Sunteți pe pagina 1din 7

Winkler modulus for axially loaded piles

G. MYLONAKI S

The problem of elastic soilpile interaction and its modelling


using the concept of a Winkler support are revisited. It is
shown that dividing the vertical soil shear tractions and the
corresponding settlements along the pile generates depth-
dependent Winkler springs that accurately describe the
pilesoil interaction, contrary to the widespread belief that
the Winkler representation is always approximate. A simpli-
ed theoretical model is then derived for analysing the
response of an end-bearing cylindrical pile in a homoge-
neous soil stratum. Explicit solutions are obtained for: (a)
pile settlement; (b) depth-dependent Winkler modulus; (c)
average (depth-independent) Winkler modulus to match pile
head settlement. Both innitely long piles and piles of nite
length are examined. The approximate analytical solution
compares favourably with nite-element and boundary-ele-
ment solutions. A simple regression formula for the average
Winkler modulus is developed.
KEYWORDS: dynamics; elasticity; piles; settlement; stiffness; soil/
structure interaction.
Nous reprenons le probleme de l'interaction elastique entre
sol et pile et sa representation en maquette par le concept du
support de Winkler. Nous montrons qu'en divisant des trac-
tions de cisaillement de sol verticales et les tassements corre-
spondants le long de la pile, on engendre des ressorts de
Winkler qui dependent de la profondeur, ressorts qui decri-
vent avec exactitude l'interaction pile-sol, contrairement a
l'opinion qui veut que la representation de Winkler soit
toujours approximative. Nous derivons ensuite une maquette
theorique simpliee pour analyser la reaction d'une pile
cylindrique porteuse en bout dans une couche de sol homo-
gene. Nous obtenons des solutions explicites pour : (a) le
tassement de la pile ; (b) le module de Winkler dependant de
la profondeur ; le module moyen de Winkler (ne dependant
pas de la profondeur) pour predire le tassement des tetes de
pile. Nous etudions a la fois les piles de longueur innie et les
piles de longueur nie. La solution analytique approximative
montre une bonne correlation avec les solutions a element
nis et les methodes marginales. Nous developpons une sim-
ple formule de regression pour le module de Winkler moyen.
INTRODUCTION
A versatile way of modelling soilpile interaction is through a
series of independent Winkler springs distributed along the pile
shaft. Although approximate, Winkler models are widely ac-
cepted in the analysis of both axially and laterally loaded piles
subjected to static or dynamic loads (Terzaghi, 1955; Coyle &
Reese, 1966; Novak, 1974; Poulos & Davis, 1980; Randolph &
Wroth, 1978; Scott, 1981; Fleming et al., 1992; Gazetas et al.,
1992). Their popularity stems primarily from their ability to
(a) yield predictions that are in satisfactory agreement with
more rigorous solutions
(b) incorporate variation of soil properties with load amplitude
(non-linearity) and depth (inhomogeneity)
(c) be extended to dynamic loads by adding pertinent
distributed dampers to the spring bed
(d) incorporate group effects through pertinent pile-to-pile
interaction models
(e) require smaller computational effort than rigorous nite-
element or boundary-element formulations.
The key problem in the implementation of Winkler models lies
in the assessment of the modulus of the Winkler springs. In the
case of axially loaded piles, the springs are dened as
k(z)
f (z)
W(z)
(1)
where f (z) denotes the vertical soil reaction per unit pile length
and W(z) the corresponding pile settlement, at depth (z).
Following early applications of the model to settlement analysis
of surface footings, k(z) is often referred to as the modulus of
subgrade reaction and is measured in units of force per length
squared.
y
It is well known that k(z) is not a property of the soil
alone, but depends on the characteristics of both pile and soil,
and varies with depth even in a homogeneous layer.
Under the assumption of negligible Poisson effects in the
pile,
{
introducing the function k(z) reduces the continuum
problem to the one-dimensional equation (Scott, 1981)
E
p
A
p
d
2
W(z)
dz
2
k(z)W(z) 0 (2)
which is amenable to analytical treatment, often leading to
closed-form solutions (Novak, 1974; Scott, 1981; Guo, 2000).
Note that using the exact ratio k(z) f (z)=W(z) in equation (2)
reproduces the pile response accurately, contrary to the wide-
spread view that the Winkler representation is always approx-
imate.
Current methods for determining k(z) can be classied into
three main groups:
(a) experimental methods
(b) calibration with rigorous numerical solutions
(c) simplied theoretical models.
In the experimental methods (group a), k(z) is obtained directly
from equation (1) by measuring the longitudinal strains along
an axially loaded pile and computing analytically the corre-
sponding distributions of vertical soil reaction and pile settle-
ment (Coyle & Reese, 1966). With the methods of group b,
average k values along the pile can be determined by matching
a key response parameter (e.g. pile head settlement) with results
from Winkler models, as for instance done by Thomas (1980)
and Sanchez-Salinero (1982) using nite-element formulations.
Group c consists of approximate analytical methods, notably the
plane-strain model (Novak, 1974; Randolph & Wroth, 1978),
which introduce simplications to derive simple theoretical
estimates of k.
Notwithstanding the signicance of the above methods in
engineering research and practice, they all can be criticised for
certain drawbacks. For instance, experimentally determined k
values (expressed through the well-known `py' or `tz' curves)
have been developed primarily for inelastic conditions, and do
not properly account for the low-strain stiffness of the soil
(Reese & Wang, 1996). On the other hand, calibrations with
455
Mylonakis, G. (2001). Geotechnique 51, No. 5, 455461
Manuscript received 16 November 2000; revised manuscript accepted 16
March 2001.
Discussion on this paper closes 2 November 2001, for further details
see the inside back cover.

Civil Engineering Department, City University of New York.


{ Note that k differs from the so-called coefcient of subgrade reaction
(denoted by k
s
), which is dened as pressure over settlement, and which
thereby has units of force per length cubed.
{ The effect of radial displacements on pile response is usually minor
as, for instance, shown by Mattes (1969) and Pak & Ji (1993).
rigorous numerical solutions in group b may encounter numer-
ical difculties in certain parameter ranges, as, for example, in
the case of long compressible piles (El-Sharnouby & Novak,
1990). Also, these approaches are often limited by the analytical
and computational complexities associated with the underlying
numerical procedures, which can make them unappealing to
geotechnical engineers. Finally, the plane-strain theories of
group c involve empirical parameters that need to be calibrated
with other methods (Randolph & Wroth, 1978), and do not
account for important factors such as the continuity of the
medium in the vertical direction and the stiffness mismatch
between pile and soil.
With reference to the methods in group c, it seems that a
simple rational model capable of providing improved estimates
of k(z) to be used in engineering applications would be
desirable. In the framework of linear elasticity, an approximate
analytical solution is presented in this paper for an axially
loaded pile in a homogeneous soil stratum. While maintaining
analytical simplicity, the proposed solution has distinct advan-
tages over the existing models in group c. Specically
(a) it accounts for the continuity of the medium in both the
horizontal and vertical directions
(b) it accounts for pilesoil stiffness ratio, pile length to
diameter ratio, and compressibility of the soil material
(c) it does not involve empirical constants.
Apart from its intrinsic theoretical interest, the proposed approx-
imate solution may be used to provide a more rational basis for
assessing and improving other related methods.
PROBLEM DEFINITION AND MODEL DEVELOPMENT
The system considered in this study is depicted in Fig. 1: a
solid cylindrical pile embedded in a homogeneous soil layer
over a rigid base, subjected to an axial head load, P. Both soil
and pile are assumed to be homogeneous, isotropic and linearly
elastic. The pile is described by its length, L, diameter, d, and
Young's modulus, E
p
, and the soil by its Young's modulus, E
s
,
and Poisson's ratio, . Stresses and displacements are assumed
to be uniformly distributed within the pile cross-section. Perfect
contact (i.e. no gap or slippage) is considered at the pilesoil
interface.
With reference to the cylindrical coordinate system of Fig. 1,
the vertical equilibrium of the soil medium in the axisymmetric
state of deformation is written as
@(
rz
r)
@r

@
z
@z
r 0 (3)
where
z
is the vertical normal stress and
rz
is the associated
shear stress.
Fundamental to the approximate analysis presented is the
assumption that the normal stress,
z
, and shear stresses,
rz
, in
the vertically loaded medium are controlled exclusively by the
vertical displacement, u
z
; the inuence of radial displacement,
u
r
, on these two stresses is considered to be negligibly small.
Based on this physically motivated simplication, the stress
displacement relations for
z
and
rz
are respectively

z
' M
@u
z
@z
(4)

rz
' G
s
@u
z
@r
(5)
where M is a pertinent constant to be discussed later on.
Equations (4) and (5) were apparently rst employed by Nogami
& Novak (1976) for analysing the corresponding dynamic
problem. In that work, however, the radial displacement of the
medium was assumed to be zero. In the present study the
assumption would be less restrictive: u
r
has negligible inuence
on
z
and
rz
, but is not zero. The importance of this modica-
tion is discussed later on.
From equations (3), (4) and (5), the equation of vertical
equilibrium of the soil medium is expressed as
@
@r
r
@u
z
@r
_ _

@
2
u
z
@z
2
r 0 (6a)
where is a dimensionless parameter given by

M
G
s
(7)
Note that if the variation with depth of the vertical normal
stress
z
is neglected (i.e. @
z
=@z 0), equation (6) simplies
to
@
@r
r
@u
z
@r
_ _
0 (6b)
which is the governing equation of the plane-strain model. The
solution to this equation is
u
z
c
1
ln r c
2
(8)
which clearly diverges with increasing radial distance from the
pile. To overcome the problem, Randolph & Wroth (1978) and
Baguelin & Frank (1979) consider an empirically determined
`magical radius' around the pile beyond which soil displacement
is assumed to be zero. As will be shown below, the solution to
equation (6a) is free of this problem.
Introducing separation of variables and accounting for the
boundary conditions of zero normal tractions at the soil surface
and bounded displacements at large radial distances from the
origin, equation (6a) yields the solution
u
z
(r, z) BK
0
(r) cos z (9)
where K
0
( ) denotes the modied Bessel function of zero order
and rst kind, and is a positive variable. B is a constant to be
determined from the boundary conditions.
Because of the approximate nature of the analysis employed,
the equilibrium of forces in the horizontal direction is not
satised in this approach, nor is the boundary condition of
vanishing shear stresses at the soil surface. Nevertheless, as
demonstrated in studies of several related probems (Tajimi,
1969; Nogami & Novak, 1976; Veletsos & Younan, 1994), these
violations have typically only a minor inuence on the solution.
This will be demonstrated further in this paper through com-
parisons with results from pertinent numerical studies.
P
L
d
z
r
u
t
u
z

rz

u
= 0
d

Fig. 1. System considered


456 MYLONAKIS
Innitely long pile
For an innitely long pile, the displacements and shear
stresses in the medium are obtained by integrating equations (5)
and (9) over the positive variable :
u
z
(r, z)
_
1
0
BK
0
(r) cos z d (10)

rz
(r, z) G
s

_
1
0
BK
1
(r) cos z d (11)
With reference to the pile, the differential equation of vertical
equilibrium is
E
p
A
p
@
2
W(z)
@z
2
d
rz
(d=2, z) F(z) (12)
where F(z) represents body forces distributed along the pile
axis. For the problem at hand, F(z) is determined by resolving
the force at the pile head into equivalent body forces through
the Cosine transformation
F(z)
_
1
0
2P

cos z d (13)
From equations (10)(13), and considering perfect bonding
at the pilesoil interface [i.e. W(z) u
z
(d=2, z)], an explicit
solution is obtained for the pile settlement W:
W(z)
2P
E
p
A
p

_
1
0
K
0
( d=2) cos z

2
K
0
( d=2)
dG
s
E
p
A
p

K
1
( d=2)
_ _ d
(14)
End-bearing pile
For a pile of nite length, one should consider the condition
of vanishing soil displacement at the base of the soil layer.
Imposing this requirement on equation (9) yields

m


2L
(2m 1), m 0, 1 . . . (15)
which corresponds to the solution of the eigenvalue problem
cos(L) 0. In addition, in the same spirit as with the innitely
long pile, the pile-head force, P, can be expanded in Cosine
series as
F(z)

1
m0
2P
L
cos
m
z (16)
The solution to equation (12) is obtained by replacing the
integrals in equations (10) and (11) with corresponding innite
sums involving
m
:
W(z)
2P
E
p
A
p
L

1
m0
K
0
(
m
d=2) cos
m
z

2
m
K
0
(
m
d=2)
d G
s
E
p
A
p

m
K
1
(
m
d=2)
_ _
(17)
where
m
is given by equation (15). The above equation can be
obtained directly from equation (14) by replacing the integral
with an innite sum, and the factor outside the integral with
L, to account for the differences in the forcing functions F(z).
Determination of coefcient
Mention has been already made of the Nogami & Novak
(1976) dynamic solution based on the assumption of vanishing
radial displacement, u
r
. The static part of that solution can be
deduced from equation (17) by assigning factor the value

2(1 )
1 2
(18)
which expresses the ratio of the constrained modulus to the
shear modulus of the soil material. A problem arising from the
use of this equation is that the solution will exhibit a high
sensitivity to Poisson's ratio (recall that the constrained modulus
tends to innity as approaches 05), a behaviour that has not
been observed in rigorous numerical solutions of such problems
(see, for instance, Buttereld & Banerjee, 1971; Selvadurai &
Rajapake 1985). As an alternative, one may assume that the two
horizontal normal stresses,
r
and

, in the vertically loaded


medium are zero, which is analogous to the assumption used by
Veletsos & Younan (1994) for the laterally-loaded problem. In
such a case, equation (18) should be replaced by

2
2(1 ) (19)
Perhaps a better choice for the problem at hand is to consider

r
0 and

0, which captures better (though only approxi-


mately) the condition of zero tangential displacement in the
domain. With the latter assumption,

2
1
(20)
Results obtained from equations (18)(20) are compared
graphically in Fig. 2. It can be seen that the predictions of
equations (19) and (20) remain close over the entire range of
values, whereas equation (18) exhibits a singular behaviour as
approaches 05 and ceases to be acceptable. Except where
specically otherwise indicated, the solutions presented herein
are based on equation (20).
MODEL VALIDATION
Figure 3 compares results for the stiffness of end-bearing
piles computed with the proposed approximate model and with
four nite-element and boundary-element solutions by Poulos &
Davis (1980), Blaney et al. (1975), Sanchez-Salinero (1982),
and El-Sharnouby & Novak (1990). It can be seen that with
small E
p
=E
s
ratios (Fig. 3(a)) the numerical results are sensitive
to the discretisation of the pile. For instance, when a small
number of elements is used, an increase in stiffness with
increasing pile length is observed in some of the solutions for
L=d .50an obviously erroneous trend for end-bearing piles.
El-Sharnouby & Novak (1990) report that a dense discretisation
(of the order of 50 pile elements) is generally needed to remove
this anomaly. In contrast, the present solution exhibits a stable
behaviour and agrees well with the most rigorous results by El-
Sharnouby & Novak. Similar good agreement is observed with
large E
p
=E
s
ratios in Fig. 3(b).
With reference to a long hollow pile in homogeneous half-
space, Table 1 compares results for pile stiffness obtained with
the proposed model and with a rigorous elasto-static solution by
Pak & Ji (1993). Although the two solutions are not strictly
comparable (hollow against solid pile), the agreement between
the predictions is very satisfactory, with the average difference
4
3
2
1
0
2
Constrained medium:
equation (18)
Unconstrained medium:
equation (19)
Partially constrained
medium: equation (20)
01 02 03 04 05
Poisson's ratio,

2
173
Fig. 2. Sensitivity of compressibility coefcient, , to Poisson's ratio
WINKLER MODULUS FOR AXIALLY LOADED PILES 457
being about 5%. The minor effect of the pile Poisson's ratio on
the solution is evident.
EVALUATION OF WINKLER MODULUS
For an innitely long pile, the Winkler modulus, k(z), is
obtained by dividing the side shear traction, f [ d
rz
(d=2)]
(equation (11)), by the corresponding pile settlement, W (equa-
tion (14)). Accordingly,
k(z) d G
s
_
1
0
K
1
( d=2) cos z d
[K
0
( d=2)
dG
s
E
p
A
p

K
1
( d=2)]
_
_
1
0
K
0
( d=2) cos z d

2
[K
0
( d=2)
dG
s
E
p
A
p

K
1
( d=2)]
(21)
For piles of nite length, the corresponding solution is
k(z) dG
s

1
m0
K
1
(
m
d=2) cos
m
z

m
[K
0
(
m
d=2)
dG
s
E
p
A
p

m
K
1
(
m
d=2)]
_

1
m0
K
0
(
m
d=2) cos
m
z

2
m
[K
0
(
m
d=2)
dG
s
E
p
A
p

m
K
1
(
m
d=2)]
(22)
With reference to an ininitely long pile, the variation of
k(z) with depth is presented in Fig. 4 as a function of pilesoil
stiffness ratio, E
p
=E
s
. A decreasing trend with depth is ob-
served in all curves. For points located between about 3 and 20
pile diameters from the surface, k(z) varies between approxi-
mately one and two times G
s
, which is in agreement with
values reported in the literature (Thomas, 1980; Sanchez-
Salinero, 1982; Fleming et al., 1992). With small E
p
=E
s
ratios,
k(z) tends to increase close to the surface but decreases more
rapidly with depth. The singularity observed at z 0 is
analogous to that encountered in elastic analyses of surface
6
5
4
25
20
15
25 50 75 100 125
Dimensionless pile length, L/d
(b)
(a)
N
o
r
m
a
l
i
s
e
d

p
i
l
e

s
t
i
f
f
n
e
s
s

P
d
/
2
W
(
0
)
E
s
A
p
Poulos
(10 elements)
Blaney
(20 elements)
El-Sharnouby
(50 elements)
Salinero
(>20 elements)
Poulos
Blaney
El-Sharnouby
Salinero
Unconstrained medium: equations (17) and (19)
Partially constrained medium:
equations (17) and (20)
Fig. 3. Normalised stiffness of end-bearing piles in a homogeneous
soil stratum over rigid bedrock; comparison of the proposed
approximate model with results from four numerical solutions. (a)
Soft piles, E
p
=E
s
100; (b) stiff piles, E
p
=E
s
1000. Modied from
El-Sharnouby & Novak (1990); 0
:
5
Table 1. Normalised stiffness of a long hollow pile of wall thickness t and Poisson's ratio
p
in a homogeneous halfspace. Comparison of the proposed approximate model with a
rigorous elasto-static solution by Pak & Ji (1993); 0
:
25, t=d 0
:
05

p
log(G
p
=G
s
) E
p

=E
s
{ Normalised pile
stiffness:
Pd
2W(0)E
s
A
p
Difference
(b) (a)
(a)
: %
Pak & Ji (1993)
(a)
Proposed model
(b)
0 10 152 094 098 42
15 481 133 144 83
20 152 200 221 55
25 481 323 348 77
025 10 190 104 105 10
15 601 147 156 61
20 190 231 241 43
25 601 364 381 47
{ E
p

E
p
[1 (1 2t=d)
2
] denotes the Young's modulus of an `equivalent' solid pile having the
same axial rigidity as the hollow pile.
0
5
10
20
15
10
4
10
3
500
10 05 15 20 25 30 35
Normalised Winkler modulus, k/G
s
D
e
p
t
h
,

z
/
d
E
p
/E
s
= 100
Fig. 4. Variation with depth of Winkler modulus for an innitely
long pile in homogeneous halfspace; 0
:
5
458 MYLONAKIS
footings, and has been reported in the literature (Pak & Ji,
1993).
The effect of pile length on k(z) for end-bearing piles is
examined in Fig. 5. It can be seen that in a short pile k(z) is
always larger than in a more slender pile having the same
E
p
=E
s
ratio. For instance, with L=d 15, k(z) can exceed the
value 2G
s
over the entire pile length, which is more than twice
that of the corresponding innitely long pile. The decreasing
trend with depth is analogous to that observed in Fig. 4.
AVERAGE (DEPTH-INDEPENDENT) WINKLER MODULUS
A common approximation in Winkler analyses is that the
ratio k(z)=G
s
is constant along the pile length. While this
introduces some error in the solution, it usually simplies the
analysis by allowing equation (2) to be solved in closed form
within a homogeneous soil layer. Corresponding average Wink-
ler moduli can be derived by matching a key response para-
meter (e.g. pile head settlement) with results from Winkler
formulations. For instance, assuming k=G
s
to be constant within
a homogeneous layer over rigid rock, the solution to equation
(2) is (Scott, 1981)
W(z)
P
E
p
A
p

(cosh z tanh L sinh z) (23)


where is a parameter (units of 1=length) given by

k
E
p
A
p

(24)
Enforcing the settlements at the pile head in equations (17) and
(23) to be equal, the following implicit solution for k is
obtained:
tanh L

k
E
p
A
p

k
E
p
A
p

2
L
2

1
m0
K
0
(
m
d=2)

2
m
K
0
(
m
d=2)
dG
s
E
p
A
p

m
K
1
(
m
d=2)
_ _ (25)
which can be easily solved iteratively once the value of the
right-hand side has been determined.
For an innitely long pile, setting tanh(hL) 1 in equation
(23) and using equation (14) leads to the explicit solution
k E
p
A
p

2
4
_
1
0
K
0
( d=2) d

2
K
0
( d=2)
dG
s
E
p
A
p

K
1
( d=2)
_ _
_

_
_

_
2
(26)
Results obtained from the above expressions are plotted in
Fig. 6. The following points are worthy of note. For the pile
lengths of the most practical interest (say 15 ,L=d ,50), k
varies between about 27 G
s
and 18 G
s
, and tends to decrease
with increasing E
p
=E
s
and L=d. In the limiting cases of
E
p
=E
s
!1 and E
p
=E
s
!0, it can be shown from equations
(25) and (26) that k tends to zero and innity respectively. For
slenderness ratios less than about 50, k is practically indepen-
dent of pilesoil stiffness ratio. In addition, the effect of
Poisson's ratio on the solution was found to be of secondary
importance, as shown in Fig. 7.
In Fig. 8, results from the model are compared graphically
against four empirical expressions from the literature. These
expressions represent average Winkler moduli obtained by curve
tting based on the numerical solutions by Banerjee, Blaney,
and Poulos (see list of references). In general, the predictions of
the model lie close to the average of the empirical values.
Nevertheless, it is evident that these formulae do not satisfy the
limiting behaviour of the solution for small and large values of
E
p
=E
s
and L=d, as discussed above. An improved regression
formula is presented below.
0
10
20
30
40
50
50
25
10 15 20 25 30 35
Normalised Winkler modulus, k/G
s
D
e
p
t
h
,

z
/
d
L/d = 15

Fig. 5. Variation with depth of Winkler modulus for end-bearing


piles in a homogeneous soil layer over rigid bedrock; E
p
=E
s
1000,
0
:
5
0
1
2
10
2
10
3
10
4
W
i
n
k
l
e
r

m
o
d
u
l
u
s
,

k
/
G
s
= 1/3
1/2
1/4
Pilesoil relative stiffness, E
p
/E
s
Fig. 7. Effect of soil Poisson's ratio on average Winkler modulus for
an innitely long pile in homogeneous halfspace
0
1
2
3
50
100
25
10
2
10
3
10
4
W
i
n
k
l
e
r

m
o
d
u
l
u
s
,

k
/
G
s
L/d = 15

Pilesoil relative stiffness, E


p
/E
s
Fig. 6. Depth-independent (`average') Winkler modulus for end-
bearing piles in a homogeneous soil layer over rigid bedrock; 0
:
5
WINKLER MODULUS FOR AXIALLY LOADED PILES 459
NON-LINEAR REGRESSION ANALYSES
For practical applications, it appears desirable to have a
simplied formula to estimate k. Using non-linear regression
analyses based on the LevenbergMarquardt method (Bevington
& Robinson, 1992), the following equation was derived:
k
G
s
' 1
:
3
E
p
E
s
_ _
1=40
1 7
L
d
_ _
0
:
6
_ _
(27)
which corresponds to the depth-independent modulus in equa-
tions (25) and (26). It is evident that this equation satises the
limiting behaviour of the solution, i.e. k 0 at E
p
=E
s
!1;
k !1 at E
p
=E
s
!0 or L=d !0; k finite as L=d !1.
The correlation coefcient between the values in Fig. 6 and
those computed with equation (27) is 0973, and the standard
error is 0114.
CONCLUSIONS
An approximate analytical model was developed for the soil
reaction along an axially loaded pile in a homogeneous elastic
soil stratum. The model was shown to be reasonably accurate
and exhibit stable behaviour over a wide range of parameters
without developing the numerical instabilities with long piles
observed in numerical solutions. The main conclusions of the
study are:
(a) k(z) is singular at the pile head, and tends to decrease with
depth. With long piles and at depths between 3 and 20 pile
diameters, k(z) varies between approximately one and two
times the soil shear modulus, G
s
. With end-bearing piles,
k(z) increases with decreasing pile length, and may exceed
the value of 2 G
s
over the entire pile length.
(b) Depth-independent (`average') Winkler moduli can be
obtained by matching predictions of pile head settlement
with results from Winkler formulations. In the parameter
range of most practical interest (say 100 ,E
p
=E
s
,10 000
and 15 ,L=d ,50), k ranges between approximately
27 G
s
and 18 G
s
, and tends to decrease with increasing
E
p
=E
s
and L=d. In the limiting cases of very stiff
(E
p
=E
s
!0) and very soft piles (E
p
=E
s
!1), k tends
to zero and innity respectively. On the other hand, k
remains nite with innitely long piles, while it tends to
innity as L=d !0. k tends to increase with increasing soil
Poisson's ratio, but this increase is of secondary importance.
NOTATION
A
p
pile cross-sectional area
B integration constant
d pile diameter
G
p
, E
p
pile shear modulus, Young's modulus
G
s
, E
s
soil shear modulus, Young's modulus
f (z) vertical soil reaction per unit pile length
F body force along pile axis
k, k(z) Winkler modulus of subgrade reaction
k
s
coefcient of subgrade reaction
L pile length (= soil thickness)
M soil compression modulus
P pile head load
r radial coordinate
t wall thickness of hollow pile
u
q
, u
r
, u
z
soil displacement components
W(z) pile displacement (settlement)
z vertical coordinate
,
m
positive variable
compressibility coefcient
Winkler parameter
soil Poisson's ratio

p
pile Poisson's ratio

rz
soil shear stress

r
,
q
,
z
soil normal stresses
REFERENCES
Baguelin, F. & Frank, R. (1979). Theoretical studies of piles using the
nite element method. In Numerical methods in offshore piling, pp.
8391. London: Institution of Civil Engineers.
Bevington, P. R. and Robinson, D. K. (1992). Data reduction and error
analysis for the physical sciences. New York: McGraw Hill.
Buttereld, R. & Banerjee, P. K. (1971). The elastic analysis of
compressible piles and pile groups. Geotechnique 21, No. 1, 4360.
Blaney, G. W., Kausel, E. & Roesett, J. M. (1975). Dynamic stiffness of
piles. Proc. 2nd Int. Conf. Num. Methods Geomech., Blacksburg 2,
10101012.
Coyle, H. M. & Reese, L. C. (1966). Load transfer for axially-loaded
piles in clay. J. Soil Mech. Found. Div., ASCE 92, No. SM2, 126.
El-Sharnouby, B & Novak, M. (1990). Stiffness constants and inter-
action factors for vertical response of pile groups. Can. Geotech. J.
27, 813822.
Fleming, W. G. K., Weltman, A. J., Randolph, M. F. & Elson, W. K.
(1992). Piling engineering, 2nd edn. John Wiley & Sons.
Gazetas, G. et al. (1992). Seismic response of soilpilefoundation
structure systems: some recent developments. In Piles under dy-
namic loads, Geotech. Special Publ. No. 34, pp. 5693. American
Society of Civil Engineers.
Guo, W. D. (2000). Vertically-loaded pile in Gibson soil. J. Geotech.
Engng, ASCE 126, No. 2, 189193.
Mattes, N. S. (1969). The inuence of radial displacement compatibility
on pile settlement. Geotechnique 19, No. 1, 157159.
Nogami, T. & Novak, M. (1976). Soilpile interaction in vertical
vibration. Earthquake Engng & Struct. Dyn 4, 277293.
Novak, M. (1974). Dynamic stiffness and damping of piles. Can.
Geotech. J. 11, No. 4, 574598.
O'Rourke, M. J. & Dobry, R. (1978). Spring and dash pot coefcients
for machine foundation on piles, American Concrete Institute,
Detroit, SP-10, 177198.
Pak, R. Y. S. and Ji, F. (1993). Rational mechanics of pilesoil
interaction. J. Engng Mech., ASCE 119, No. 4, 813832.
Poulos, H. G. & Davis, E. (1980). Pile foundation analysis and design,
John Wiley & Sons.
Randolph, M. F. & Wroth, C. P. (1978). Analysis of deformation of
vertically loaded piles. J. Geotech. Engng, ASCE 104, No. 12,
14651488.
Reese, L. C. & Wang, S. T. (1996). Computer program GROUP:
05
10
15
20
25
35
30
100 1000 10000
N
o
r
m
a
l
i
s
e
d

W
i
n
k
l
e
r

m
o
d
u
l
u
s
,

Pilesoil relative stiffness, E


p
/E
s
Sanchez-Salinero (1982):
75(E
p
/E
s
)
02
Thomas (1980):
255(E
p
/E
s
)
0033
O'Rourke & Dobry (1978):
1 + 55(L/d)
074
+ 06(L/d)
13
(E
p
/E
s
)
104
Modified
Randolph & Wroth (1978):
2/ln(L/d)
This paper
(Eqn 27) (Eqn 25)
Fig. 8. Depth-independent Winkler modulus for end-bearing piles in
a homogeneous soil layer over rigid bedrock. Comparison of results
from the proposed model with four empirical expressions from the
literature: L=d 50; 0
:
5
460 MYLONAKIS
Analysis of a group of piles subjected to axial and lateral loading.
Austin, Texas: Ensoft, Inc.
Sanchez-Salinero, I. (1982). Static and dynamic stiffness of single piles,
Geotech. Engng Rep. GR82-31. Austin: University of Texas.
Scott, R. F. (1981). Foundation analysis. Prentice Hall.
Selvadurai, A. P. S. & Rajapakse, R. K. N. D. (1985). On the load-
transfer from a rigid cylindrical inclusion into an elastic half-space.
Int. J. Solids Struct. 21, 12131229.
Terzaghi, K. (1955). Evaluation of coefcients of subgrade reaction.
Geotechnique 5, No. 4, 297326.
Tajimi, H. (1969). Dynamic analysis of a structure embedded in an
elastic stratum. Proc. 4th WCEE, Chile.
Thomas, G. E. (1980). Equivalent spring and damping coefcient for
piles subjected to vertical dynamic loads. MS thesis, Rensselaer
Polytechnic Institute, New York.
Veletsos, A. S. & Younan, A. H. (1994). Dynamic soil pressures on
rigid cylindrical vaults. Earthquake Engng Struct. Dyn. 23, 645
669.
WINKLER MODULUS FOR AXIALLY LOADED PILES 461

S-ar putea să vă placă și