc expressed directly
penetration resistance in geologically recent, uncemented, clean in terms of density and stress, the other on simple expressions of
sands was developed by Salgado et al. 共1997b兲 and codified into qc in terms of relative density and stress state, are given. Ex-
the program CONPOINT. Predictions by this theory were exten- amples illustrate the application of these different methods to
sively and successfully tested against actual measurements of cone resistance interpretation.
penetration resistance made in calibration chambers 共Salgado
et al. 1997a,b; 1998兲 and in the field 共Lee et al. 1999; Bonita
2000兲. Cavity Expansion Analysis
1
Professor, School of Civil Engineering, Purdue Univ., 550 Stadium
Analytical Solutions
Mall Dr., West Lafayette, IN 47907-2051. E-mail: rodrigo@
ecn.purdue.edu Originally, cavity expansion analysis aimed to solve problems of
2
Assistant Professor, School of Civil Engineering, Purdue Univ., 550 metal indentation 共Bishop et al. 1945; Hill 1950兲, which became
Stadium Mall Dr., West Lafayette, IN 47907-2051. E-mail: mprezzi@ more important as the industrial revolution intensified in the last
ecn.purdue.edu part of the 19th Century and first half of the 20th Century. So, in
Note. Discussion open until January 1, 2008. Separate discussions
its origins, cavity expansion analysis was very much linked to a
must be submitted for individual papers. To extend the closing date by
one month, a written request must be filed with the ASCE Managing penetration process, but a penetration process in a much simpler
Editor. The manuscript for this paper was submitted for review and pos- material 共one typically assumed to follow the Tresca or Mises
sible publication on June 9, 2005; approved on May 25, 2006. This paper yield criterion and to have zero dilatancy兲. It is interesting that the
is part of the International Journal of Geomechanics, Vol. 7, No. 4, results these pioneers obtained support facts we now accept for
August 1, 2007. ©ASCE, ISSN 1532-3641/2007/4-251–265/$25.00. penetration processes in soil, such as the value of the ratio of qc to

INTERNATIONAL JOURNAL OF GEOMECHANICS © ASCE / JULY/AUGUST 2007 / 251

Int. J. Geomech., 2007, 7(4): 251-265


undrained shear strength in clays and the need to push a pen-
etrometer down several diameters before a steady-state pressure is
achieved.
After metal indentation, weapons research, stimulated by the
cold war, motivated further research on cavity expansion, notably
the work of Chadwick 共1959, 1962兲, who first introduced friction
into the analysis, adopting the Mohr–Coulomb criterion with an
associated flow rule. The interests then were related to explosions
within the ground and how the stress waves generated by these
Downloaded from ascelibrary.org by UFRGS - Universidade Federal do Rio Grande do Sul on 01/09/20. Copyright ASCE. For personal use only; all rights reserved.

explosions would propagate. Geomechanics with more of a geo-


technical engineering flavor followed, notably with works by
Ladanyi 共1963兲, who was interested in cavity expansion in clays,
and Palmer 共1972兲, who studied applications of cavity expansion
analyses to the pressuremeter test. There was some further work
by Vésic 共1972兲 and Baligh 共1976兲, who attempted to capture the
important feature of soil stress–strain nonlinearity, but in a way
that was empirically based. The next generation of analyses ap-
peared in the 1980s, 1990s, and 2000s 共notably, Carter et al. 1986;
Yu and Houlsby 1991; Collins et al. 1992; Salgado et al. 1997a,b;
Salgado and Randolph 2001兲. Fig. 1. Cavity expansion: Cavity and the plastic and elastic zones
In this paper, our focus is on penetration processes, which are
associated with cavity creation in the soil. If a cavity is expanded
from a finite initial radius, an increasing pressure is always re-
quired for continued expansion. In contrast, in a cavity creation created, the horizontal stress was equal to ␴h = p0 everywhere in
process, the cavity is expanded from zero initial radius and there- the soil. After cavity creation, a plastic zone is created in the
fore immediately reaches a steady-state condition. In the steady- immediate vicinity of the cavity. The radius of the plastic zone is
state condition, ongoing expansion happens at constant cavity R. The plastic zone is bounded by a nonlinear elastic zone, which
pressure; this pressure is referred to as the limit pressure pL. extends from R to a radius A and is itself bounded by a linear
elastic zone. The shear strains in the linear elastic zone are
smaller than the threshold strain 共typically between 10−6 and 10−4兲
Relationship to Penetration Resistance below which soil can be assumed to behave as a linear elastic
material having a shear modulus G0 and Poisson’s ratio ␯0. In the
Penetration resistance, whether cone resistance qc or limit unit
nonlinear elastic zone, the stresses have not reached the failure
base resistance qbL in the case of a flat-base penetrometer or pile,
criterion in terms of peak strength, but the shear strains are larger
can be related to the relevant intrinsic soil variables and soil state
than the elastic threshold strain. Behavior there, even if in reality
variables. The intrinsic variables depend only on the nature of the
inelastic, can be modeled as nonlinear elastic, with the elastic
sand particles 共mineralogy, size, shape, and surface characteristics
parameters changing continuously from the interface between the
of the soil particles兲. In uncemented, purely frictional soils, the
linear and the nonlinear elastic zones to the elastic–plastic inter-
relevant state variables are relative density DR and effective stress
face. There is also nonlinearity in the plastic zone: when steady-
state 共␴v and ␴h兲. Note that we omit the prime symbol because all
state expansion is in progress 共i.e., a limit pressure has been
of the stress variables in this paper are effective stresses and there
reached兲, the friction angle varies from a value equal to the
is no need to differentiate between total and effective stresses.
critical-state friction angle ␾c at the cavity wall to a value equal to
When a penetrometer is pushed vertically in the soil, it creates
the peak friction angle ␾ p at the elastic–plastic interface.
and expands a cylindrical cavity. Thus, there is a relationship
The cavity expansion analysis is done in terms of radii r. A
between penetration resistance and the pressure required to ex-
typical thin shell is bounded by inner and outer radii ri and r j
pand a cylindrical cavity in the soil from zero initial radius. The
measured from the center of the cavity, which has radius a. A
cylindrical cavity expansion pressure is a function of the initial
combination of the yield criterion
lateral effective stress ␴h = Ko␴v in the soil. Sands with a given
vertical effective stress can have different lateral effective
stresses, and, therefore, different qc values, if the overconsolida- ␴r = Nij␴␪ 共1兲
tion ratios 共and therefore Ko兲 are different. Experimental evidence for purely frictional soils, with the equilibrium equation
共Jamiolkowski et al. 1985; Baldi et al. 1986; Houlsby and Hitch-
man 1988; Houlsby and Wroth 1989; Mayne and Kulhawy 1991兲 d␴r ␴r − ␴␪
indicates that qc correlates quite well with the stress state as ex- + =0 共2兲
dr r
pressed by the lateral effective stress ␴h, providing additional
support for relating qc 共and qbL兲 to the cylindrical cavity limit yields
pressure. We return to this point later in the paper.

冉冊
Nij−1
rj Nij
Numerical Cavity Expansion Formulation ␴ ri = ␴ r j 共3兲
ri
Plastic Zone where ␴r⫽effective radial stress; ␴␪⫽effective hoop stress around
Fig. 1 shows a horizontal cross section of an expanding cylindri- the expanding cylindrical cavity; Nij⫽flow number at the center
cal cavity in an infinite soil mass. Initially, before the cavity was of the element, given by

252 / INTERNATIONAL JOURNAL OF GEOMECHANICS © ASCE / JULY/AUGUST 2007

Int. J. Geomech., 2007, 7(4): 251-265


Table 1. Properties of the Sands Used in the Calibration Chamber Tests
␥d,min ␥d,max D50 D10 ␾c
Sand 共kN/ m3兲 共kN/ m3兲 Gs emin emax U 共mm兲 共mm兲 共deg兲 Cg eg ng
H 14.3 17.2 2.72 0.548 0.871 1.91 0.45 0.21 36.0 942 1.96 0.46
M0 14.0 16.6 2.65 0.570 0.860 1.60 0.37 0.25 37.0 326 2.97 0.50
O 14.6 17.6 2.65 0.48 0.78 1.48 0.39 0.23 29.0 612 2.17 0.44
T1 13.6 16.7 2.68 0.574 0.931 — 0.58 0.36 34.8 647 2.27 0.43
T2 13.6 16.7 2.68 0.574 0.931 — 0.58 0.36 34.8 647 2.27 0.43
Downloaded from ascelibrary.org by UFRGS - Universidade Federal do Rio Grande do Sul on 01/09/20. Copyright ASCE. For personal use only; all rights reserved.

T4 13.6 16.7 2.68 0.575 0.924 1.50 0.54 0.36 34.8 647 2.27 0.43
T9 13.6 16.4 2.68 0.605 0.930 1.35 0.55 0.36 34.8 647 2.27 0.43
Y 13.1 16.1 2.65 0.611 0.985 1.27 0.23 — 35.1 900 2.17 0.40
Note: Data after Bolton 共1986兲, Lo Presti 共1987兲, and Lo Presti et al. 共1992兲. H⫽Hokksund; M0⫽Monterey No. 0; O⫽Ottawa Sand; Y⫽Toyoura Sand;
T⫽Ticino Sand; ␥d,min determined using ASTM D4254-83; and ␥d,max determined using ASTM D4253-83 for Monterey No. 0 sand and pluviation 共Miura
and Toki 1982兲 for the other sands.

1 + sin ␾ij in n elements ij 共with n being a large number兲. If we look at the


Nij = 共4兲 stresses and strains experienced by the very last element of the
1 − sin ␾ij
plastic zone 共the one adjacent to the elastic–plastic interface兲 at a
and ␾ij⫽friction angle within the very thin shell ij. The friction given time and by the element next to it 共in the direction of the
angle, in this paper, is assumed to vary according to the Bolton cavity兲 at the end of the previous time increment, when it was the
model 共1986兲. It is also possible to use other constitutive models outermost element of the plastic zone, we see that they are iden-
in the cavity expansion analysis. Salgado and Randolph 共2001兲, tical. This is true of any other pair of elements in the plastic zone:
for example, used the state parameter model 共Been and Jefferies the stresses and strains in an element at a given time are the same
1985兲. However, for uncemented sands, the Bolton 共1986兲 model as the ones experienced by the element interior to it at the end of
is a simple model that works quite well because it captures the the previous time increment. The outer element experiences con-
essence of soil behavior. In addition, the Bolton 共1986兲 model ditions identical to those experienced by the inner element one
accounts for the fact that the critical state line of sands is not time increment earlier in the cavity expansion process. This fact
straight, as assumed in the classical state parameter model. For can be used to calculate the strain increments for a given element
plane-strain conditions, which are operative in cylindrical cavity when the plastic radius R increases by some small amount during
expansion, the Bolton 共1986兲 model gives the relative dilatancy cavity expansion. These strain increments are tied together by the
index IR, friction angle ␾ and dilatancy angle ␺ within an element dilatancy angle, following the flow rule of Rowe 共1962兲. Salgado
ij as and Randolph 共2001兲 present the mathematics of how this is done

IR,ij =
100
冋 冉 冊册
DR,ij
Q + ln
pA
100pij
− RQ 共5兲
in cavity expansion analysis. Starting from the plastic radius,
where stresses, strains and the radial displacement are known,
calculations proceed inward, toward the cavity, element by ele-
ment. At the elastic–plastic interface, the radial stress ␴R and the
␾ij = ␾c + 5IR,ij 共6兲 mean effective stress pR are given by 共Salgado et al. 1997b;
Salgado and Randolph 2001兲:
IR,ij
sin ␺ij = 共7兲 2N p
6.7 + IR,ij ␴R = p0 共8兲
Np + 1
where DR,ij⫽relative density ranging from 0 to 100% within the
thin ij shell; Q, RQ⫽fitting parameters that depend on the sand
characteristics; pij⫽mean effective stress at the center of the ij
shell; and PA⫽reference stress⫽100 kPa= 0.1 MPa⬇ 1 tsf. Al-
1
pR = 关1 + ␮R兴 1 +
3
冉 冊
1
Np
␴R 共9兲

though the values of Q and RQ have been specifically determined with


for very few sands 共see Table 1兲, the values proposed by Bolton 1
共1986兲, Q = 10 and RQ = 1, are recommended for use in the ab- ␮R = 2 共1 + sin ␾ p sin ␺ p兲 共10兲
sence of laboratory testing that would allow making a specific
determination of these values, as they produce satisfactory results uR N p − 1 p0
for most silica sands. ␧T = = 共11兲
R N p + k 2G
Equation 共3兲 expresses how the radial stress varies within a
thin shell bounded by internal radius ri and external radius r j. A
␧r共r = R兲 = ␧R = k␧T 共12兲
relationship also exists between the accumulated volumetric, ra-
dial, and hoop strains in an element; ri and r j; and the radial where p0 = ␴h; ␧T⫽hoop strain at the elastic–plastic interface;
displacements ui and u j at r = ri and r = r j. Cavity expansion analy- ␧r⫽radial strain; N p⫽peak flow number, related to ␾ p through Eq.
sis has the mathematical property of self-similarity. The easiest 共4兲; and k = 1 for cylindrical cavity expansion and k = 2 for spheri-
way to understand this is to consider a cavity undergoing steady- cal cavity expansion. Note that iterations are required to obtain
state expansion both at a fixed moment in time and at various compatible values of ␾ p, ␴R, pR, and ␮R 共all at the elastic–plastic
times during expansion. Let us first consider that at some moment interface兲. These iterations involve using pij = pR in Eq. 共5兲 to
in time the cavity has expanded to the point such that the plastic calculate ␾ p using Eq. 共6兲. At some point in this process, an
radius is equal to R and that we have subdivided the plastic zone element will be reached for which the inner boundary displace-

INTERNATIONAL JOURNAL OF GEOMECHANICS © ASCE / JULY/AUGUST 2007 / 253

Int. J. Geomech., 2007, 7(4): 251-265


ment is equal to the inner radius of the element. The implication dR R
is that the inner boundary of the element coincides with the wall = 共15兲
da a
of a cavity expanded from zero initial radius. The limit pressure is
the internal stress for this element. Combination of Eq. 共15兲 with the linear elastic, perfectly
plastic stress rate–strain rate relationship and the small strain-
Elastic Zone displacement relations leads to an equation from which the limit-
The stress ␴R at the elastic–plastic interface induces a stress and a ing ratio 共R / a兲L can be obtained. It is not strictly correct to use
strain field outside the plastic zone that must be consistent, for this criterion if the soil is not linear elastic, perfectly plastic.
every radius r ⬎ R, with the value of the shear modulus there. Use Although it is possible to develop numerical algorithms that
Downloaded from ascelibrary.org by UFRGS - Universidade Federal do Rio Grande do Sul on 01/09/20. Copyright ASCE. For personal use only; all rights reserved.

of a nonlinear elastic stress–strain relationship ensures this con- would produce equivalent linear material properties that would
sistency. It is possible, by integration of strains, to obtain an then be fed into the criterion to determine whether a limit condi-
equivalent linear value of G for the elastic zone. This value of G tion has been reached, what happens at the cavity wall happens in
can be used to calculate the strains for the first element of the response to material properties that are local and vastly different
plastic zone 共the element adjacent to the elastic–plastic interface兲, from equivalent linear parameters. It turns out that it is key to
which allows the results of the calculations for the elastic zone to capture this local response if we desire an accurate cavity expan-
feed into the calculations for the plastic zone. sion analysis.
The initial soil stiffness is represented by the small-strain shear Chadwick 共1959兲 and latter Yu and Houlsby 共1991兲 took the
modulus G0. The small-strain shear modulus G0 of sand can be steady state as that for which the limit of the ratio of the initial to
computed from the following expression 共Hardin and Black the current cavity radius approaches zero. This does indeed rep-
1968兲: resent the notion of cavity creation and is well suited to analytical

冉 冊
solutions based on a simple constitutive relation, for which math-
G0 共eg − e兲2 ␴m ng
ematical limits can be taken, but is difficult to use effectively in
= Cg 共13兲
pA 1+e pA numerical solutions. The numerical difficulty resides in ad-
equately capturing this limit process. If calculations are done by
where e⫽initial void ratio; ␴m = p⫽initial mean effective stress;
expanding a cavity from some initial radius, the criterion tells us
and Cg, eg, ng⫽intrinsic variables of the soil.
that we would need to keep expanding this cavity forever in order
to reach a solution. So the numerical difficulty is in answering the
Relationship to Analytical Cavity Expansion Solutions question: at what cavity expansion radius short of infinity have
we obtained a sufficiently good approximation? Although ap-
Limit Cavity Pressure proximations can be found, it would be difficult to control the
In the discussion that follows, we focus on the closed-form ana- accuracy of a numerical solution based on this criterion.
lytical solutions of Carter et al. 共1986兲 and Yu and Houlsby In this paper, the steady-state criterion used is that the calcu-
共1991兲 for linear elastic, perfectly plastic soil. In a cavity creation lated displacement at the inner boundary of the cylindrical ele-
process, once the cavity is expanded from zero initial radius and ment must equal its inner radius. This expresses the physical
reaches a steady-state condition, ongoing expansion happens at notion of cavity creation rigorously and has the advantage of
constant limit cavity pressure. It is necessary to express math- being easily implemented in a numerical formulation. So compu-
ematically the requirements for a steady-state condition to be tations proceed from the plastic radius towards the cavity. The
reached and thus for the limit pressure to be calculated. plastic radius, the starting radius, can be any number, as cavity
Closed-form analytical solutions can be obtained for the limit expansion in the free field is a self-similar problem and what
cavity pressure if the assumptions of linear elasticity and perfect matters is the ratio of R to a, and not their absolute values. We
plasticity can be made. The limit pressure can be linked with the took R = 100 in our calculations.
radial stress at the elastic-plastic interface using Eq. 共3兲 with ri
= a, r j = R and a single value of N 共which, for a perfectly plastic Equivalent Linear Values of Plastic and Elastic Parameters
material, is constant throughout the plastic zone兲: Engineers wishing to use the analytical cavity expansion solutions

冉冊 共N−1兲/N need guidance on which values of shear modulus and friction and
R dilatancy angles to use. The present analysis can produce these
pL = ␴R 共14兲
a L values. It is of particular interest to investigate how the equivalent
In Eq. 共14兲, N must be an equivalent linear flow number in linear values of ␾, ␺, E, and ␯ vary with initial soil state, i.e.,
order to produce the same limit pressure as a full nonlinear analy- initial relative density and stress state. For this purpose, calcula-
sis. The subscript L indicates that the R / a ratio is a limiting tions are done for the following values of the intrinsic values Cg,
共maximum兲 value. Eq. 共14兲 cannot be used directly to calculate eg, ng, emin, emax, Q, and RQ:
the limit pressure because the ratio of the plastic to cavity radius Cg = 650, eg = 2.2, ng = 0.45
is not known a priori; however, combined with a suitable steady-
state criterion, an equation can be derived, based on Eq. 共14兲, that emin = 0.60, emax = 0.95
can be used to calculate the limit pressure.
Q = 10, RQ = 1
Steady-State Criteria
Carter et al. 共1986兲 proposed a steady-state criterion based on Eq. These values are representative of typical silica sands. The
共14兲. For the pressure pL to indeed be a limit pressure, R / a needs relative density values considered are DR = 10, 20, 30, 40, 50, 60,
to be equal to the limit value 共R / a兲L. Stated in another way, in 70, 80, 90, and 100%. The coefficient of active earth pressure at
steady-state cavity expansion, even as the cavity radius a and the rest is in the 0.4–0.5 range in most cases in practice, with dense
plastic radius R increase, their ratio remains the same. That hap- sands tending to have lower, and loose sands having higher val-
pens if and only if: ues. For simplicity, a single value of 0.45 is taken for all values of

254 / INTERNATIONAL JOURNAL OF GEOMECHANICS © ASCE / JULY/AUGUST 2007

Int. J. Geomech., 2007, 7(4): 251-265


Downloaded from ascelibrary.org by UFRGS - Universidade Federal do Rio Grande do Sul on 01/09/20. Copyright ASCE. For personal use only; all rights reserved.

Fig. 2. G / G0 for clean sand according to the Ishibashi and Zhang


共1993兲 model

relative density. The maximum vertical effective stress considered


is that for a depth z = 50 m and water table deeper than 50 m.
Taking the unit weight as roughly 20 kN/ m3, this would corre-
spond to 1 MPa vertical effective stress. Values of vertical effec-
tive stress are therefore varied from 50 kPa to 1 MPa in 50 kPa
increments.
The equivalent linear values of the elastic parameters in the
elastic zone are estimated, as indicated earlier, by integration of
the radial strains in the elastic zone, which in turn are obtained by
considering a nonlinear stress–strain relationship, and using the
amount of compression there to obtain the elastic constant values
that would produce the same compression for a linear elastic soil.
We have found that the nonlinear elastic model does not need to
be overly sophisticated to produce satisfactory results. The results
presented in this paper are for the Ishibashi and Zhang 共1993兲
model, which, for sands, takes the form:
G
= K␴mn 共16兲
G0 Fig. 3. Equivalent linear shear modulus as function of initial soil
where K and n are given by state: 共a兲 ␾c = 29°; 共b兲 ␾c = 36°

再 冋 冉冉 冊 冊册冎
K = 0.5 1 + tanh ln
0.000102

0.492
共17兲

再 冋 冉冉 冊 冊册冎
n = 0.272 1 − tanh ln
0.000556

0.4
共18兲
The equivalent linear shear modulus for steady-state cavity
expansion, normalized with respect to the small-strain shear
modulus, G / G0, is plotted in Fig. 3 for ␾c = 29° and ␾c = 36°. It
where ␥⫽engineering shear strain 共as a number, not a can be seen that G / G0 does not vary much with relative density or
percentage兲. ␾c, but does vary significantly with lateral effective stress. To
The radial and hoop stresses are computed at the center of thin some extent, this is due to the nature of the Ishibashi and Zhang
cylindrical shell elements in the nonlinear elastic zone using the 共1993兲 model, but the calculations show that the cavity expansion
equations for an elastic hollow cylinder of infinite radius 共in the process does not introduce a strong dependence on variables other
case of cavity expansion in the field, as opposed to in a calibration than the stress. For ␴h = 22.5 kPa, G / G0 lies in the 0.52–0.6 range
chamber, when the equations for a hollow cylinder with a defined for ␾c = 29° and in the 0.5–0.58 range for ␾c = 36°. For ␴h
outer boundary are used兲 with internal pressure equal to ␴R. These = 450 kPa, G / G0 lies in the 0.86–0.9 range for ␾c = 29° and in the
stresses then are used to compute the corresponding strains. Here 0.8–0.88 range for ␾c = 36°.
iterations are necessary, as we initially assume a value of G then The linear-equivalent flow number N in the plastic zone at
iterate with Eq. 共16兲 to guarantee that the calculated strain and the steady state can be determined from the rigorous analysis by con-
modulus are compatible. Fig. 2 shows the modulus degradation sidering directly Eq. 共14兲. For a given plastic radius R, the analy-
共G / G0 versus ␥兲 curves for various mean effective stress values sis produces values of ␴R, pL and a at steady state. It follows that
produced by the Ishibashi and Zhang 共1993兲 model. N can be calculated as

INTERNATIONAL JOURNAL OF GEOMECHANICS © ASCE / JULY/AUGUST 2007 / 255

Int. J. Geomech., 2007, 7(4): 251-265


ln 冉冊
R
a

冉冊 冉 冊
L
N= 共19兲
R pL
ln − ln
a L ␴R
The linear-equivalent friction angle ␾ is then obtained from N
through the general equation relating flow number to friction
angle:
Downloaded from ascelibrary.org by UFRGS - Universidade Federal do Rio Grande do Sul on 01/09/20. Copyright ASCE. For personal use only; all rights reserved.

N−1
sin ␾ = 共20兲
N+1
Once ␾ is known, the linear-equivalent dilatancy angle can be
obtained from the values of ␾ and ␾c using Eqs. 共5兲–共7兲. The ␾
calculated using Eq. 共20兲 is plotted in Fig. 4. Fig. 4 shows a plot
of ␾ – ␾c versus initial lateral stress for three values of critical-
state friction angle ␾c: 29, 33, and 36°. This covers the full range
of ␾c values for silica sands. It can be seen from Fig. 4 that the
plots are, in essence, independent of ␾c for DR = 0 – 60% but start
separating for larger relative densities. Even so, for DR = 100%,
the difference between ␾ – ␾c values for ␾c = 29 and 36° is only of
the order of 1°. For very loose sand, the equivalent ␾ was allowed
to dip below 0° 共which it does by no more than a few degrees兲. In
this, we followed the argument put forth by Bolton 共1986兲, who
linked the occurrence of ␾ values below ␾c using his equation to
the fact that, for very loose sands, the peak and critical-state fric-
tion angles 共which would be very close兲 would occur at such large
shear strains that it would be acceptable to use ␾ ⬍ ␾c.
It is interesting to explore how the ratio 共R / a兲L of the plastic
radius to the cavity radius at steady state 共when the limit pressure
has been reached兲 varies with initial soil state. This is a quantity
of interest not only in that it is an integral part of the calculations
of limit pressure but also in that the question of how far the
effects of cavity creation extend is often asked in practical prob-
lems. For example, if experiments are being conducted on pen-
etrometers, the question of how far a boundary can be placed
without excessively distorting the mechanism the experiments
aim to study is an important one, and the extent of the plastic
zone can guide us on that account.
Fig. 5 shows 共R / a兲L versus lateral effective stress for ␾c = 29
and 36° and DR = 10– 100%. Note first that the values of 共R / a兲L
are practically independent of ␾c. Note also that the plastic radius
to cavity radius ratio is greater than 100 for dense sands and as
high as 70 for medium dense sands at low confining stresses
共0 – 50 kPa兲, which correspond to depths often of interest in geo-
technical engineering problems. This means that the use of cali-
bration chambers or centrifuges to study cone penetration must be
guided by rigorous theory so that due account can be taken of
boundary effects. Insufficient distance between boundaries and
the cone allows boundaries to interfere with the formation of the Fig. 4. Equivalent linear friction angle as function of initial soil state:
plastic zone around the cone, and this interference is substantial at 共a兲 ␾c = 29°; 共b兲 ␾c = 33; and 共c兲 ␾c = 36°
low confinements.
The plots of 共R / a兲L can be used to obtain suitable values of
共R / a兲L for use in Eq. 共14兲. Since the radial stress at the elastic– Penetration Resistance Analysis
plastic interface follows from Eq. 共8兲, limit cavity pressures can
then be calculated using Eq. 共14兲. A more direct way to obtain
Basic Analysis for Cone Penetrometers
limit pressure is to use the charts in Fig. 6, in which limit pressure
versus initial lateral effective stress is plotted for DR = 10– 100% The penetration resistance of a penetrometer with flat or conical
and ␾c ranging from 29 to 36°. There is a gradual shift to the right tip may be calculated from the cylindrical limit pressure pL deter-
in the pL versus stress curves as ␾c increases from 29 to 36°. The mined in the previous section. Spherical cavity pressures have
maximum limit pressure, calculated for DR = 100% and 1 MPa also been proposed in the past for the same purpose, although in
vertical stress 共450 kPa lateral stress兲, is approximately 15 MPa. papers typically focusing on clays. For sands, Mitchell and Tseng

256 / INTERNATIONAL JOURNAL OF GEOMECHANICS © ASCE / JULY/AUGUST 2007

Int. J. Geomech., 2007, 7(4): 251-265


conical tip tends to be horizontal, and can thus be linked to cy-
lindrical cavity expansion. A stress rotation analysis can then be
used to account for the rotation of principal stresses from vertical
immediately below the conical tip to horizontal away from it.
Although a rigorous analysis that incorporates all the main fea-
tures of sand mechanical response would be very difficult, an
analysis with sufficient accuracy can be obtained if a few simpli-
fying assumptions are made.
Fig. 7 shows a plane-strain, kinematically admissible slip
Downloaded from ascelibrary.org by UFRGS - Universidade Federal do Rio Grande do Sul on 01/09/20. Copyright ASCE. For personal use only; all rights reserved.

mechanism that can be used to develop a suitable analysis. If we


can assume a similar, axisymmetric mechanism exists for a pen-
etrometer with a conical tip, the conical surface, during penetra-
tion, is in effect a slip surface with interface friction angle ␦c. The
semiapex angle of the cone is ␪c. From the point of view of the
stress field, there is a transition zone between the cone and the
zone where the major principal effective stress is horizontal and
associated with the expansion of a cylindrical cavity. The transi-
tion zone is composed of an infinite number of log-spiral surfaces,
each represented by an equation with the form:

s = s0 exp共⌬␭ tan ␺T兲 共21兲


where s⫽log-spiral radius measured from point B, the edge of the
cone where the conical surface meets with the cylindrical shaft of
the penetrometer; s0⫽log-spiral radius BC; ⌬␭⫽angle between
the log-spiral radius s at which the principal stress becomes hori-
zontal and s0 = BC; ␺T⫽operative dilatancy angle for the transi-
tion zone. The initial point of the log-spiral, lying on the conical
surface, has radius r0, measured from the axis of the cone; the end
point, lying on the line where the major principal stress turns
horizontal, has radius r given by

r = rc共1 + C␭兲 − r0C␭ 共22兲


where rc⫽penetrometer radius,


sin共⌬␭ − ␪c兲
exp共⌬␭ tan ␺T兲 if ␾T ⱖ ␾c
sin ␪c
C␭ = 共23兲
sin共⌬␭ − ␪c兲
if ␾T ⬍ ␾c
sin ␪c

Fig. 5. Ratio of plastic radius to cavity radius when cylindrical cavity ␲ ␦c


⌬␭ = + + ␪c 共24兲
limit pressure is reached: 共a兲 ␾c = 29°; 共b兲 ␾c = 36° 4 2
and ␾T⫽operative friction angle for the transition zone.
Following a stress rotation analysis 关such as done by Bolton
共1990兲 observed that pure spherical cavity pressures would not 共1982兲 for footings兴, the major principal stresses in zone P and on
work and explored averaging with cylindrical cavity pressures. the face Q of the cone, separated by the transition zone with angle
Spherical cavity expansion analysis is possible for a soil medium ⌬␭, are related through:
subjected to an isotropic stress state. Accordingly, for an aniso-
tropic stress state 共with ␴v different from ␴h兲, the mean stress ␴m, ␴Q1 = ␴1P exp共2⌬␭ tan ␾T兲 共25兲
as an approximation, is used in calculations. In contrast, cylindri- where
cal cavity expansion can be directly analyzed for a soil medium in ␾T = ␾c + 5IRT⫽transition zone friction angle; ␴Q1 ⫽major prin-
which an anisotropic stress field is established, and the lateral cipal effective stress acting on the conical surface at the initial
stress ␴h is used in calculations. Aside from the fact that the cone point of a given log-spiral; IRT⫽value of IR in the transition zone;
plus the rods to which it is attached 共like a pile兲 is a cylindrical
and ␴1P⫽major principal effective stress acting in the horizontal
body that must expand a cylindrical cavity in the soil in order to
direction at the end point of the same log-spiral 共which is also the
advance, the most compelling argument in favor of calculating qc
radial stress due to cavity expansion acting at the end of the log
from cylindrical cavity limit pressure 共aside from the fact that it
spiral兲, given by
works兲 is the very strong correlation between qc and ␴h, not ␴m,
observed in calibration chamber tests done by Houlsby and Hitch-
man 共1988兲 and Schnaid and Houlsby 共1991兲. Salgado et al.
共1997a,b兲 also showed that, for sands, the displacement field in-

␴1P = pL 1 + C␭ − C␭
r0
rc
冊 ␩−1
共26兲

duced by cone penetration outside the immediate area below the where

INTERNATIONAL JOURNAL OF GEOMECHANICS © ASCE / JULY/AUGUST 2007 / 257

Int. J. Geomech., 2007, 7(4): 251-265


Downloaded from ascelibrary.org by UFRGS - Universidade Federal do Rio Grande do Sul on 01/09/20. Copyright ASCE. For personal use only; all rights reserved.

Fig. 6. Cylindrical cavity limit pressure as a function of initial soil state for a normally consolidated sand with critical-state friction angle equal
to: 共a兲 29; 共b兲 30; 共c兲 31; 共d兲 32; 共e兲 33; 共f兲 34; 共g兲 35; and 共h兲 36°

␩=1−
NT − 1
NT
共27兲 冉
␴Q1 = pL exp共2⌬␭ tan ␾T兲 1 + C␭ − C␭
r0
rc
冊 ␩−1
共28兲

The associated minor principal effective stress is given by


rc = 0.5 dc⫽cone radius and NT⫽transition zone flow number,
computed from the friction angle ␾T operative within the stress
rotation zone. ␴Q1
Note, from Eq. 共26兲, that ␴1P decreases monotonically from a ␴Q3 = 共29兲
Nc
value equal to pL at the cone edge with distance r from the axis of
the cone. The stress field in zone P is defined by the cylindrical
cavity expansion solution; the stresses within the mechanism and, where the critical-state flow number Nc is used because ␴Q1 and
in particular, at the cone face are determined by Eq. 共25兲. ␴Q3 are the principal stresses at the cone surface, where the sand
Cone resistance qc depends on the friction angle ␾T, which in deformation is large and critical-state conditions have been
turn depends on qc because qc creates large confining stresses in reached. Knowing the values of ␴Q1 and ␴Q3 at each point of the
the zone around the cone, inhibiting dilatancy and keeping ␾T conical surface, it is possible, through integration over the area of
close in value to ␾c. As qc depends on ␾T, and ␾T depends on qc the cone surface, to obtain the total vertical force opposing pen-
through the confining stresses that qc creates, the calculation of qc etration. Dividing the resulting vertical force by the projected
is done iteratively. Using Eqs. 共25兲 and 共26兲: cross-sectional area of the cone gives the cone resistance qc:

258 / INTERNATIONAL JOURNAL OF GEOMECHANICS © ASCE / JULY/AUGUST 2007

Int. J. Geomech., 2007, 7(4): 251-265


Downloaded from ascelibrary.org by UFRGS - Universidade Federal do Rio Grande do Sul on 01/09/20. Copyright ASCE. For personal use only; all rights reserved.

Fig. 6. 共Continued兲.

共1 + C␭兲␩+1 − C␭共␩ + 1兲 − 1
qc = 2f v pL exp共2⌬␭ tan ␾T兲 共30兲
C␭2 ␩共␩ + 1兲

with

fv =
1
2
冋冉 冊 冉 冊
1+
1
Nc
+ 1−
1
Nc
共cos ␦c cot ␪c − sin ␦c兲 册 共31兲

The starting point for finding the average mean effective stress
in the transition zone, required for the calculation of ␾T, is to
realize that qc⫽average vertical stress acting on the face of the
cone. Working back from that, we find that the average radial
stress acting on the conical face of the penetrometer is given by

共1 + C␭兲␩+1 − C␭共␩ + 1兲 − 1
¯␴r = 2pL 共32兲
C␭2 ␩共␩ + 1兲

and the associated mean effective stress can be written as Fig. 7. Assumed slip mechanism for cone penetration in sand

INTERNATIONAL JOURNAL OF GEOMECHANICS © ASCE / JULY/AUGUST 2007 / 259

Int. J. Geomech., 2007, 7(4): 251-265


p̄i =
1
2
冉 冊
1+
1
NT
¯␴r 共33兲
p̄ = p̄i exp关2共⌬␭ − ␰兲tan ␾T兴 共34兲

Consider the slipline starting at the point on the conical sur- where ␰⫽angle ranging from 0 to ⌬␭ between an arbitrary log-
face where ␴v = qc 共which we may choose to call the “average” spiral radius s and s0 = BC.
slip line兲. The mean effective stress varies along this slip line Taking the average of p̄ over the slip line, we obtain the aver-
following: age mean stress p̄T in the transition zone:
Downloaded from ascelibrary.org by UFRGS - Universidade Federal do Rio Grande do Sul on 01/09/20. Copyright ASCE. For personal use only; all rights reserved.


p̄i exp共2⌬␭ tan ␾T兲 − exp共⌬␭ tan ␺T兲
if ␾T ⱖ ␾c
2 tan ␾T cot ␺T exp共⌬␭ tan ␺T兲 − 1
p̄T = 共35兲
exp共2⌬␭ tan ␾T兲 − 1
p̄i if ␾T ⬍ ␾c
2⌬␭ tan ␾T

Computation of qc from pL is done iteratively as follows: The predictive power of cavity-expansion-based cone resis-
1. Assume value for ␾T 共and thus for ␺T and NT兲; tance theory is further illustrated by Fig. 10 for Ticino sand. Fig.
2. Compute ␩, ⌬␭, and C␭ using Eqs. 共27兲, 共24兲, and 共23兲, 10 shows qc curves obtained for Ticino sand using CONPOINT
respectively; plotted together with data points from calibration chamber tests
3. Compute ␾T using the Bolton relationship 关Eqs. 共5兲–共7兲 with on Ticino sand. The CONPOINT curves approximate the cone
pij = p̄T兴; resistance measured in calibration chambers reasonably well for
4. Iterate until convergence of assumed and calculated values of
comparable relative densities.
␾T is reached; and
5. Compute qc using Eqs. 共30兲 and 共31兲.
There is no single value of friction angle in the immediate
neighborhood of the cone, as the stress and strain fields are com- Correlations for Penetration Resistance
plex and have sharp gradients. Along the conical surface and the
penetrometer shaft, where shear strains are quite large, ␾ is equal
to ␾c both for contractive and dilative sands. Accordingly, the Cone Resistance Charts
interface friction angle is also a critical-state value ␦c that is taken
as a fraction of ␾c. However, near the cone surface but slightly The preceding analysis allows us to develop charts and correla-
away from the steel–soil interface, stresses are very high and the tions for qc in terms of key intrinsic and state variables of the
sand is likely to undergo some crushing, which is difficult to sand. Fig. 11 shows qc in terms of relative density and stress state
model. for eight different values of ␾c. These charts can be used directly
Bolton 共1986兲 suggests it may be necessary to allow the ␾ to estimate qc if the sand properties are similar to those used to
calculated using Eqs. 共5兲–共7兲 to be less than ␾c under some con-
prepare the charts 共recall that these values are Cg = 650, eg = 2.2,
ditions, as very contractive sands would require such large strains
ng = 0.45, emin = 0.60, emax = 0.95, Q = 10, and RQ = 1兲. For different
to develop ␾c that the operative, mobilized friction angle would
be less than ␾c under many conditions of practical interest. For values of these parameters, use of CONPOINT is indicated. Al-
dilative sands, either sands with high relative densitites or low to ternatively, with knowledge of K0 共perhaps from knowledge of
moderate initial confining stresses, ␾ values tend to be greater the geological history of the soil deposit兲 and qc, the relative
than ␾c even within the transition zone. Fig. 8 shows values of density may be estimated.
␾T – ␾c as a function of relative density and lateral stress for ␾c Observing Fig. 11, we note that qc values range from zero at
= 29, 33, and 36°. The curves are similar, with ␾T – ␾c tending to the ground surface due to lack of confinement to as much as
increase with increasing relative density, decreasing stress and 80 MPa at a vertical effective stress of 1 MPa in a normally-
decreasing ␾c. Note that this approach indirectly accounts for consolidated deposit with DR = 100% 共admittedly, an eminently
sand “compressibility” 共a term found in the literature to describe theoretical case兲. For cases of more relevance in on-shore prac-
what, in more appropriate usage, is contractiveness兲.
tice, a depth of 30 m with a deep water table would give us a
As the analysis is applicable to both field and calibration
more credible “large” stress. That would correspond to a vertical
chamber conditions, the quality of the predictions made using the
analysis can be checked by comparison with calibration chamber effective stress on the order of 600 kPa and a lateral effective
penetration tests. As an illustration of the excellent match be- stress of the order of 270 kPa. For a very high critical-state fric-
tween predicted and measured values, Fig. 9 shows how well qc tion angle of 36° and a relative density of 90%, qc would be about
values predicted using CONPOINT compare with values mea- 70 MPa, which would constitute a point beyond refusal for a
sured in chamber tests done on West Kowloon sand. These studies typical CPT system. For a sand with 29° critical-state friction
were done by Lee et al. 共1999兲 in the course of the engineering of angle, the same relative density 共DR = 90% 兲 and stresses would
large, hydraulically placed marine sand fills in Hong Kong, for lead to qc = 35 MPa, which might be reachable by sufficiently
which the ability to effectively interpret CPT results was very powerful systems even without friction reducers.
important.

260 / INTERNATIONAL JOURNAL OF GEOMECHANICS © ASCE / JULY/AUGUST 2007

Int. J. Geomech., 2007, 7(4): 251-265


Downloaded from ascelibrary.org by UFRGS - Universidade Federal do Rio Grande do Sul on 01/09/20. Copyright ASCE. For personal use only; all rights reserved.

Fig. 8. Transition zone friction angle ␾T as function of initial soil state: 共a兲 ␾c = 29°; 共b兲 ␾c = 33°; and 共c兲 ␾c = 36°

A ratio of interest is the ratio of qc to cylindrical cavity limit Most expressions in the form of Eq. 共36兲 have been obtained
pressure pL, which can be extracted from Figs. 6 and 11. The ratio by regressions of qc values obtained from penetration tests done
ranges from 3.2 for loose sands with high ␾c and large ␴h values in large calibration chambers 共Bellotti et al. 1982; Villet and
to 7.3 for dense sands with low ␾c and low ␴h values. Mitchell 1981; Ghionna and Jamiolkowski 1991兲. Proposal of C1,
C2, and C3 values for Eq. 共36兲 based on all available calibration
Simple Correlations chamber test results have been made by, for example, Baldi et al.
A simpler approach to penetration resistance estimation is to use 共1983, 1985, 1986兲. However, as different sands have different
an equation of the form: intrinsic variables and qc is sensitive to these variables, particu-
larly to ␾c, the use of a single expression of the form of Eq. 共36兲
qc
pA
= C1 冉 冊
␴h
pA
C2
exp共C3DR兲 共36兲
to calculate qc for all silica sands cannot be expected to have high
accuracy. Additionally, it is impossible to find constants C1, C2,
and C3 that can truly fit qc values for all possible relative densities
where DR⫽relative density, ranging from 0 to 100%; and lateral effective stresses. In fact, they cannot be constants at
pA⫽reference stress 共=100 kPa= 0.1 MPa兲 in the same units as
all if they are to provide a satisfactory fit. They must themselves
␴h; and C1, C2, and C3⫽regression coefficients. This type of ex-
be functions of the relative density and the critical-state friction
pression has been proposed by Baldi et al. 共1983, 1985, 1986兲,
Schmertmann 共1976兲, and others. angle.

INTERNATIONAL JOURNAL OF GEOMECHANICS © ASCE / JULY/AUGUST 2007 / 261

Int. J. Geomech., 2007, 7(4): 251-265


Downloaded from ascelibrary.org by UFRGS - Universidade Federal do Rio Grande do Sul on 01/09/20. Copyright ASCE. For personal use only; all rights reserved.

Fig. 9. Comparison of predicted with measured values of cone resis- Fig. 10. Cone resistance calculated both using CONPOINT with
tance for calibration chamber tests on West Kowloon sand 共adapted Ticino Sand properties and cone resistance measured in calibration
from Lee et al. 1999兲 chamber tests on normally, 1D consolidated samples of Ticino Sand
with DR within the 10–20%, 50–60%, and 90–100% ranges 共data
adapted from Salgado 1993 and Salgado et al. 1997a,b兲

Fig. 11. Cone penetration resistance using CONPOINT as a function of initial soil state for a normally consolidated sand with ␾c equal to: 共a兲
29; 共b兲 30; 共c兲 31; 共d兲 32; 共e兲 33; 共f兲 34; 共g兲 35; and 共h兲 36°

262 / INTERNATIONAL JOURNAL OF GEOMECHANICS © ASCE / JULY/AUGUST 2007

Int. J. Geomech., 2007, 7(4): 251-265


Downloaded from ascelibrary.org by UFRGS - Universidade Federal do Rio Grande do Sul on 01/09/20. Copyright ASCE. For personal use only; all rights reserved.

Fig. 11. 共Continued兲.

The following equation was found to fit qc values calculated


using the analysis in this paper quite well:

qc
= 1.64 exp关0.1041␾c + 共0.0264 − 0.0002␾c兲DR兴
pA

⫻ 冉 冊
␴h
pA
0.841−0.0047DR
共37兲

where ␾c has units of degrees. Recall that our calculations were


done for values of Cg = 650, eg = 2.2, ng = 0.45, emin = 0.60, emax
= 0.95, Q = 10, and RQ = 1.
Eq. 共37兲 predicts values for ␾c in the 29–36° range and DR in
the 0 to 90% range with error at no time greater than 10% with
respect to the values calculated using the analysis discussed here.
Given that direct use of the theory is expected, based on the
experience with both chamber tests 共Salgado et al. 1997a,b; 1998兲
and field observations 共Lee et al. 1999兲, to provide qc values
accurate to ±30% with respect to experimentally measured val-
ues, this means Eq. 共37兲 is accurate to ±40% with respect to
measured qc values, which, for such a simple equation and in light
of the many variabilities in both soil variables and experimental
setups, is quite satisfactory. It should be noted that, given the
exponential relationship of qc with DR, the inverse calculation 共DR Fig. 12. CPT log from a site in Evansville, Ind. 共test performed by
from qc兲 will be significantly more “accurate.” That equation is Dr. Junhwan Lee兲

INTERNATIONAL JOURNAL OF GEOMECHANICS © ASCE / JULY/AUGUST 2007 / 263

Int. J. Geomech., 2007, 7(4): 251-265


ln 冉 冊
qc
pA
− 0.4947 − 0.1041␾c − 0.841 ln冉 冊
␴h
pA
can be estimated as 220 kPa using 19 kN/ m3 for the soil unit
weight. A K0 of 0.4, consistent with a young NC deposit, gives

冉 冊
DR = ⱕ 100% ␴h = 88 kPa. From Fig. 12, qc = 16 MPa. Taking all of this into Eq.
␴h 共38兲, we get DR = 62%. Using Fig. 11, DR = 61%. This value of DR
0.0264 − 0.0002␾c − 0.0047 ln
pA can now be used, for example, to estimate friction angles or as-
共38兲 sess the susceptibility of the soil to liquefaction.

Summary and Conclusions


Downloaded from ascelibrary.org by UFRGS - Universidade Federal do Rio Grande do Sul on 01/09/20. Copyright ASCE. For personal use only; all rights reserved.

Cone Resistance Calculation and CPT Interpretation


The capability of accurately estimating penetration resistance for
Cone resistance calculations may be necessary for a number of sand deposits is important in applications such as interpretation of
different purposes in a number of different scenarios. In one com- CPT results for determination of relative density and stress state,
mon situation, an engineer may wish to estimate, from the results estimaton of friction angles, assessment of soil liquefaction sus-
of a CPT investigation, the relative density of the deposit at dif- ceptibility, and prediction of the capacity of pile foundations to
ferent depths, which would then allow estimation of a number of support vertical loads.
additional parameters, such as friction angle and stiffness. In the Calculation of qc can be done relatively simply if linear elas-
simplest such case, geologic information is sufficient for the in- ticity and perfect plasticity can be assumed for the soil. Values for
ference to be made that the soil deposit is normally consolidated, G, ␯, and ␾ of a linear-elastic, perfectly plastic material in terms
and thus has a K0 = 0.40– 0.45. Vertical effective stresses can be of the initial relative density and lateral stress were developed. It
calculated with reasonable certainty from soil unit weights and was shown that ␾ – ␾c is essentially independent of ␾c. The ratio
knowledge of the location of the groundwater table. Given the of plastic radius to cone penetrometer radius is also essentially
cone resistance and sufficiently accurate information on the value independent of ␾c. The representative friction angle ␾T within the
of lateral stress, Eq. 共38兲 or the charts of Fig. 11 should provide a transition zone around a penetrating cone was found to depend on
reasonable estimate of the value of DR. initial density, initial lateral effective stress and ␾c.
When we determine the value of qc at a given depth to be used Summary charts for qc directly in terms of soil state were
in calculations, we need to make an allowance for the stochastic developed. Another way of computing qc is to use a single ex-
component of qc, associated with local variations in density and pression that gives qc as a function of the lateral effective stress
particle size along the penetration path. In practice, a reasonable, ␴h and the relative density DR. This expression was derived by
simple way to do that, is to select qc by avoiding sharp peaks and fitting the results of the complete analysis over the entire range of
valleys in the qc profile. relative density and stress states considered in the paper 共0 ⱕ DR
As far as the values of the intrinsic variables 共emax, emin, Cg, eg, ⱕ 100% and 22.5ⱕ ␴h ⱕ 450 kPa兲. Example calculations show
ng, ␾c, Q, RQ兲 to use in calculations, they could in practice be that two different methods of estimating qc are easy to apply and
determined from triaxial compression tests, preferably coupled give consistent values. An equation was also provided to obtain
with bender element testing 共for estimates of Cg, eg, ng兲 and index DR in terms of qc and ␴h.
testing. CONPOINT could then be used to establish the relation-
ship between qc and soil state 共␴h and DR兲. Although feasible in
larger projects, in more routine projects it is likely that estimates
of these variables will have to be made largely based on judgment References
and analysis of such things as the grain size distribution of the
sand and the shape of the sand particles. Baldi, G., et al. 共1985兲. “Laboratory validation of in-situ tests.” AGI
Example 1. The cone resistance is to be estimated at a depth Golden Jubilee Volume for XI ICSMFE, San Francisco, 217–239.
Baldi, G., Bellotti, R., Ghionna, V. N., Jamiolkowski, M., and Pasqualini,
of 25 m within a clean sand deposit with water table at 5 m. The
E. 共1983兲. “Prova penetrométrica stática e densitá relativa della sabbia
sand is assumed to be normally consolidated with K0 = 0.45. The
– Atti del.” XV Congegno Nazionale di Geotecnica, Spoleto, 4–6,
relative density is 50%. The critical-state friction angle estimated Maggio, Vol. 1.
from triaxial compression tests is 30°. Following Bolton 共1986兲, Baldi, G., Bellotti, R., Ghionna, V. N., Jamiolkowski, M., and Pasqualini,
Q = 10 and RQ = 1. The vertical effective stress is estimated to be E. 共1986兲. “Interpretation of CPT’s and CPTU’s—second part:
296 kPa from a total unit weight of 20 kN/ m3; thus, ␴h Drained penetration of sands.” 4th Int. Conf. on Field Instrumentation
= 133 kPa. and In Situ Measurement, Nanyang Technological Institute, Sin-
• Method 1 关using Eq. 共37兲兴: qc = 12.2 MPa; and gapore, 143–156.
• Method 2 关using the cone resistance charts, Fig. 11共b兲兴: qc Baligh, M. M.. 共1976兲. “Cavity expansion in sands with curved enve-
= 12.1 MPa. lopes.” Proc. ASCE, J. Geotech. Eng. Div. 102共11兲, 1131–1146.
Example 2. Fig. 12 shows cone resistance qc versus depth for Been, K., and Jefferies, M. G. 共1985兲. “A state parameter for sands.”
a sand site in Evansville, Ind. Soil from a depth of 14.0 m was Geotechnique, 35共2兲, 99–112.
tested in the laboratory. The sand was found to have specific Bishop, R. F., Hill, R., and Mott, N. F. 共1945兲. “The theory of identation
and hardness tests.” Proc. Phys. Soc. 57, 147–159.
gravity of 2.67, emax of 0.826, and emin 0.454. The fines content is
Bellotti, R., Bizzi, G., and Ghionna, V. 共1982兲. “Design, construction and
less than 7%; the fines are mostly non-plastic silt. The water table
use of a calibration chamber.” Proc., ESOPT II, Vol. 2, Amsterdam,
was located at a depth of 9 m at the time of the test. Taking a
439–446.
critical-state friction angle of 33°, which is consistent with the Bolton, M. D. 共1982兲. A guide to soil mechanics, MacMillan, New York.
general properties of the sand and sand particles, we estimate the Bolton, M. D. 共1986兲. “The strength and dilatancy of stands.” Geotech-
relative density at 14 m. From knowledge of geologic conditions nique, 36共1兲, 65–78.
of the area, the deposit is a young normally consolidated sand Bonita, J. A. 共2000兲. “The effects of vibration on the penetration resis-
with low fines content throughout. The vertical effective stress tance and pore water pressure in sands.” Dissertation in partial fulfill-

264 / INTERNATIONAL JOURNAL OF GEOMECHANICS © ASCE / JULY/AUGUST 2007

Int. J. Geomech., 2007, 7(4): 251-265


ment of the requirements for a Ph.D. degree in Civil Engineering, Huang, ed., Elsevier, New York, 257–264.
Virginia Polytechnic Institute and State Univ. Blacksburg, Va. Mitchell, J. K. 共1988兲. “New developments in penetration tests and equip-
Carter, J. P., Booker, J. R., and Yeung, S. K. 共1986兲. “Cavity expansion in ment.” Penetration testing 1988, ISOPT-1, J. De Ruiter, ed., A. A.
cohesive-frictional soils.” Geotechnique, 36, 349–353. Balkema, Vol. 1, Rotterdam, The Netherlands, 245–261.
Chadwick, P. 共1959兲. “The quasistatic expansion of a spherical cavity in Mitchell, J. K., and Tseng, D.-J. 共1990兲. “Assessment of liquefaction
metals and ideal soils. Part I.” Q. J. Mech. Appl. Math., 12, 52–71.
potential by cone penetration resistance.” H. Bolton Seed Memorial
Chadwick, P. 共1962兲. “Propagation of spherical plastic-elastic distur-
Symp., J. M. Duncan, ed., Vol. 2, BiTech, Berkeley, Calif., 335–350.
bances from an expanded cavity.” Q. J. Mech. Appl. Math., 15共3兲,
Miura, S., and Toki, S. 共1982兲 “A sample preparation method and its
349–376.
effect on static and cyclic deformation—strength properties of sand.”
Collins, I. F., Pender, M. J., and Yan, W. 共1992兲. “Cavity expansion in
Downloaded from ascelibrary.org by UFRGS - Universidade Federal do Rio Grande do Sul on 01/09/20. Copyright ASCE. For personal use only; all rights reserved.

Soils Found., 22共1兲, 61–77.


sands under drained loading conditions.” Int. J. Numer. Analyt. Meth.
Geomech., 16, 3–23. Palmer, A. C. 共1972兲. “Undrained plane-strain expansion of a cylindrical
Ghionna, V., and Jamiolkowski, M. 共1991兲. “A critical appraisal of cali- cavity in clays.” Geotechnique 22共3兲, 451–457.
Rowe, P. W. 共1962兲. “The stress-dilatancy relation for static equilibrium
bration chamber testing of sands.” Proc., 1st Int. Symp. on Calibration
of an assembly of particles in contact.” Proc. R. Soc. London, Ser. A,
Chamber Testing (ISOCCT1), Potsdam, N.Y., A. B. Huang, ed., A269, 500–527.
Elsevier, New York, 13–39. Salgado, R. 共1993兲. “Analysis of penetration resistance in sands.” Ph.D.
Hardin, B., and Black, W. 共1968兲. “Shear modulus and damping in soils.” thesis, Univ. of California, Berkeley, Calif.
J. Soil Mech. and Found. Div., 94共2兲, 353–369. Salgado, R., Boulanger, R. W., and Mitchell, J. K. 共1997a兲. “Lateral stress
Hill, R. 共1950兲. “Mathematical theory of plasticity.” Oxford Univ. Press. effects on CPT liquefaction resistance correlatons.” J. Geotech.
Houlsby, G. T., and Hitchman, R. 共1988兲. “Calibration chamber tests of a Geoenviron. Eng., 123共8兲, 726–735.
cone penetrometer in sand.” Geotechnique, 38共1兲, 39–44. Salgado, R., Mitchell, J. K., and Jamiolkowski, M. 共1997b兲. “Cavity Ex-
Houlsby, G. T., and Wroth, C. P. 共1989兲. “The influence of soil stiffness pansion and Penetration Resistance in Sand.” J. Geotech. Geoenviron.
and lateral stress on the results of in situ tests.” Proc., XII ICSMFE, Eng., 123共4兲, 344–354.
Rio de Janeiro, Vol. 1, 227–232. Salgado, R., Mitchell, J. K., and Jamiolkowski, M. B. 共1998兲. “Calibra-
Ishibashi, I., and Zhang, X. 共1993兲. “Unified dynamic shear moduli and tion chamber size effects on penetration resistance in sand.” J. Geo-
damping ratios of sand and clay.” Soils Found., 33共1兲, 182–191. tech. Geoenviron. Eng., 124共9兲, 878–888.
Jamiolkowki, M., Ladd, C, Germaine, J., and Lancellotta, R. 共1985兲. Salgado, R., and Randolph, M. F. 共2001兲. “Analysis of cavity expansion
“New developments in field and laboratory testing of soils.” Proc., XI in sand.” Int. J. Geomech., 1共2兲, 175–192.
ICSMFE, San Francisco. Schmertmann, J. 共1976兲. “An updated correlation between relative den-
Ladanyi, B. 共1963兲. “Expansion of a cavity in a saturated clay medium.” sity DR, and Fugro-type electric clone bearing qc.” Contract Rep. No.
J. Soil Mech. and Found. Div. ASCE 90共SM4兲, 127–161. DACW 39-76 M6646. Waterways Experimental Station.
Lee, K. M., Shen, C. K., Leung, D. H. K., and Mitchell, J. K. 共1999兲. Schnaid, F., and Houlsby, G. 共1991兲. “An assessment of chamber size
“Effects of placement method on geotechnical behavior of hydraulic effects in the calibration of in situ tests in sand.” Geotechnique, 41共3兲,
fill sands.” J. Geotech. Geoenviron. Eng., 125共10兲, 832–846. 437–445.
Lo Presti, D. C. F. 共1987兲. Ph.D. thesis, Politécnico di Torino, Turin, Italy. Vésic, A. S. 共1972兲, “Expansion of cavities in infinite soil mass.” J. Soil
Lo Presti, D. C. F., Pedroni, S., and Crippa, V. 共1992兲. “Maximum dry Mech. and Found. Div., 98共3兲, 265–290.
density of cohesionless, soils by pluviation and by ASTM D4253-83: Villet, W. C. B., and Mitchell, J. K. 共1981兲. “Cone resistance, relative
A comparative study.” Geotech. Test. J., 15共2兲, 180–189. density and friction angle.” Proc., ASCE Symp. on Cone Penetration
Mayne, P. W., and Kulhawy, F. H. 共1991兲. “Calibration chamber database Testing and Experience, Norris and Holtz, eds., St. Louis, 178–208.
and boundary effects correction for CPT data.” Proc., 1st Int. Symp. Yu, H. S., and Houlsby, G. T. 共1991兲. “Finite cavity expansion in dilatant
on Calibration Chamber Testing (ISOCCT1), Potsdam, N.Y., A. B. soils: Loading analysis.” Geotechnique 41共2兲, 173–183.

INTERNATIONAL JOURNAL OF GEOMECHANICS © ASCE / JULY/AUGUST 2007 / 265

Int. J. Geomech., 2007, 7(4): 251-265

You might also like