Sie sind auf Seite 1von 11

Lu, Q., Randolph, M. F., Hu, Y. & Bugarski, I. C. (2004). Géotechnique 54, No.

4, 257–267

A numerical study of cone penetration in clay


Q. L U, * M . F. R A N D O L P H , * Y. H U † a n d I . C . B U G A R S K I ‡

This paper presents results from large-displacement Cet exposé présente les résultats d’une analyse d’éléments
finite-element analysis of cone penetration into clay. The finis à grand déplacement lors de la pénétration d’un cône
soil is idealised as a homogeneous elastic-perfectly plastic dans de l’argile. Le sol est idéalisé en tant que matière
material obeying a Tresca yield criterion, and the analysis élastique homogène parfaitement plastique obéissant au
is carried out using an ‘arbitrary Lagrangian–Eulerian’ critère de fluage de Tresca, et l’analyse est faite en utilisant
technique, with periodic remeshing and interpolation of une technique ‘arbitraire Langrangienne-Eulerienne’, avec
all field values. This allows the cone to be advanced by reconnexion et interpolation périodiques de toutes les
several diameters, thus achieving steady-state conditions. valeurs de terrain. Ceci permet d’avancer le cône de
A full parametric study has been undertaken, quantifying plusieurs diamètres, obtenant ainsi des conditions d’état
the influences of the rigidity index, in situ stress anisotro- stable. Nous avons effectué une étude paramétrique com-
py and the cone roughness. A theoretical correlation for plète, quantifiant les influences de l’indice de rigidité, de
the cone factor, Nkt , is developed from this study, and l’anisotropie de contrainte in situ et de la rugosité du cône.
compared with previous correlations developed using the Nous développons une corrélation théorique pour le fac-
strain path method. Characteristics of the stress distribu- teur cône, N kt , à partir de cette étude et nous la comparons
tion around the cone, the extent of the plastic zone and avec les corrélations précédentes développées en utilisant
apparent incremental movements are discussed, allowing la méthode de cheminement de déformation. Nous discu-
new insights into this problem. tons des caractéristiques de la distribution de contrainte
autour du cône, de l’étendue de la zone plastique et des
mouvements incrémentiels apparents, ce qui donne une
KEYWORDS: clays; plasticity nouvelle façon d’envisager ce problème.

INTRODUCTION range 9–17, with the shear strength measured in triaxial


The cone penetration test is one of the most widely used in compression generally used to normalise the cone resistance.
situ tests for assessing the strength profile in soils. In In principle, comparison of the empirically derived cone
saturated clays and other fine-grained soils, the test is carried factor with the theoretical values for idealised soil provides
out at a penetration rate that does not permit drainage, and additional information on the soil characteristics. Indeed, it
the cone resistance may then be interpreted directly as a may also be appropriate to adjust theoretical geotechnical
measure of the undrained shear strength of the clay. Conven- design equations in a manner that is consistent with any
tionally, the shear strength is derived by dividing the (net) difference between empirical and theoretical cone factors.
cone resistance by a cone factor and, ideally, it would be The current theoretical basis for cone factors is discussed
helpful to have a sound theoretical basis for this cone factor. later, but arguably the most rigorous approach to date is
Real soils render this task difficult because of complex based on the strain path method (Baligh, 1985), with small-
rheological characteristics, where the shear strength is a displacement finite-element analysis used to establish final
function of the rate of strain, the particular stress (or strain) equilibrium stress fields (Teh & Houlsby, 1991). Conven-
path imposed, and other factors such as the physical struc- tional small-strain finite-element analysis alone is not suita-
ture of the deposit. The same limitations apply to any type ble for determining cone factors, as large displacements are
of bearing capacity calculation, and yet much of geotechni- necessary to establish a steady-state (residual) stress field.
cal engineering design is based on theoretical plasticity Although a Eulerian finite-element approach to cone penetra-
solutions that are derived from idealised models of the soil tion problems was outlined by van den Berg (1994), it does
response. Logically, therefore, a starting-point for interpreta- not appear to have been pursued, perhaps because of pre-
tion of cone resistance is to establish theoretical cone factors vious limitations in computing power.
for similarly idealised soil models. This paper presents results of a parametric study under-
In practical applications, it may still be appropriate to taken using large-displacement finite-element analysis, in
establish an empirical cone factor, for example by correlat- order to evaluate steady-state cone penetration resistance. The
ing the cone resistance with shear strengths determined in large-displacement approach follows the RITSS (Remeshing
the laboratory or with an alternative in situ device. Typical and Interpolation Technique combined with Small Strain)
ranges of cone factor, Nkt (see later, equation (3)) may be finite-element approach (Hu & Randolph, 1998a). Conven-
found in the literature (Lunne et al., 1997), and can range tional small-strain finite-element analysis is combined with
from as low as 6 to over 20, but more commonly are in the frequent remeshing and interpolation of all field values, such
as stresses and material properties. The cone is advanced by
Manuscript received 23 July 2002; revised manuscript accepted 2 several diameters until steady-state conditions are achieved.
February 2004. The soil is modelled as a homogeneous elastic-perfectly
Discussion on this paper closes 1 November 2004, for further plastic material obeying the Tresca yield criterion, and with
details see p. ii. low compressibility. The influences of cone roughness, Æc ,
* Centre for Offshore Foundation Systems, The University of
Western Australia.
soil rigidity index, Ir , and initial stress anisotropy, ˜, on the
† School of Engineering, Curtin University of Technology, Aus- cone resistance have been quantified, leading to an expression
tralia. for the cone factor that allows for these effects. The extent of
‡ Department of Civil Engineering and Surveying, Darmstadt the plasticity zone, stress distributions around the cone and
University of Technology, Germany. apparent incremental movements are also explored.

257

Downloaded by [ Universidad de Concepcion] on [02/01/18]. Copyright © ICE Publishing, all rights reserved.
258 LU, RANDOLPH, HU AND BUGARSKI
 
PREVIOUS STUDIES AND RESULTS 4 G 4
Teh & Houlsby (1991) have reviewed previous semi- ł s ¼ su 1 þ ln þ p0 ¼ su ð1 þ ln I r Þ þ p0 (2)
3 su 3
analytical approaches that have been used to estimate cone
resistance. They comment that conventional approaches for where G is the soil shear modulus; Ir is the rigidity index,
the bearing capacity of shallow objects are inappropriate for G/su ; and p0 is the initial mean stress. Expressing the cone
deep penetration. The transition from shallow failure, where resistance in terms of the conventional cone factor, Nkt ,
the mechanism extends to the free surface, to deep failure, relating the net cone resistance to the shear strength (Lunne
where the volume of the penetrometer is accommodated by et al., 1997):
elastic deformation of the soil, occurs at penetrations as low qc ¼ Nkt su þ  v0 (3)
as 1–2 diameters in homogeneous soil (Hu et al., 1999) and
within 3–4 diameters for soil where the strength increases where v0 is the in situ vertical stress (taken as equivalent to
proportionally with depth. p0 in the cavity expansion analogy), the cone factor is then
given by
pffiffiffi
Nkt ¼ 1:33 þ 1:33 ln I r þ 3Æc (4)

Cavity expansion For typical clays with rigidity index, Ir , in the range 50–
The analogy between deep penetration and cavity expan- 500, this expression gives a range for Nkt from 6.55 to
sion draws on the similarities of an inner plastic region with 11.35. As will be shown later, this range is somewhat lower
outward motion of the soil accommodated by elastic defor- than that obtained from more rigorous numerical analysis
mations in the far field. While the eventual displacement that takes proper account of the actual soil deformation
field surrounding the shaft of a penetrometer is closely during steady-state cone penetration.
analogous to cylindrical cavity expansion (Steenfelt et al.,
1981), the soil displacements in front of the advancing
penetrometer tip may be considered closer to those for Strain path method
spherical cavity expansion. The limiting pressure for spheri- The strain path method first proposed by Baligh (1985)
cal cavity expansion has formed the basis for estimating pile has been used extensively as a basis for estimating the cone
tip resistance at deep penetrations, a similar problem to cone resistance. The soil is treated as a viscous fluid, and a flow
resistance (Vesic, 1972). field is established from a potential function, in order to
A full review of cavity expansion theory and its various satisfy incompressibility. The strain rates may then be
applications has been provided by Yu (2000). The high derived for each soil element by differentiating the velocity
symmetry of cavity expansion allows relatively simple analy- field. Corresponding deviatoric stresses are derived by inte-
tical solutions to be obtained, which is the principal attrac- grating the relevant constitutive model along the streamlines,
tion of its use for deep penetration problems in spite of their and the mean stress is determined from either one of the
evident asymmetry. The relationship between the spherical equilibrium equations (radial or axial). However, in general
cavity expansion limit pressure and cone resistance is gen- the deduced stresses will not satisfy the other equilibrium
erally based on an approach such as that shown in Fig. 1 equation, with the discrepancy reflecting the error in the
(Ladanyi & Johnson, 1974). The cone tip is replaced by a initial flow field.
hemispherical surface with the same diameter as the cone, Teh & Houlsby (1991) have discussed a number of meth-
within which the cavity expansion limit pressure, łs , is ods to establish equilibrium following the initial estimate of
assumed to act. Additionally, shear stress, ı, taken as Æc su stresses using the strain path method. They concluded that
(where Æc represents a friction ratio for the cone–soil inter- the most effective approach was to take the estimated stress
face, and su is the soil shear strength) is assumed to act on field as initial stresses in a small-strain finite-element analy-
the face of the cone. Thus for a 608 cone the penetration sis, and establish equilibrium during further advance of the
resistance is expressed as cone (typically by 20% of the cone radius). This approach
will not necessarily establish the correct residual stress field
pffiffiffi behind the advancing cone.
qc ¼ łs þ 3Æc su (1) A further limitation, intrinsic to the strain path method, is
that the cone–soil interface is assumed to be smooth. Teh &
If the soil obeys a Tresca yield criterion, łs is given by Houlsby (1991) addressed this in two ways. First, using the
(Vesic, 1972) (uncorrected) strain path method and estimating the effect of
cone tip roughness using the quasi-analytical approach dis-
cussed in relation to Fig. 1, they arrived at
Nkt ¼ 1:25 þ 1:84 ln I r þ 2Æc  2˜ (5)
where ˜ is the in situ deviator stress normalised by the
shear strength, ˜ ¼ (v0  h0 )/2su . The coefficient 2 for the
D
cone tip roughness, rather than ˇ3 in equation (4), is due to
the adoption of a von Mises failure criterion for the soil,
with su being the strength in triaxial compression.
Their second method was a full strain path and finite-
q
element analysis, with the effect of cone tip and shaft
roughness (friction ratios, Æc and Æs ) estimated from finite-
element analysis conducted on pre-installed cones. The
τ τ resulting expression for the cone factor was
 
: Ir
ψs ψs Nkt ¼ 1 67 þ ð1 þ ln I r Þ þ 2:4Æc  0:2Æs  1:8˜
1500
(6)
Fig. 1. Transformation of cavity expansion results to cone
penetration Comparing equations (5) and (6), the latter equation gives

Downloaded by [ Universidad de Concepcion] on [02/01/18]. Copyright © ICE Publishing, all rights reserved.
A NUMERICAL STUDY OF CONE PENETRATION IN CLAY 259
higher values of Nkt for values of rigidity index greater than Analysis details
100, and both expressions give higher values than from the A standard cone with tip angle of 608 has been consid-
spherical cavity analogue. ered. Although the analysis could, in principle, be carried
out with penetration of the cone from the free surface, as
the main objective is to evaluate the cone resistance at
depth, the cone was pre-embedded to a depth of 5 cone
LARGE-DISPLACEMENT FINITE-ELEMENT ANALYSIS
diameters (D). This is below the depth at which the free
As pointed out by Houlsby et al. (1985), conventional
surface is likely to affect the penetration resistance. All
small-strain finite-element analyses (e.g. Griffiths, 1982; De
distances and penetrations of the cone have been normalised
Borst & Vermeer, 1984) are unable to generate the necessary
by the cone radius (R) or diameter (D), and the results are
residual stress field around the cone, and thus to achieve an
independent of the absolute size of the cone. Fig. 2 shows
appropriate ultimate cone resistance. In order to simulate
an example of an initial mesh and boundary conditions used
sufficient displacement, a large-deformation model is essen-
in the analysis. Note that the interface between cone and soil
tial. An updated Lagrangian approach, as used by Kiousis et
was modelled in such a way as to prevent any soil move-
al. (1988), couples the element nodes in the mesh to the
ment normal to the interface, but allowing sliding motion at
material points, with the mesh being updated after every
a predetermined limiting shear stress.
increment. However, as concluded by van den Berg &
The size of the finite-element domain was chosen to
Vermeer (1988), during continuous deformation the elements
ensure that the boundaries were well outside the plastic
become severely distorted in shape, leading to numerical
zone. In order to optimise computation times, the radial
problems in the solution. An alternative approach for large-
extent of the domain was varied with the rigidity index, Ir ,
deformation analysis is a Eulerian scheme, where the ele-
noting that solutions for cavity expansion from zero initial
ment mesh is fixed and the material flows through the
radius show that the ratio of plasticpffiffiffiffi radius to the current
elements (van den Berg, 1994). In this approach, extra terms
cavity radius remains
p ffiffiffiffi constant at I r for cylindrical cavity
are included in the governing equations to account for the
expansion and 3 I r for spherical cavity expansion (Vesic,
rotation of the material and convection of the stress field,
1972). The radial extent
pffiffiffiffi of the domain, AB, was therefore
which can lead to mathematical difficulties when complex
taken as about 1:5 I r R (where R is the cone radius), and
constitutive models are adopted.
the overall depth of the domain, AH, was fixed at 45R
An alternative ‘steady-state’ finite-element approach has
(22.5D) except for the check analysis for cone penetration
been presented by Yu et al. (2000), in which the stress–
of 25 diameters discussed later. No attempt was made to
strain history of elements flowing past a fixed cone is
optimise the domain further, by reducing either dimension,
essentially integrated along the flow path. The analysis
but Fig. 3 indicates that increasing the radial extent of the
includes some approximations, such as setting zero nodal
domain gave no difference in the penetration response. The
force at the very tip of the cone, and also integrating the
two analyses shown are for Ir ¼ 150, so that the criterion
strain history along vertical columns of elements, which
above would give a domain radius of 20R ¼ 10D (where
appears not to take full account of the effects of radial
R is the radius of the cone).
motion. Their expression for the cone factor was
All analyses reported here are based on a simple elastic,
 perfectly plastic soil model, with a Tresca failure criterion.
Nkt ¼ 0:33 þ 2 ln I r þ 2:37  1:83˜ (7)
cs This model gives identical shear strength, su , irrespective of
the strain path; in contrast, a von Mises model, such as that
where  and cs are the interface friction angle (over cone adopted in the work of Teh & Houlsby (1991), gives a shear
tip and shaft) and critical state friction angle for the soil strength in triaxial compression, which is 15% smaller than
respectively. for plane strain conditions (simple shear). Plasticity analyses
of axisymmetric or three-dimensional problems reported in
the literature generally show that similar normalised capa-
Combined remeshing and small-strain analysis cities are obtained from the two different models (within a
The form of analysis undertaken here is referred to as few per cent), depending on which shear strength is used as
arbitrary Lagrangian–Eulerian (ALE), whereby a series of the normalising quantity (e.g. Murff & Hamilton, 1993).
small-strain analysis increments are followed by complete Poisson’s ratio was taken as 0.49, and the rigidity index was
remeshing, and then interpolation of the field quantities varied between 50 and 500. Nodal force interface elements
(stress and material properties) between the Gauss points in
the new mesh and those in the old mesh. The coordinate R (D/2)
positions of the Gauss point in the old mesh are first
updated, according to the cumulative displacements in the
preceding increments of analysis. This process was proposed
5D

by Hu & Randolph (1998a) and referred to as RITSS


(Remeshing and Interpolation Technique combined with A
Small Strain).
The sequence of small-strain increments, remeshing and
interpolation is repeated until sufficient penetration of the
AH

cone has been achieved, which may be several diameters. As


the remeshing and interpolation aspects are essentially in-
dependent of the incremental analysis, the procedure may be
linked with virtually any finite-element code, and complex
constitutive models present no particular difficulty. Detailed
discussion of strategies for the size of increments and A
frequency of remeshing may be found in Hu & Randolph
(1998a, b), Hu et al. (1999) and Lu et al. (2000, 2001), AB
where the robustness of the method has also been demon-
strated. Fig. 2. Initial mesh for analysis

Downloaded by [ Universidad de Concepcion] on [02/01/18]. Copyright © ICE Publishing, all rights reserved.
260 LU, RANDOLPH, HU AND BUGARSKI
14 14
1
12 2 12
11·3
10 10

9·3
8 8

q/su
AH ⫽ 22·5D
q/su

6 1: AB ⫽ 20D 6 Small strain


2: AB ⫽ 10D
4 Large displacement
4

2
2
0
0 0 1 2 3 4 5 6 7
0 2 4 6 8 Normalised penetration: diameters
Normalised penetration: diameters
Fig. 4. Identification of steady state (Ir 150, smooth cone,
Fig. 3. Effect of domain width (Ir 150, rough cone, smooth smooth shaft)
shaft)
approach is to take the point at which the incremental
(e.g. Herrmann, 1978; Lu et al., 2000) were used between stiffness has reduced by a certain factor compared with the
the cone and the soil, with the nodal forces representing the initial stiffness. Although this is sufficient from a practical
integrated shear and normal stresses acting on the interface point of view for certain applications, the definition is at
area represented by each node. The interface elements best arbitrary, and would lead to significant underestimation
allowed the limiting interfacial shear stress to be varied from of cone resistance.
zero to 100% of the soil shear strength. Figure 4 shows a comparison of conventional small-strain
Except where otherwise stated, all results presented are and large-displacement (RITSS) analysis of cone penetration.
for an additional cone penetration of 6–7 diameters (a total This analysis is for a smooth cone tip (compared with a
of 11–12 diameters below the soil surface), in order to rough cone tip in Fig. 3), rigidity index, Ir ¼ 150, and
achieve steady-state conditions. A check analysis was also isotropic initial stresses. The small-strain analysis has been
carried out where the cone was penetrated 25 diameters continued until a displacement of 1 cone diameter. However,
from its initial position. This gave a very slight increase in the accuracy of the small-strain analyses becomes doubtful
cone resistance (some 3% higher) compared with a penetra- for displacements beyond 0.5D, owing to gross distortion of
tion of 6 diameters. the mesh, as indicated by the divergence of the two curves.
As discussed later, the cone resistance reported here is the The large-displacement analysis indicates that the limiting
force on the cone tip itself, divided by the projected cross- resistance is not achieved until a penetration of approxi-
sectional area of the cone. In practice, field cones include a mately 5–6D, beyond which the normalised penetration
small part of the shaft (7–10 mm in a standard 36 mm resistance is reasonably constant at 11.35. This is 22%
diameter cone) in the load cell measurement (ISSMGE, greater than the resistance of 9.3 at a displacement of 0.5D,
1999). For typical friction ratios of 3–4%, this would the practical limit for conventional small-strain analysis. An
amount to an increase in the cone factor by a similar amount analysis was also undertaken for a rough cone, using a more
(3–4%). extensive but slightly coarser mesh, where the cone was
pushed 25 diameters beyond its pre-embedment position.
That analysis gave a 3% increase in penetration resistance
RESULTS AND DISCUSSION beyond 6 diameters.
Identification of steady state Stress contours (with r , v representing radial and verti-
As pointed out previously, conventional small-strain finite- cal stress respectively) from the conventional small-strain
element analysis has difficulty in identifying the ultimate analysis are compared with those from the large displace-
bearing capacity for deep penetration problems. A common ment analysis in Fig. 5 for a smooth cone (as analysed in

O O
1 7 7 7
3 5 5 1 1 7
3 5 5
3 1
3

Small strain Large displacement Small strain Large displacement

(a) (b)

Fig. 5. Stress distributions at penetration of 0.5D (Ir 150, smooth cone, smooth shaft):
(a) contours of r /su ; (b) contours of v /su

Downloaded by [ Universidad de Concepcion] on [02/01/18]. Copyright © ICE Publishing, all rights reserved.
A NUMERICAL STUDY OF CONE PENETRATION IN CLAY 261
Fig. 4), for a cone displacement of 0.5D. The two sets of residual stress field created behind the advancing cone,
contours are very similar, confirming that the RITSS ap- which is similar to that obtained by Teh (1987), and Teh &
proach does not distort the stress contours, although the Houlsby (1991).
interpolation procedure has led to a smoothing of the However, the stress contours revealed here show more
contours compared with those from the small-strain analysis. gradual reduction in stresses below the cone. The stress
contours in Figs 5 and 6 indicate a centre, O, about one
radius of the cone below the shoulder. At the level of this
centre, the radial stress changes are larger than the vertical
Stress distributions around cone
stress, whereas ahead of the cone the vertical stress changes
The evolution of the stress distributions around a rough
. are larger. Note that the slight concavity seen in the two
cone as the displacement increases from 0 5D to 6D and
lowest vertical stress contours corresponds to a transition
beyond is shown in Fig. 6, for the same soil properties as
from downward vertical displacement below the cone to
for Fig. 3. Although the patterns of stress contours at the
upward vertical displacement, as indicated later in Fig. 18.
different displacements are generally similar, the stress levels
The analogy between cone penetration and spherical cav-
increase significantly with displacement. The increase in
ity expansion is illustrated well in Fig. 7(a), where the plot
radial and vertical stresses adjacent to the cone between
shows the direction of the major principal stress changes, 1
penetrations of 0.5–6D may be estimated approximately
and 3 , and the plot in Fig. 7(b) shows magnitudes of 1 /su
from the closest contours (7 7su to 9 5su for r , 9su to 11su
. .
along different lines radiating (approximately) from a no-
for z as the cone displacement increases from 0.5D to 6D);
tional expansion centre, O, midway between cone tip and
the increase is 22–23%, which is broadly consistent with the
. . shoulder. The values of 1 /su vary with the angle, Ł, near
change in the cone resistance from 10 7 to 12 7 (19%)
the cone but then converge with increasing distance. Inter-
between these two displacements (Fig. 3).
estingly, however, the rate of decay is much more gradual
For comparison, stress contours for a displacement of 25D
than the theoretical curve from spherical cavity expansion
are also shown in the right-hand half of each diagram, with
(Vesic, 1972).
successive contours always lying just outside the correspond-
ing contours for a displacement of 6D. It is interesting to
note that, although the cone resistance increased by only 3%
between penetrations of 6D and 25D, significant changes Extent of plastic zone
occurred in the stress contours, particularly at some distance During deep penetration, the plastic zone is confined by
from the cone. The most notable difference in the stress elastically deforming soil. The extent of the plastic zone will
contours for displacements of 6D and 25D is for the lowest- depend on various factors, such as cone roughness, soil
level contours (r /su and v /su equal to 1), where the analy- rigidity and initial stress anisotropy. These effects are exam-
sis to 25D displacement gives a more realistic picture of the ined in this section.

1 1

1
1

9·5 11
7·7 3
1 3 9
1
3 3 3
3
5 7 9
7 5 5
5 7 9
7

0·5D 6D
0·5D 6D

25D
25D

(a) (b)

Fig. 6. Stress distributions at different penetrations (Ir 150, rough cone, smooth shaft): (a)
contours of r /su ; (b) contours of v /su

Downloaded by [ Universidad de Concepcion] on [02/01/18]. Copyright © ICE Publishing, all rights reserved.
262 LU, RANDOLPH, HU AND BUGARSKI

R O

Zp
(a) Rp

12

α ⫽ 0° (a) (b)
10
α ⫽ 45°
Fig. 8. Current plastic zones during steady state for different
8 α ⫽ 90°
soil rigidity indices: (a) Ir 150, rough cone, smooth shaft; (b)
Spherical cavity O
d Ir 300, rough cone, smooth shaft
expansion
σ1/su

6
α
20
4

2 Cylindrical cavity
15
Zp /R (RITSS)
0
Zp /R, Rp /R

0 2 4 6 8 10 12
d/R 10 Rp /R (RITSS)
(b)

Fig. 7. Principal stresses around a penetrating cone (Ir 150,


5 Spherical cavity
smooth cone, smooth shaft): (a) orientation of principal stresses;
(b) major principal stress variation with radial distance from O

0
0 200 400 600
Figure 8 shows the plastic zones for rough cones (Æc ¼ 1) Soil rigidity index, Ir
with smooth shafts (Æs ¼ 0) penetrating into homogeneous
soils with isotropic initial stresses. The results are for cone Fig. 9. Variation of plastic zone with soil rigidity index
displacements of 6D, although the size was found not to
change significantly for the analysis taken to 25D. The size
of the plastic zone in the vicinity of the cone tip may be
characterised approximately by radius Rp and depth Zp, as
indicated in Fig. 8, and the evident dependence of the size
on the soil rigidity index is quantified in Fig. 9. The plastic
radii for cylindrical and spherical cavity expansion solutions
(Vesic, 1972) are plotted for comparison. It may be seen that
the general trend of the variation obtained from the finite-
element analyses agrees well with the solution from spheri-
cal cavity theory, especially in terms of the radial extents,
which are only marginally above the spherical cavity solu-
tion.
In the studies by Teh & Houlsby (1991) and Yu et al.
(2000), the plastic zone was found to extend back along the
shaft, with the boundary of the plastic zone above the cone
shoulder being almost parallel to the shaft. The horizontal
extent of the plastic zone in that region was found to
correlate well with the plastic radius from cylindrical cavity
expansion. Although, in the current study, the plastic zone
does not appear to extend as far radially, as noted in Fig. 9,
this is due to slight unloading behind the cone, limiting the (a) (b)
extent of soil that is currently plastic.
Figure 10 indicates little influence of roughness of the Fig. 10. Effect of roughness on the extent of current plastic
cone tip and shaft on the development of the plastic zone. zone (Ir 150): (a) smooth cone, smooth shaft; (b) rough cone,
Fig. 11 together with Fig. 8(a) shows that anisotropic initial rough shaft

Downloaded by [ Universidad de Concepcion] on [02/01/18]. Copyright © ICE Publishing, all rights reserved.
A NUMERICAL STUDY OF CONE PENETRATION IN CLAY 263
area of the embedded shaft of the cone. As given in equa-
tion (3), the cone factor is then derived as the ratio of the
ultimate net cone resistance to the shear strength of the soil:
qc   v0
Nkt ¼ (9)
su
As noted previously, the cone factor defined in this way will
be a slight underestimate (for Æs . 0) for the standard cone
geometry defined in ISSMGE (1999), as the short length of
shaft above the conical tip has been ignored. Analyses
undertaken with different (isotropic) initial stress levels gave
identical cone factors, regardless of the level of initial stress.
It was also found that the cone tip resistance was indepen-
dent of the amount of shaft friction, once corrected accord-
ing to equation (8), as indicated in Fig. 13.

Effect of soil rigidity index. It has been widely recognised


that the cone factor must vary with soil rigidity, owing to the
accommodation of the cone volume by ‘elastic’ strains in the
(a) (b) soil outside the plastic zone. To quantify the effect of soil
rigidity on the cone factor, two steps have been taken. First,
Fig. 11. Effect of stress anisotropy on the extent of current calculations have been carried out for five different soil
plastic zone (Ir 150, rough cone, smooth shaft): (a) ˜ 0.5;
(b) ˜ 0.5
rigidity indices assuming isotropic initial stresses and a fully
rough cone tip (but smooth shaft). Results from these
analyses are summarised in Fig. 14. Second, analyses have
stresses, quantified by ˜ ¼ (vo  ho )/2su , seems to result been carried out for a range of initial stress ratios and values
in a slight change in the plastic domain. For a more of cone tip roughness to validate the trends from the first set
complete picture, the sizes of plastic zone for different of analyses.
combinations of soil rigidity and stress anisotropy are com- Figure 14 shows that the trend derived here using the
pared in Fig. 12. It is interesting to note that, for higher soil RITSS finite-element approach agrees well with the solution
rigidity indices, the case of isotropic initial stresses tends to
produce the largest vertical extent of the plastic zone and 60
the smallest radial extent. Rough cone, rough
shaft (F/Asu)
50

Cone factor
(F–f )/A/su or F/A/su

40
The cone resistance is computed from the displacement-
controlled finite-element analyses by taking the total force
30
on the cone, F, which is the sum of the vertical components
of all nodal forces along the interface, subtracting any shaft
friction, f, and dividing by the cross-sectional area of the 20
cone, A, to give Rough cone, smooth shaft F/A/su

F f 10 Rough cone, rough shaft (F–f )/A/su


qc ¼ (8)
A
For most of the results presented here, the cone shaft has 0
0 1 2 3 4 5 6 7
been modelled as a smooth interface, so that f is zero. In
Normalised penetration: diameters
other cases, a limiting shaft friction of Æs su has been
modelled, so that f is the product of Æs su times the surface Fig. 13. Effect of shaft roughness, Æs

14
18
∆ Zp /R Rp /R
12 16
0·5 SPFE
RITSS
14
0·0
Zp /R, Rp /R

10 ⫺0·5 12
10 SP
Nkt

8 8
6
6 4
2
4 0
0 100 200 300 400 500 600 0 100 200 300 400 500 600
Soil rigidity index, Ir Soil rigidity index, Ir

Fig. 12. Variation of plastic zone with soil rigidity index and Fig. 14. Variation of Nkt with soil rigidity index (isotropic initial
stress anisotropy (rough cone, smooth shaft) stresses)

Downloaded by [ Universidad de Concepcion] on [02/01/18]. Copyright © ICE Publishing, all rights reserved.
264 LU, RANDOLPH, HU AND BUGARSKI
obtained by Teh & Houlsby (1991), using a pure strain path Values of Nkt from the first set of analyses fall approxi-
(SP) method. The discrepancy in magnitude between these mately on a straight line with a gradient of 1.82, as shown
two solutions may be attributed mainly to the effect of cone by line 2 in Fig. 16. This gradient is almost exactly the
roughness, although the gradient of Nkt against ln Ir from the same as that obtained by Teh & Houlsby (1991) from their
current analyses is slightly lower than that from the strain finite-element adjusted strain path analyses (necessarily for a
path method (1.59 instead of 1.84, a reduction of 14%). The rough cone), although it differs by 10% from the coefficient
difference between the current results and those from Teh & of 2.0 obtained directly from the strain path method
Houlsby (1991) for a rough cone, after using small-strain (necessarily for a smooth cone). The latter coefficient may
finite-element analyses to correct equilibrium errors (see be argued theoretically for the particular boundary condi-
equation (6), plotted as SPFE in Fig. 14), is significant, tions assumed in the strain path method (uniform vertical
especially for high values of rigidity index. For the case of a soil velocity relative to the cone at large z), but seems not to
fully rough cone (Æc ¼ 1) in initially isotropic soil, the hold once small elastic deformations are allowed for at
relationship shown in Fig. 14 is depth. As pointed out by Teh & Houlsby (1991), though, the
Nkt ¼ 4:62 þ 1:59 ln I r (10) net effect is that the total cone resistance is much more
strongly associated with the in situ horizontal stress than the
Results from the additional analyses to check the effect of vertical stress.
cone roughness and initial stress ratio on the dependence of The dependence of the cone factor on the initial stress
Nkt on rigidity index are presented in Fig. 15. The variation state can also be explained by the induced stress distribu-
of the gradient of the best-fitting lines shown is very slight, tions shown in Fig. 17 (at displacements of 6D). It can be
less than 5%, and so for simplicity the average gradient of seen that higher initial horizontal stress (lower ˜) results in
1.6 is recommended for all combinations. slightly lower horizontal stress changes and much higher
vertical stress changes, which in turn result in a higher cone
factor. There are also discernible differences in the incre-
Effect of in situ stress. A similar strategy as in the last mental movements. These have been obtained using software
section has been followed to investigate the effect of in situ from fluid mechanics, where flow-lines are fitted through
stress anisotropy. First, a fixed soil rigidity index of 150 and current directions of incremental movement. As shown in
fully rough cone surface were adopted, while the initial stress Fig. 18, the soil appears to flow more horizontally for high
anisotropy parameter, ˜ ¼ (vo  ho )/2su , was varied be- values of ˜.
tween 0.5 and 0.5. Then, the variation of Nkt with stress Figure 16 demonstrates that the gradient of Nkt with ˜ is
anisotropy was examined for various combinations of soil hardly affected by soil rigidity or cone roughness. The
rigidity index and cone roughness. All results are plotted in maximum error that might be attributed to any difference in
Fig. 16, including two additional analyses where the range of gradient is about 0.13 for the extreme values of ˜ ¼ 1,
˜ was extended to 0.75, confirming a linear trend. which is less than 1.3% of the cone factor. The gradient
may therefore be taken as the average value, 1.9.
16
Effect of cone roughness. In many of the previous studies,
it was impossible to control the shearing boundary condition
14
on the cone face or on the surface of the shaft. In the current
analyses, nodal joint elements have been adopted at the
interface to achieve this control. The formulation of this type
Nkt

12 1
1: ∆ ⫽ ⫺0·5, αc ⫽ 1·0 of interface element follows that proposed by Herrmann
4
2 2: ∆ ⫽ 0, αc ⫽ 1·0
(1978), and the efficacy has been examined by Lu et al.
10 3
(2000) for penetration of cylindrical and spherical penet-
3: ∆ ⫽ 0·5, αc ⫽ 1·0 rometers. The roughness factors Æc (on the cone face) and Æs
4: ∆ ⫽ ⫺0·5, αc ⫽ 0·5 (on the shaft) have been varied between 0 and 1. As
8 discussed previously, the roughness of the cone shaft makes
10 100 1000 no difference to the actual tip resistance, although it clearly
Soil rigidity index, Ir affects the total cone resistance (see Fig. 13). Therefore
attention here is focused on the roughness of the cone tip.
Fig. 15. Variation of Nkt with soil rigidity index Note that, by analogy with typical values of Æ deduced
during continuous penetration of piles and measured on cone
friction sleeves, field values of Æ are likely to lie in the range
1: Ir ⫽ 500, αc ⫽ 1·0
0.2–0.6.
20 2: Ir ⫽ 150, αc ⫽ 1·0 Figure 19 suggests that it is sufficient to adopt a linear
3: Ir ⫽ 50, αc ⫽ 1·0 variation in the cone factor with cone roughness, even
18
though there is a slight curvature in the results for Æc ¼ 0,
4: Ir ⫽ 150, αc ⫽ 0·5
0.5 and 1. The best-fitting lines for different initial stress
16
1 5: Ir ⫽ 150, αc ⫽ 0·0 anisotropy are almost parallel, with a gradient of about 1.3.
Nkt

14 2
4
Summary cone factor relationship. The final relationship
12
5 for the cone factor may be obtained by combining the results
3
discussed in the previous sections, to arrive at
10 Nkt  3:4 þ 1:6 ln I r  1:9˜ þ 1:3Æc (11)
The accuracy of this expression has been assessed against
8
⫺1 ⫺0·5 0 0·5 1 each case of calculation involved in this study by comparing
∆ the back-calculated undrained shear strength according to
equation (9) with the value assumed. The maximum error is
Fig. 16. Variation of Nkt with stress anisotropy 1.7% (for 95% of the cases, the error is between plus and

Downloaded by [ Universidad de Concepcion] on [02/01/18]. Copyright © ICE Publishing, all rights reserved.
A NUMERICAL STUDY OF CONE PENETRATION IN CLAY 265

7 5 7 5 3 1 7 5 3 1
3
1

∆ ⫽ ⫺0·5 ∆⫽0 ∆ ⫽ 0·5

(a)

7
7
7
5
5 5
3
3
3
1
1 1

∆ ⫽ ⫺0·5 ∆⫽0 ∆ ⫽ 0·5

(b)

Fig. 17. Effect of stress anisotropy on stress changes due to penetration: (a) contours of radial stress
change, (r 2 h0 )/su ; (b) contours of vertical stress change, (v 2 v0 )/su

Table 1. Comparison of cone factors derived using different methods


Ir Cavity expansion Strain path Strain path and Large strain/displacement finite
finite element element.
Ladanyi & Baligh Yu Teh & Houlsby Van den Berg RITSS (this paper)
Johnson
S R S R S R S R S R S R S R
50 6.5 8.5 — 15.9 8.5 13.8 8.5 (10.5) (8.4) 10.8 9.5 13.8 9.6 10.9
150 8.0 10.0 — 17.0 9.8 15.0 10.5 (12.5) (10.6) 13.0 — — 11.3 12.7
300 8.9 10.9 — 17.7 10.6 15.8 11.7 (13.7) (12.5) 14.9 — — 12.4 13.8
500 9.6 11.6 — 18.2 11.2 16.4 12.7 (14.7) (14.4) 16.8 12.0 17.5 13.2 14.5

Downloaded by [ Universidad de Concepcion] on [02/01/18]. Copyright © ICE Publishing, all rights reserved.
266 LU, RANDOLPH, HU AND BUGARSKI
is with the solutions for a rough cone expressed in equation
(5) obtained by Teh & Houlsby (1991), using a pure strain
path method but including a quasi-analytical term for the
cone roughness.

CONCLUDING REMARKS
An extensive parametric study has been carried out for
cone penetration in soft clay, using large-displacement finite-
element analysis, with the cone penetrometer being advanced
several diameters in order to achieve a steady state. A
simple elastic perfectly plastic soil model has been adopted
deliberately, with the main aim being to establish a robust
theoretical framework for cone factors, using a soil model
that is consistent with typical geotechnical design. The
∆ ⫽ 0·5 ∆ ⫽ ⫺0·5
large-displacement finite-element scheme is based on the
Fig. 18. Effect of stress anisotropy on incremental movements RITSS approach of Hu & Randolph (1998a), with frequent
remeshing and interpolation of all field values.
The main variables that have been investigated are the
effects of (a) soil rigidity index, (b) in situ stress anisotropy,
and (c) roughness of the cone tip and shaft. Based on the
∆⫽0
results of the analyses, an updated expression for the cone
14 factor in terms of these parameters has been proposed (see
∆ ⫽ 0·5
equation (11)). Detailed stress distributions and incremental
∆ ⫽ ⫺0·5 movements have been investigated, and the findings have
13
been compared with solutions from other approaches, includ-
ing cavity expansion and the strain path method. The cone
factor was found to vary linearly with the logarithm of the
Nkt

12
soil rigidity index, but with a gradient of 1.6, which is
intermediate between the values of 1.33 from spherical
cavity expansion and 1.84 from the strain path approach of
11
Teh & Houlsby (1991).
The analogy of spherical cavity expansion for the cone
penetration resistance is partly supported by the resulting
10
0 0·5 1 stress distributions and the size of the plastic zone around
Cone roughness, αc the cone. In this respect, it is interesting that the radial
extent of the plastic zone appears significantly less than
Fig. 19. Effect of cone roughness, Æc , on cone factor predicted by cylindrical cavity expansion, even where the
cone was advanced 25 diameters through the soil.
minus 1.0%), and the standard error is about 0.6%. There-
fore the approximation in the selections of coefficients ACKNOWLEDGEMENTS
accounting for different factors is practically reasonable. The work described in this paper forms part of the
Equation (11) is applicable to any combination of soil activities of the Special Research Centre for Offshore Foun-
rigidity index, Ir , from 50 to 500, soil anisotropy parameter, dation Systems, established and supported under the Austra-
˜, between 0.5 and 0.5, and cone roughness, Æc , between lia Research Council’s Research Centres programme.
0 and 1. It should be noted that these coefficients are based
on the Tresca soil model adopted in the present analyses,
and would vary slightly if alternative soil models were
NOTATION
adopted. A cross-section area of the cone
Following the parametric study presented above, it is D cone diameter
helpful to compare the resulting values of Nkt with results d distance away from cone
presented from previous studies. Table 1 presents this com- F total force driving the cone
parison, all for isotropic initial stress conditions, denoting f shaft friction
smooth and rough cone tip by S and R respectively. The G soil shear modulus
bracketed solutions are obtained by correcting the basic Ir soil rigidity index, ¼ G=su
solution for the roughness of the cone tip, according to the N kt cone factor
suggestions by the original authors. p0 initial mean stress
q cone resistance
It may be seen that the factors derived from the present
qc limiting cone tip resistance
work are generally higher than the spherical cavity solutions R cone radius
given by equation (4), and are lower than the cavity expan- Rp horizontal extent of the plastic zone
sion solution by Baligh (1975), which is referred to by Yu su soil shear strength
& Mitchell (1998). The values are comparable to the other Zp vertical extent of the plastic zone
solutions listed in Table 1, although there are some outlying Æ angle from vertical at cone tip
results, such as the values for a rough cone from van den Æc cone roughness
Berg (1994). With both cone roughness and soil rigidity Æs shaft roughness
taken into consideration, the range from the present work is  interface friction angle
slightly narrower than the solutions by Yu (1993), Teh & ˜ initial stress anisotropy, ¼ ( v0   h0 )=2su
cs critical state friction angle for the soil
Houlsby (1991) and van den Berg (1994). Closest agreement 1 major principal stress

Downloaded by [ Universidad de Concepcion] on [02/01/18]. Copyright © ICE Publishing, all rights reserved.
A NUMERICAL STUDY OF CONE PENETRATION IN CLAY 267
3 minor principal stress footings and plate anchors embedded in permafrost. Can. Geo-
 v0 initial vertical stress tech. J. 11, 531–553.
 h0 initial horizontal stress Lu, Q., Hu, Y. & Randolph, M. F. (2000). FE analysis for T-bar and
r radial stress ball penetration in cohesive soil. Proc. 10th Int. Offshore and
v vertical stress Polar Engineering Conf. ISOPE 2000, Seattle 2, 617–623.
łs cavity expansion limit pressure Lu, Q., Hu, Y. & Randolph, M. F. (2001). Deep penetration in soft
clay with strength increasing with depth. Proc. 11th Int. Off-
shore and Polar Engineering Conf. ISOPE 2001, Stavanger 2,
REFERENCES 453–458.
Baligh, M. M. (1985). Strain path method. J. Soil Mech. Found. Lunne, T., Robertson, P. K. & Powell, J. J. M. (1997). Cone
Div., ASCE 111, 9, 1180–1186. penetration testing in geotechnical practice. London: Blackie
De Borst, R. & Vermeer, P. A. (1984). Possibilities and limitations Academic and Professional.
of finite elements for limit analysis. Géotechnique 34, 2, 199– Murff, J. D. & Hamilton, J. M. (1993). Pultimate for undrained
210. analysis of laterally loaded piles. J. Geotech. Engng, ASCE 119,
Griffiths, D. V. (1982). Elasto-plastic analysis of deep foundations No. 1, 91–107.
in cohesive soil. Int. J. Numer. Anal. Methods Geomech. 6, Steenfelt, J. S., Randolph, M. F. & Wroth, C. P. (1981). Model tests
211–218. on instrumented piles jacked into clay. Proc. 10th Int. Conf. Soil
Herrmann, L. R. (1978). Finite element analysis of contact pro- Mech. Found. Engng, Stockholm 2, 857–864.
blems. J. Engng Mech., ASCE 104, 1043–1059. Teh, C. (1987). An analytical study of the cone penetration test.
Houlsby, G. T., Wheeler, A. A. & Norbury, J. (1985). Analysis of DPhil thesis, University of Oxford.
undrained cone penetration as a steady flow problem. Proc. 5th Teh, C. I. & Houlsby, G. T. (1991). An analytical study of the cone
Int. Conf. on Numerical Methods in Geomechanics, Nagoya, 4, penetration test in clay. Géotechnique 41, No. 1, 17–34.
1767–1773. van den Berg, P. (1994). Analysis of soil penetration. Delft: Delft
Hu, Y. & Randolph, M. F. (1998a). A practical numerical approach University Press.
for large deformation problem in soil. Int. J. Numer Anal. van den Berg, P. & Vermeer, P. A. (1988). Undrained strength from
Methods Geomech. 22, 5, 327–350. CPT and finite element computations. Proceedings of a Confer-
Hu, Y. & Randolph, M. F. (1998b). Deep penetration of shallow ence on Numerical Methods in Geomechanics, Innsbruck, 1095–
foundations on non-homogeneous soil. Soils Found. 38, 1, 241– 1100. Balkema: Rotterdam: Balkema.
246. Vesic, A. S. (1972). Expansion of cavities in infinite soil mass. J.
Hu, Y., Randolph, M. F. & Watson, P. G. (1999). Bearing response Soil Mech. Found. Div., ASCE 98, No. 3, 265–290.
of skirted foundations on non-homogeneous soil with large Yu, H. S. (1993). Discussion on: Singular plastic fields in steady
deformation. J. Geotech. Engng, ASCE 125, No. 17, 924–935. penetration of a rigid cone. J. Appl. Mech. 60, 1061–1062.
ISSMGE (1999). International reference test procedure for the cone Yu, H. S. (2000). Cavity expansion methods in geomechanics.
penetration test. Report of the ISSMGE Technical Committee 16 Rotterdam: Balkema.
in Proc. 12th Eur. Conf. on Soil Mech. Found. Engng., Amster- Yu, H. S. & Mitchell, J. K. (1998). Analysis of cone resistance:
dam 3, 2195–2222. review of methods. J. Geotech. Geoenviron. Engng, ASCE 124,
Kiousis, P. D., Voyiadjis, G. Z. & Tumay, M. T. (1988). A large No. 2, 140–149.
strain theory and its application in the cone penetration mechan- Yu, H. S., Herrmann, L. R. & Boulanger, R. W. (2000). Analysis of
ism. Int. J. Numer. Anal. Methods Geomech. 12, 45–60. steady cone penetration in clay. J. Geotech. Geoenviron. Engng,
Ladanyi, B. & Johnson, G. H. (1974). Behaviour of circular ASCE 126, No. 7, 594–605.

Downloaded by [ Universidad de Concepcion] on [02/01/18]. Copyright © ICE Publishing, all rights reserved.

Das könnte Ihnen auch gefallen