Anda di halaman 1dari 25

EARTHQUAKE ENGINEERING AND STRUCTURAL DYNAMICS

Earthquake Engng Struct. Dyn. 2009; 38:1281–1305


Published online 3 February 2009 in Wiley InterScience (www.interscience.wiley.com). DOI: 10.1002/eqe.892

A model for the 3D kinematic interaction analysis of pile


groups in layered soils

Francesca Dezi1 , Sandro Carbonari2 and Graziano Leoni3, ∗, †


1 Dipartimento di Ingegneria Meccanica e Civile, Università di Modena e Reggio Emilia, Modena, Italy
2 Dipartimento D.A.C.S., Università Politecnica delle Marche, Ancona, Italy
3 Dipartimento ProCAm, Università di Camerino, Ascoli Piceno, Italy

SUMMARY
The paper presents a numerical model for the analysis of the soil–structure kinematic interaction of single
piles and pile groups embedded in layered soil deposits during seismic actions. A finite element model
is considered for the pile group and the soil is assumed to be a Winkler-type medium. The pile–soil–pile
interaction and the radiation problem are accounted for by means of elastodynamic Green’s functions.
Condensation of the problem permits a consistent and straightforward derivation of both the impedance
functions and the foundation input motion, which are necessary to perform the inertial soil–structure
interaction analyses. The model proposed allows calculating the internal forces induced by soil–pile and
pile-to-pile interactions. Comparisons with data available in literature are made to study the convergence
and validate the model. An application to a realistic pile foundation is given to demonstrate the potential
of the model to catch the dynamic behaviour of the soil–foundation system and the stress resultants in
each pile. Copyright q 2009 John Wiley & Sons, Ltd.

Received 23 November 2007; Revised 20 December 2008; Accepted 22 December 2008

KEY WORDS: dynamics; frequency domain; kinematic interaction; impedances; numerical modelling;
pile groups; soil–structure interaction

1. INTRODUCTION

Field evidence and subsequent analyses after earthquakes have confirmed that soil–structure inter-
action plays an important role in evaluating structural performance [1–6].
When non-linear effects are negligible, according to the sub-domain method [7, 8], the soil–
structure interaction may be analyzed by studying the kinematic and the inertial interaction sepa-
rately. The first considers the soil–foundation system (without the superstructure) subjected to the

∗ Correspondence to: Graziano Leoni, Dipartimento ProCAm, Università di Camerino, Ascoli Piceno, Italy.

E-mail: graziano.leoni@unicam.it

Copyright q 2009 John Wiley & Sons, Ltd.


1282 F. DEZI, S. CARBONARI AND G. LEONI

free-field motion due to the earthquake and is responsible for a stress component in the founda-
tion and the motion transmitted to the superstructure. The second considers the superstructure on
deformable restraints with dynamic impedances of the soil–foundation system and is responsible
for the effects of the earthquake on the superstructure and for the additional stresses in the foun-
dation. The action consists of the motion at the foundation level from the kinematic interaction
analysis.
In the case of pile foundations, the kinematic interaction caused by wave propagation in the soil
induces stresses along the pile which, depending on the stratigraphy of the soil, may be of the same
magnitude as those induced at the pile head by the inertial interaction with the superstructure. In
some cases soil–structure interaction may also significantly affect the dynamics of the system and
consequently the design of the superstructure.
The dynamic response of piles during earthquake loading has been analysed in detail in recent
years, and several studies investigated the nature of input ground motion and the mechanism of soil–
pile interaction to determine seismic design loads for pile-supported structures. Modern seismic
codes have acknowledged these aspects and suggest that the effects of soil–structure interaction
should be taken into account when designing the foundation and superstructure.
Some researchers have developed numerical methods using the direct approach where the system
as a whole (soil, pile group and superstructure) is modelled and the seismic response obtained
in just one step. Among others, Yegian and Wright [9], Angelides and Roesset [10], Randolph
[11], Faruque and Desai [12], Trochanis et al. [13, 14], Wu and Finn [15], and Bentley and
El Naggar [16] used the finite element method (FEM), whereas Sanchez [17], Kaynia and
Kausel [5], and Sen et al. [18] implemented the boundary element method (BEM) for the dynamic
response analysis of piles. Both FEM and BEM consider soil as a continuum and give the advantage
of making it possible to perform fully coupled seismic soil–pile–structure interaction analyses.
The two methods are not however commonly used in design mainly because of their excessive
computational cost and complexity in the case of common pile foundation dynamic analyses.
Alternatively, the domain decomposition technique is an approach that may account for soil-
foundation–structure interaction and is commonly used in professional engineering and research
practice. It consists of dividing the domain into sub-domains, where the problem may be studied
by using different models and solving procedures, and introducing suitable impedances at the
interface.
For design purposes the beam model on Winkler soil is commonly used to study pile-soil
interaction because of its versatility in accounting for various complex conditions [19–22]. Closed-
form expressions have also been derived for computing the maximum steady-state bending moment
at the interface between two consecutive soil layers [23–25]. All these methods are based on the
use of dynamic interaction factors by considering two piles at a time: the loaded source pile and
the unloaded receiver pile. This represents a simplification of the pile-to-pile interaction problem,
which is removed in this paper, that presents a numerical method for the kinematic interaction
analysis of pile foundations with a generic number of piles, pile-group geometry and soil profile.
The numerical model is derived discretizing piles and soil in the vertical direction by considering
a formal finite element procedure based on the Lagrange-D’Alembert principle. The pile group is
modelled by means of beam finite elements and the soil is assumed to be a horizontally layered
half space. Both the piles and soil are considered to behave linearly. The pile–soil–pile interaction
is taken into account in the frequency domain by considering the elastodynamic Green’s functions
that make it possible to express the mutual interactions between all the piles of the group and
the radiation problem consistently without using the stepped analysis generally adopted in the

Copyright q 2009 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2009; 38:1281–1305
DOI: 10.1002/eqe
A MODEL FOR THE 3D KINEMATIC INTERACTION ANALYSIS 1283

technical literature. The presence of a rigid cap is accounted for by constraining the displacements
of the pile heads. This allows obtaining the impedances of the pile group and the motion at the cap
necessary to study the inertial interaction with the superstructure in a straightforward manner. In
particular, the procedure makes it possible to calculate all the components of the impedance matrix
including the rocking-translation coupling. It is worth mentioning that the assumption of rigid cap
may be relaxed if necessary to account for flexible foundation structures. The input motion may be
constituted simply by harmonic excitations or by transient earthquake motions. In the latter case,
the incoming free field may be derived from local one-dimensional or spatial analysis depending
on the complexity of the site.
The proposed model is validated by comparing the results obtained for sample pile groups
with other results available in the literature. It revealed to be efficient and able to predict the
soil–foundation dynamic behaviour, without significant loss of accuracy, as compared with the
rigorous thin-layered solution of Kaynia and Kausel [5].
Finally, the analysis of a realistic pile foundation system is performed to demonstrate the potential
of this method in catching the dynamic behaviour of the pile group.
The model presented may be particularly attractive in the real design of pile foundations
embedded in layered soils since it allows combining kinematic and inertial response analyses, as
no interaction factors between piles need to be employed.

2. SOIL–PILE FOUNDATION INTERACTION ANALYSIS

2.1. Analytical model


This section presents a model for the soil–pile foundation kinematic interaction analysis. The
problem is formulated in the frequency domain so that the forces and displacements reported are
implicitly assumed to be the Fourier Transforms of the corresponding quantities expressed in the
time domain.
A group of n circular piles having length L and diameter  is considered. A right-handed
reference system frame {0; x1 , x2 ; z} is defined as shown by Figure 1. Each pile is assumed to be
a Euler–Bernoulli beam, namely it is assumed that the cross section remains plane and orthogonal
to the deformed longitudinal axis. If  is the circular frequency, the displacements at depth z are
described in the frequency domain by the complex-valued vector
uT (; z) = [uT1 · · · uTp · · · uTn ] (1)
which groups the displacements measured at the axis of the n piles constituting the group. Each
sub-vector u p (; z), referring to the pth pile, contains the displacement components u p1 , u p2 and
u p3 along the directions x1 , x2 and z, respectively. According to the Euler–Bernoulli model, the
strains are described by the pile curvatures on planes orthogonal to xi and by the overall normal
strain; similar to the displacements, they are conveniently grouped in the vector

DuT (; z) = [D̃uT1 · · · D̃uTp · · · D̃uTn ] (2)


obtained by applying the formal differential operator
 2 2 
D̃u p (; z) = * u p1 − * u p2
T *u p3 (3)
*z 2 *z 2 *z

Copyright q 2009 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2009; 38:1281–1305
DOI: 10.1002/eqe
1284 F. DEZI, S. CARBONARI AND G. LEONI

xi

z = + +

uff uff -r r

(a) (b) (c)

Figure 1. Decomposition of the domain: (a) free field; (b) soil subjected to the interaction forces; and
(c) foundation subjected to the interaction forces.

The piles are subjected to forces distributed on the lateral surface due to the soil reaction. The
complex-valued vector

rT (; z) = [rT1 · · · rTp · · · rTn ] (4)

groups the resultants of such forces in the frequency domain. Similar to the displacements, each
sub-vector r p (; z) contains the force components r p1 ,r p2 and r p3 . Notice that moments are
neglected as usual for pile foundations.
If the piles are subjected to the motion (1), inertia forces 2 Mu(; z) arise. The mass matrix
of the pile group assumes the form

⎡ ⎤
I ··· 0 ··· 0
⎢ ⎥
⎢. .. .. ⎥
⎢ .. . .⎥
⎢ ⎥
⎢ ⎥
⎢ ⎥
M = p A ⎢0 · · · I · · · 0⎥ (5)
⎢ ⎥
⎢ ⎥
⎢ .. .. .. ⎥
⎢. . .⎥
⎣ ⎦
0 ··· 0 ··· I

where A is the pile cross section area,  p is the material density and I is the identity matrix of
size 3. Here again, as usual for piles, rotational inertias are neglected.
Under the assumption of linearly elastic material with Young’s modulus E, and by denoting the
moment of inertia of the pile cross section by I , the pile stress resultants (bending moments and
axial force) are given by the linear relationship

s(; z) = KDu(; z) (6)

Copyright q 2009 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2009; 38:1281–1305
DOI: 10.1002/eqe
A MODEL FOR THE 3D KINEMATIC INTERACTION ANALYSIS 1285

where

⎡ ⎤
K1 ··· 0 ··· 0
⎢ .. .. .. ⎥
⎢ ⎥
⎢ . . . ⎥
⎢ ⎥
⎢ ⎥
K=⎢ 0 ··· Kp ··· 0⎥ (7)
⎢ ⎥
⎢ .. .. .. ⎥
⎢ ⎥
⎣ . . . ⎦
0 ··· 0 · · · Kn

is the real-valued stiffness matrix of the pile group obtained by assembling the pile cross section
sub-matrices

⎡ ⎤
I 0 0
⎢ ⎥
Kp = E ⎣0 I 0⎦ (8)
0 0 A

The equilibrium condition of the pile group (Figure 1(c)) may be expressed in weak form by the
Lagrange-D’Alembert principle by assuming that the work resulting from external forces and inertia
forces acting through every virtual consistent displacement field û(z) is equal to that resulting from
stresses acting through every virtual strain field Dû(z). In the frequency domain, this provides the
following equation:


L
L
L
KDu(; z)·Dû(z) dz − r(; z)· û(z) dz − 2
Mu(; z)· û(z) dz = 0 ∀û  = 0 (9)
0 0 0

The soil reaction forces r develop along the pile as a result of the pile–soil–pile interaction.
Under the assumptions of linear behaviour for the soil and that no gaps arise during the motion,
the compatibility condition between the pile and the soil displacements is expressed by the integral
relationship


L
u(; z) = uff (; z)− D(; , z)r(; ) d (10)
0

The right-hand side of Equation (10) represents the soil displacements measured at the pile locations
expressed by superimposing the free-field motion uff (Figure 1(a)) and the displacements induced
by the soil–pile interaction forces (Figure 1(b)). In particular, the incoming free-field motion is
described by the complex-valued vector

uTff (; z) = [uT1ff . . . uTpff . . . uTn ff ] (11)

Copyright q 2009 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2009; 38:1281–1305
DOI: 10.1002/eqe
1286 F. DEZI, S. CARBONARI AND G. LEONI

obtained by assembling the soil displacements at the pile location as in Equation (1). Furthermore,
the kernel
⎡ ⎤
D11 . . . D1q . . . D1n
⎢ . .. .. ⎥
⎢ . ⎥
⎢ . . . ⎥
⎢ ⎥
⎢ ⎥
D(; , z) = ⎢ D p1 . . . D pq . . . D pn ⎥ (12)
⎢ ⎥
⎢ . .. ⎥
.. ⎥
⎢ .
⎣ . . . ⎦
Dn1 ... Dnq ... Dnn
is a complex-valued matrix consisting of the sub-matrices D pq (; , z) that contain the elastody-
namic Green’s functions [8, 26] which express the soil displacement at the location of the pth
pile at depth z, due to a time-harmonic unit point load acting at the location of the qth pile
at depth . In generic cases, all the components of the D pq matrices are non-zero according to
the non-local nature of the problem. Equation (10) describes the complex pile–soil–pile dynamic
interaction and, considering that the Green’s functions are defined for an infinite domain, it is also
able to describe the radiation problem which is an important energy dissipation mechanism of the
foundation system. It is worth noticing that Equation (10) is a formal expression and may be used
in practical cases only if the closed-form expressions of the elastodynamic Green’s functions are
known.
The problem may be simplified substantially by making use of the Baranov assumption, namely
the extension of the well-known Winkler model to dynamic loads. This implies introducing the
following form for the kernel (12):

D(; , z) = D̂(; z)(z −) (13)


where (z −) is Dirac’s delta function and D̂ is a complex-valued matrix containing the elasto-
dynamic Green’s function defined on the infinite plane at depth z.
By taking into account the property of Dirac’s delta function, Equation (10) transforms into

u(; z) = uff (; z)− D̂(; z)r(; z) (14)


namely, the displacements at depth z are assumed to be dependent only on the forces applied at the
same depth. This is equivalent to supposing that the soil consists of independent thin horizontal
layers [19].
Since matrix D̂ is not singular, the soil complex-valued impedance matrix may be defined as

K S (; z) = D̂(; z)−1 (15)


and thus the soil–pile interaction forces may be evaluated from
r(; z) = −K S (; z)[u(; z)−uff (; z)] (16)
By substituting (16) into (9), the following equation is obtained:

L
L
L
L
KDu·Dû dz + K S u· û dz −2 Mu· û dz = K S uff · û dz ∀û  = 0 (17)
0 0 0 0

Copyright q 2009 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2009; 38:1281–1305
DOI: 10.1002/eqe
A MODEL FOR THE 3D KINEMATIC INTERACTION ANALYSIS 1287

Equation (17) represents a global balance condition and may be used to solve the problem with
variational methods like Ritz’s or FEMs. In this paper the latter is used to obtain a tool for the
practical analysis of pile foundations. Contrary to Dobry and Gazetas [4], Makris et al. [27] and
Mylonakis [6], who obtained the response of a pile group by means of interaction factors derived
by considering only two piles at a time, with this approach the pile-to-pile interaction is captured
between all the piles of the group at the same time. In other words, this model does not require a
stepped analysis based on the active and passive pile concepts and that of scattered motion.

2.2. Finite element solution


In the displacement-based approach, the solution of the problem may be achieved numerically by
the FEM, by dividing each pile into e elements and approximating the motion at the interior of the
piles by interpolating the displacements at the end nodes. The following third-order polynomials
are adopted for transverse displacements:

3z 2 2z 3 z 2z 2 z 3
n 1 (z) = 1− 2 + 3 , n 2 (z) = L − 2+ 3 (18a, b)
L L L L L
2 2
3z 2z 3 z z3
n 3 (z) = − 3 , n 4 (z) = L − 2 + 3 (18c, d)
L2 L L L
whereas first-order polynomials are adopted for longitudinal displacements
z z
n 5 (z) = 1− , n 6 (z) = (19a, b)
L L
Five degrees of freedom are thus associated with the ith node of the mesh, namely three translations
and two rotations, which are grouped in the nodal vector
diT () = [u 1i , u 2i , u 3i , 1i , 2i ] (20)
If N p (z) is the matrix of the interpolating polynomials,
⎡ ⎤
n1 0 0 0 n2 n3 0 0 0 n4
⎢ ⎥
N p (z) = ⎣ 0 n 1 0 −n 2 0 0 n 3 0 −n 4 0⎦ (21)
0 0 n5 0 0 0 0 n6 0 0
the displacements in the element of the pth pile may be expressed as
u p (z; ) ∼
= N p (z)dep () (22)
where dep is the nodal displacement vector of the pth pile element having the end nodes h and k
T
dep () = [u 1h , u 2h , u 3h , 1h , 2h , u 1k , u 2k , u 3k , 1k , 2k ] p (23)
It is worth mentioning that the twisting rotation is not considered to avoid singular problems due
to the assumption of zero torsional stiffness for piles.
The vector containing the displacements of the n piles of the group defined by (1) is conveniently
approximated in the form
u(; z) ∼
= Nde (24)

Copyright q 2009 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2009; 38:1281–1305
DOI: 10.1002/eqe
1288 F. DEZI, S. CARBONARI AND G. LEONI

where de is the vector that groups the nodal displacements of all the piles.
T T T T
de = [de1 · · · dep · · · den ] (25)
and N(z) is the matrix of the interpolating polynomial functions for the group obtained by assem-
bling matrices N p
⎡ ⎤
N1 · · · 0 · · · 0
⎢ . .. .. ⎥
⎢ . ⎥
⎢ . . . ⎥
⎢ ⎥
⎢ ⎥
N(z) = ⎢ 0 · · · N p · · · 0 ⎥ (26)
⎢ ⎥
⎢ . .. ⎥
.. ⎥
⎢ .
⎣ . . . ⎦
0 ··· 0 · · · Nn
By considering the approximation (24) and the contribution of all the elements the global balance
condition (17) becomes

Le
Le
Le

E E e 
E e
K(DN)de ·(DN)d̂e dz + K S Nde ·Nd̂ dz −2 MNde ·Nd̂ dz
e=1 0 e=1 0 e=1 0


E Le
e
= K S uff ·Nd̂ dz ∀d̂e  = 0 (27)
e=1 0

By appropriately assembling the node displacements in a unique displacement vector d(), standard
considerations make it possible to obtain the complex linear equation system
(K̄ P −2 M̄+ K̄ S )d = f (28)
where


E Le
K̄ P () = (DN)T K(DN) dz (29a)
e=1 0


E Le
M̄ = NT MN dz (29b)
e=1 0


E Le
K̄ S () = NT K S N dz (29c)
e=1 0

are the global stiffness matrix of the piles, the global mass matrix of the piles, and the global
impedance matrix of the soil, obtained by assembling the relevant contributions of all the
elements, and

Le
E
f() = NT K S uff dz (30)
e=1 0

is the vector of the external loads due to the free-field motion.

Copyright q 2009 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2009; 38:1281–1305
DOI: 10.1002/eqe
A MODEL FOR THE 3D KINEMATIC INTERACTION ANALYSIS 1289

Φ3
Φ2
Φ1 U1

U2
U3

Figure 2. Schematic model of the pile group embedded in the Winkler-type


medium and connected by a rigid cap.

It is worth noticing that both the global impedance matrix of the soil and the vector of external
loads are defined consistently with the finite element admissible displacements avoiding to lump
the soil contributions.
When piles are connected to a rigid cap, a constraint must be imposed to the pile heads. This is
obtained by introducing a master node with the six generalized displacement components collected
in the vector

UTF () = [U1 ,U2 ,U3 , 1 , 2 , 3 ] (31)

In Figure 2, Ui and i are the displacement and the rotation components of the rigid cap,
respectively. It is worth mentioning that in this case the torsional rotation component is also
taken into account. Notice also that the position of the master node may be generic; in practical
applications this should be placed at the connection between the foundation cap and the elements
of the superstructure.
Analytically, the constraint may be written as

d() = Ad̃() (32)

where A is the geometric matrix of the rigid constraint, and


 
UF
d̃() = - - - (33)
dE

is the vector of the displacements necessary to describe the foundation motion consisting of six
generalized displacements of the rigid cap (sub-vector U F ) and by the displacements at the nodes of
the embedded pile sections (sub-vector d E ). The application of the constraint transforms Equation
(28) into

AT (K̄ P −2 M̄+ K̄ S )Ad̃ = AT f (34)

Copyright q 2009 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2009; 38:1281–1305
DOI: 10.1002/eqe
1290 F. DEZI, S. CARBONARI AND G. LEONI

which may be written in the concise form

K̃()d̃() = f̃() (35)

where K̃() is the complex-valued dynamic stiffness matrix of the foundation and f̃() is the
complex-valued vector of the generalized applied forces due to the free-field motion.
System (35) permits analysing kinematic soil–pile interaction. Once the unknown displacements
are evaluated, the stress resultants in the pth pile may be calculated from Equation (6) taking into
account the approximation of displacements (24) that yields

s(; z) = KDN(z)de () (36)

For the sake of simplicity only the case of one rigid cap was presented but obviously the
formulation may easily be extended to the case of multiple rigid caps or deformable caps. Multiple
rigid caps may be studied by introducing different rigid body constraints whereas deformable caps
have to be modelled by introducing suitable finite elements.

2.3. Impedance of the pile group and foundation input motion


With reference to a group of piles connected at the heads by a rigid cap, it is interesting to
determine the forces necessary to induce time-harmonic unit displacements at the cap master node
(impedances) and the motion of the master joint due to the application of free-field displacements
(foundation input motion). This makes it possible to perform the inertial soil–structure interaction
by considering only the superstructure with deformable restraints having suitable dynamic stiffness
and damping.
The foundation impedance matrix and the foundation input motion may easily be calculated by
partitioning Equation (35) consistently with vector d̃
    
--------

K̃FF K̃FE U F f̃ F
- - -------------- - - - - - = - - - (37)
K̃EF K̃EE dE f̃ E
After simple manipulations, the following equation is obtained:

(K̃FF − K̃FE K̃−1 −1


EE K̃EF )U F = f̃ F − K̃FE K̃EE f̃ E (38)

from which

Z F () = (K̃FF − K̃FE K̃−1


EE K̃EF ) (39)

is the complex-valued foundation impedance matrix.


Finally, the motion of the master node (foundation input motion) may be expressed as

U F () = Z−1 −1
F [f̃ F − K̃FE K̃EE f̃ E ] (40)

2.4. Free-field motion


The analysis methodology developed may be used to assess the effects of kinematic interaction in
a generic soil–foundation system. Generic pile group geometry, layered soil profile and free-field
motion may be considered.

Copyright q 2009 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2009; 38:1281–1305
DOI: 10.1002/eqe
A MODEL FOR THE 3D KINEMATIC INTERACTION ANALYSIS 1291

The proposed model is the core of the soil–structure interaction analysis for practical applications.
It provides the impedances of the soil–pile system and the displacements of the master node for the
inertial analysis of the superstructure and the displacements of the embedded piles. In particular
the following steps are necessary to define the free-field input motion:
1. selection of the accelerograms at outcropping soil;
2. derivation of the free-field motion at bedrock;
3. local site response analysis to obtain the three translational displacement components of the
soil at the location of the embedded pile nodes;
4. Fourier Transform of the free-field motion if the local site response analysis is performed in
the time domain.
If the soil layers are horizontal and the site is sufficiently far from the hypocenter, it is common
practice in geotechnical earthquake engineering to use the well-known one-dimensional equivalent
linear response analysis to estimate the ground response.
In the analyses reported in this paper, one-dimensional site response analyses were performed
to obtain the components of the soil motion at the location of each pile of the group, nevertheless,
the local site response analysis is independent from the model and may be performed with the
suitable degree of completeness and accuracy.

2.5. The elastodynamic Green’s functions


The application of the proposed procedure is straightforward once the elastodynamic Green’s
functions for the soil domain are given in closed form. In this paper, they are obtained starting
from impedances and dynamic interaction factors available in the literature [4, 21, 28].
Considering Winkler’s assumption, the problem is studied separately for each layer by consid-
ering vertical and horizontal time-harmonic unit point loads. The first is assumed to induce only
shear waves propagating radially from the source point: consequently, only vertical displacements
are induced in the layer. Horizontal loads are assumed to induce in-plane displacement fields. In
view of the particular problem symmetry, the displacements of the points placed along the load
axis arise in the same direction and are due to the propagation of a pressure wave. Furthermore,
displacements of the points placed along the line passing through the source point orthogonally
to the load axis are due to the propagation of shear waves and are assumed to be parallel to the
load direction. For generic points of the layer, the displacements induced by the horizontal load
have two components, i.e. one radial and one circumferential with respect to the application point
of the load, and vary with the first harmonic of  [29] corresponding to the angle formed by
the direction of the load and the line passing through the load application and the generic points
(Figure 3(a)).
Analytically, if f is a time-harmonic point load, the displacement field is described by the linear
application as

u(; s, ; z) = D̃(; s, ; z)f(; z) (41)

where s and are polar coordinates defined as in Figure 3(b) and D̃ is a 3×3 matrix.
According to Wolf [29] (Figure 3) and by using the impedances of the single pile and the
dynamic interaction factors proposed by Dobry and Gazetas [4], Makris and Gazetas [21, 28] and

Copyright q 2009 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2009; 38:1281–1305
DOI: 10.1002/eqe
1292 F. DEZI, S. CARBONARI AND G. LEONI

P waves x1

Point load ψ fx1


θ
fx2 s
SH waves
x2 ~ ~
D11(s, θ) fx1+ D21(s, θ) fx2

~ ~
(a) (b) D12(s, θ) fx1+ D22(s, θ) fx2

Figure 3. (a) Wave propagation from a point load and (b) conventions adopted in the paper.

Mylonakis Gazetas [30] the following components of D̃ are derived:


1/2
kh −ich −(
+i)(s−d/2)/VLa −(
+i)(s−d/2)/Vs d
D̃11 (; s, ) = 2 [e cos +e
2
sin ]
2
(42a)
kh + ch
2 2 2s
kh −ich  −(
+i)(s−d/2)/VLa
2

D̃22 (; s, ) = 2 e cos −
kh +2 ch2 2
 d 1/2
−(
+i)(s−d/2)/Vs 2
+e sin − (42b)
2 2s
kh −ich
D̃12 (; s, ) = D̃21 (; s, ) =
kh2 +2 ch2
1/2
−(
+i)(s−d/2)/VLa −(
+i)(s−d/2)/Vs d
×[e −e ] (sin cos ) (42c)
2s
1/2
kv −icv d
D̃3 (; s) = 2 e−(
+i)(s−d/2)/Vs (42d)
kv +2 cv2 2s

where d is the pile diameter,


is the soil hysteretic damping, Vs is the velocity of the shear waves
and VLa is the ‘Lysmer’s analogue’ velocity [31, 32] defined as
3.4
VLa = Vs (43)
(1− )
where is Poisson’s ratio of the soil. Furthermore, if E s and s are the soil Young’s modulus and
density, respectively, for the layer at depth z it follows that

kh = 1.2E s (44a)
  −0.25
VLa d kh
ch () = 2ds Vs 1+ +2
(44b)
Vs Vs 

Copyright q 2009 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2009; 38:1281–1305
DOI: 10.1002/eqe
A MODEL FOR THE 3D KINEMATIC INTERACTION ANALYSIS 1293

  
1 d
kv () = 0.6E s 1+ (44c)
2 Vs
−0.25
d kv ()
cv () = 1.2 ds Vs +2
(44d)
Vs 
are the soil frequency-dependent stiffness and damping in the horizontal (h) and vertical (v)
directions. The latter equations also represent the soil impedances used in the soil–pile interaction
in the case of single piles.
By considering the load application point coincident with the pile q location and the generic
point coincident with the pile p location, matrix D̂ that appears in Equation (13) may be obtained
by assembling sub-matrices

D̂ pq (; z) = D̃(; sq p , q p ; z) (45)

3. VALIDATION OF THE MODEL

This section reports some of the results of the analyses performed to validate the proposed model.
The efficacy of the model to describe the kinematic response of pile groups is investigated with
reference to the foundation impedances, the dimensionless displacement response factors and the
pile stress resultants. In the sequel, reference will be made to the notations defined hereafter. By
assuming square pile patterns and suitably fixing the reference system, the impedance matrix Z F
assumes the following form:
⎡ ⎤ ⎡ ⎤
Kh 0 0 0 K hr 0 Ch 0 0 0 C hr 0
⎢ ⎥ ⎢ ⎥
⎢ Kh 0 −K hr 0 0⎥ ⎢ Ch 0 −C hr 0 0⎥
⎢ ⎥ ⎢ ⎥
⎢ ⎥ ⎢ ⎥
⎢ Kv 0 0 0⎥ ⎢ Cv 0 0 0⎥
ZF = ⎢⎢ ⎥ ⎢ ⎥ (46)
⎥ +i ⎢ ⎥
⎢ Kr 0 0⎥ ⎢ Cr 0 0⎥
⎢ ⎥ ⎢ ⎥
⎢ 0⎥ ⎢ 0⎥
⎣ Sym Kr ⎦ ⎣ Sym Cr ⎦
Kt Ct

in which the real and imaginary parts are made explicit. Furthermore, the dimensionless displace-
ment response factors are defined according to the equations

Up
Iu = (47)
Uff
pd
I = (48)
Uff
where Uff is the amplitude of the free-field motion measured at the ground surface, d is the pile
diameter and U p and  p are the amplitudes of the rigid cap horizontal displacement and rocking,
respectively [33].

Copyright q 2009 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2009; 38:1281–1305
DOI: 10.1002/eqe
1294 F. DEZI, S. CARBONARI AND G. LEONI

3.1. Numerical convergence


In order to evaluate the effects of the finite element size on the results obtained with the proposed
numerical model, accuracy analyses were performed taking into account different finite element
lengths L e , the soil profiles and pile group patterns. For the sake of brevity, the results only for a
2×2 floating pile group embedded in a two-layered soil profile are discussed (Figure 4). Analyses
were repeated by reducing the finite element size through to convergence between two consecutive
cases.

0 10 20 30 f [Hz] 50 0 10 20 30 f [Hz] 50
15⋅105 60⋅105
Le = 4 m
[kN/m] Le = 2 m [kN/m]
Le = 1 m
Le = 0.5 m ωCh
10⋅105 Kv
Le = 0.2 m 40⋅105
Kh - ωCh

Kv - ωCv
75⋅104
Kh 30⋅105
50⋅104
20⋅105
25⋅104 ωCv
0 10⋅106

-25⋅104 0
30⋅105 15⋅105
Kr [kNm/rad]
[kNm/rad]

10⋅105
5
20⋅10
ωCt
Kr - ωCr

75⋅104

Kt - ωCt
15⋅105 ωCr 50⋅104

25⋅104
10⋅105
0
5⋅105 Kt -25⋅104

0 -50⋅104
90⋅104 0 10 20 30 f [Hz] 50
[kN/rad]

70⋅104 x1
60⋅10 4
ωChr x2
Khr - ωChr

50⋅104 Vs1 = 200 m/s


ρs1 = 1.5 Mg/m³
40⋅10 4
12 m z ν = 0.4
30⋅104 Khr 1.7m = 2η = 10%
Vs2 = 400 m/s
20⋅104 8m ρs2 = 1.8 Mg/m³
10⋅10 4 ν = 0.4
0.6 m = 2η = 10%
0
0 10 20 30 f [Hz] 50

Figure 4. Dynamic impedances of the soil–pile group system for different finite element sizes.

Copyright q 2009 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2009; 38:1281–1305
DOI: 10.1002/eqe
A MODEL FOR THE 3D KINEMATIC INTERACTION ANALYSIS 1295

1.40 0.20
Le = 4 m
Iu Le = 2 m

Le = 1 m
1.00 Le = 0.5 m
0.12 Le = 0.2 m
0.80

0.60 0.08
0.40
0.04
0.20

0 0
0 10 20 30 f [Hz] 50 0 10 20 30 f [Hz] 50

Figure 5. Dimensionless displacement response factors for different finite element sizes.

-100 Mx2 [kNm] 100 -60 Vx1 [kN] 60 -50 N [kN] 50


0
Le = 4 m
z Le = 2 m
Le = 1 m
Le = 0.5 m
Le = 0.2 m

10

[m]

20

Figure 6. Stress resultants for different finite element sizes.

Figure 4 shows the effects of the mesh sizing with respect to the components of the impedance
matrix (46). In the frequency range 0–50 Hz a fast convergence is obtained and element sizes
smaller than 2 m give the same results.
Figure 5 reports the displacement response factors. In this case convergence is fast for low
frequencies (0–10 Hz) while refined meshes are required for higher frequencies (10–50 Hz). This
result was predictable since the number of nodes necessary to catch the displacement profile of
the propagating waves increases with the wave number [34].
Figure 6 shows the envelopes of stress resultants for one of the four piles. Bending moments
vary linearly within the elements whereas shear and axial forces are constant due to the shape
function used to interpolate the displacements. Discontinuities of solutions deriving from the
displacement-based approach of the finite element procedure are no more significant for mesh
sizing under 1.0 m.

Copyright q 2009 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2009; 38:1281–1305
DOI: 10.1002/eqe
1296 F. DEZI, S. CARBONARI AND G. LEONI

3.2. Soil–foundation impedance functions


The proposed model was validated by comparing the impedance functions calculated as shown
in the previous section with those available in the literature. Normalized dynamic stiffnesses and
dampings of simple pile-group foundation systems are compared with those achieved by Kaynia
and Kausel [5], with a boundary element approach, and by Dobry and Gazetas [4] and Mylonakis
[6] with other simplified methods. The elastodynamic Green’s functions were developed starting
from the dynamic impedances and the attenuation laws proposed by Dobry and Gazetas [4] and
Makris and Gazetas [21, 28] as discussed in the previous section.
Groups of 2×2 and 3×3 equally spaced floating cylindrical piles embedded in a uniform
homogeneous soil, connected by a rigid cap, were considered. Each pile was divided into 20
elements in order to achieve convergence. By considering s as the pile spacing and d as the
diameter, the ratios s/d = 2, 5 and 10 were taken into account. The data reported by Figure 7 were
adopted in the analyses.

3.2.1. Vertical response. The solutions provided by the proposed model were compared with those
obtained by Dobry and Gazetas [4] and Mylonakis [6]. Figures 8(a) and (b) show the dynamic
stiffness and the dynamic damping of the pile group versus the non-dimensional frequency factor
a0 = d/Vs for the 2×2 pile group. The values are normalized by dividing the dynamic stiffness
K v , and the dynamic damping Cv , by the sum of the static stiffness of the single pile K̄ v .
The proposed model well fits the results obtained by Dobry and Gazetas [4] even if in the case
of s/d = 5 the curves obtained shifted slightly towards higher values of a0 . Results also agree well
with the solutions proposed by Mylonakis [6].
It is worth mentioning that at zero frequency, namely at static conditions, the model may provide
the static group efficiency factors. Furthermore, it also captures the wave interference phenomena
that beyond certain frequencies (depending on the s/d ratio), dominate the response of the pile
group-soil system. For low s/d ratios the interference phenomena are not evident for the frequency
range examined.

3.2.2. Lateral response. In this case, comparisons are made with the impedance functions obtained
by Kaynia and Kausel [5] and Mylonakis [6]. The dynamic stiffness K h and damping C h plotted
in Figures 9(a) and (b) refer to the 3×3 pile group: as in the previous case, values are normalized
with respect to the group reference static stiffness calculated as the sum of the static stiffness of
the single pile K̄ h .
Here again the proposed model behaves well. The results of Mylonakis [6] are reproduced
precisely, while comparisons with the results obtained by Kaynia and Kausel [5] are somewhat
less accurate.

x1 x2
Elastic modulus ratio Ep/Es = 1000
x2 z d Density ratio ρp /ρs = 1.5
L
Soil Poisson modulus ν = 0.4
Soil hysteretic damping = 10%
s

Figure 7. Application data.

Copyright q 2009 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2009; 38:1281–1305
DOI: 10.1002/eqe
A MODEL FOR THE 3D KINEMATIC INTERACTION ANALYSIS 1297

4.0 4.0
Proposed model
L/d = 15 s/d = 10
3.0 3.5 s/d = 5
s/d = 2
Mylonakis, 1995
2.0 3.0 s/d = 10
s/d = 5
s/d = 2 L/d = 15
1.5 2.5 Dobry and Gazetas, 1988
s/d = 10
K v/4K v

ωC v/4K v
s/d = 5
0.0 2.0 s/d = 2
s/d = 10
s/d = 5 Proposed model
-1.0 s/d = 2 1.5

s/d = 10
-2.0 s/d = 5 Mylonakis, 1995 1.0
s/d = 2

-3.0 s/d = 10 0.5


s/d = 5 Dobry and Gazetas, 1988
s/d = 2
-4.0 0.0
(a) 0.0 0.2 0.4 0.6 a0 1.0 (b) 0.0 0.2 0.4 0.6 a0 1.0

Figure 8. (a) Vertical dynamic stiffness and (b) vertical dynamic damping.

3.0 3.5
Proposed model Proposed model
L/d = 20
s/d = 10 L/d = 20 s/d s/d
= 10= 10
s/d = 5 s/d s/d
= 5= 5
2.5 s/d = 2 3.0
s/d s/d
= 2= 2
Mylonakis, 1995 Mylonakis, 1995
s/d = 10
s/d s/d
= 10= 10
2.0 s/d = 5 2.5
s/d = 2 s/d s/d
= 5= 5
Kaynia and Kausel, 1982 s/d s/d
= 2= 2
s/d = 10 Kaynia and Kausel, 1982
1.5 2.0
Kh/9Kh

s/d s/d
= 10= 10
ωC h/9Kh

s/d = 5
s/d = 2 s/d s/d
= 5= 5
s/d s/d
= 2= 2
1.0 1.5

0.5 1.0

0.0 0.5

-0.5 0.0
(a) 0.0 0.2 0.4 0.6 a0 1.0 (b) 0.0 0.2 0.4 0.6 a0 1.0

Figure 9. (a) Lateral dynamic stiffness and (b) lateral dynamic damping.

3.2.3. Rocking response. The 3×3 pile group is considered and the results of the proposed model
are compared with those of Kaynia and Kausel [5] and Mylonakis [6]. Dynamic stiffness 2K r and
damping Cr are normalized with respect to the static rocking stiffness evaluated as xi K̄ v .

Copyright q 2009 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2009; 38:1281–1305
DOI: 10.1002/eqe
1298 F. DEZI, S. CARBONARI AND G. LEONI

4.5 4.5
Proposed model Proposed model
s/d = 10 s/d = 10
4.0 s/d = 5 4.0 s/d = 5
s/d = 2 s/d = 2
Mylonakis, 1995 Mylonakis, 1995
3.5 s/d = 10 3.5 s/d = 10
s/d = 5 s/d = 5
3.0 s/d = 2 L/d = 20 3.0 s/d = 2 L/d = 20
Kaynia and Kausel, 1982 Kaynia and Kausel, 1982

ωCr/Σxi2Kv
s/d = 10
Kr/Σxi2Kv

s/d = 10
2.5 s/d = 5 2.5 s/d = 5
s/d = 2 s/d = 2
2.0 2.0

1.5 1.5

1.0 1.0

0.5 0.5

0.0 0.0
(a) 0.0 0.2 0.4 0.6 a0 1.0 (b) 0.0 0.2 0.4 0.6 a0 1.0

Figure 10. (a) Rocking dynamic stiffness and (b) rocking dynamic damping.

The results shown in Figures 10(a) and (b) again demonstrate the validity of the proposed model.
The results of Mylonakis [6] are well caught while those of Kaynia and Kausel [5], deriving from
a more refined boundary element solution, are somewhat less accurate.

3.3. Foundation input motion


In addition to the foundation impedances, the foundation input motion is another important result of
the kinematic interaction required to carry out the consequent inertial interaction of the compliance-
based structure. In order to compare this substantially with other models, dimensionless displace-
ment response factors, expressed as the ratio between the soil–foundation system motion measured
at the pile cap and the free ground surface response, are considered. The data reported by Fan
et al. [33] are used as benchmarks.
Floating single free-head piles and pile groups of different geometries were considered. Two
E p /E s ratios and different pile spacing-diameter ratios were also taken into account. All the piles
are embedded in a homogeneous soil deposit having density s = 0.7 p , constant Poisson ratio
= 0.4 and constant hysteretic damping = 2
= 10%. Each soil–pile group system is excited by
vertically propagating harmonic shear waves that produce a horizontal free-field oscillation of
amplitude Uff at the ground surface. The kinematic response factors (47) and (48) are plotted
versus the non-dimensional frequency factor a0 .
Figures 11(a) and (b) plot the kinematic response factors Iu and I for a single free-head pile.
The proposed model well fits the benchmark data.
Figure 12 shows the kinematic response factors obtained for a 2×1 pile group foundation.
Again, comparisons with the benchmarks demonstrate the validity of the proposed model: only in
the case of E p /E s = 10 000 slightly differences are evident for high values of the non-dimensional
frequency factor a0 , especially in the case of the response factor I .

Copyright q 2009 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2009; 38:1281–1305
DOI: 10.1002/eqe
A MODEL FOR THE 3D KINEMATIC INTERACTION ANALYSIS 1299

1.50 0.50
Free-head pile Free-head pile
Iu d

1.00
0.30 L
Proposed model
0.75 L/d = 20 Ep /Es = 1000
L/d = 40 Ep /Es = 1000
L/d = 20 Ep /Es = 10000 0.20
0.50 L/d = 40 Ep /Es = 10000
Fan et al.
L/d = 20 Ep /Es = 1000 0.10
0.25 L/d = 40 Ep /Es = 1000
L/d = 20 Ep /Es = 10000
L/d = 40 Ep /Es = 10000
0 0
0 0.1 0.2 0.3 a0 0.5 0 0.1 0.2 0.3 a0 0.5

Figure 11. Kinematic response factors Iu and I for single free-head pile.

1.20 0.10
Ep/Es = 1000
L/d = 20
Iu Iφ
0.80 Proposed model L s
s/d = 3 0.06
s/d = 5
0.60 s/d = 10
s/d = 20 d
s/d = 40 0.04
0.40 Fan et al.
s/d = 3
s/d = 5
0.20 s/d = 10 0.02
s/d = 20
s/d = 40
0 0
0 0.1 0.2 0.3 a0 0.5 0 0.1 0.2 0.3 a0 0.5
1.20 0.10
L/d = 20 Ep/Es = 10000
Iu

0.80 L s
Proposed model 0.06
s/d = 3
0.60 s/d = 5
s/d = 10
s/d = 20 d
s/d = 40
0.04
0.40 Fan et al.
s/d = 3
0.20
s/d = 5 0.02
s/d = 10
s/d = 20
s/d = 40
0 0
0 0.1 0.2 0.3 a0 0.5 0 0.1 0.2 0.3 a0 0.5

Figure 12. Kinematic response factors Iu and I for a 2×1 pile group in a homogeneous soil.

Figure 13 shows the results obtained from a pile group of two, three, four, six and nine piles in
a row. Only the values of the non-dimensional frequency factor a0 greater than 0.4 proved to be
insufficiently accurate; in the remaining frequency range the proposed model behaves very well.

Copyright q 2009 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2009; 38:1281–1305
DOI: 10.1002/eqe
1300 F. DEZI, S. CARBONARI AND G. LEONI

1.20 0.10
L/d = 20 Ep/Es = 1000
Iu s/d = 5

0.80 L s
0.06
Proposed model
0.60 Single pile
1x2
1x3
Fan et al. 0.04 d
1x4
0.40 1x6 Single pile
1x9 1x2
1x3
1x4 0.02
0.20 1x6
1x9

0 0
0 0.1 0.2 0.3 a0 0.5 0 0.1 0.2 0.3 a0 0.5

Figure 13. Kinematic response factors Iu and I for different pile groups in a homogeneous soil.

1.20 0.10
Ep/Es = 1000
L/d = 20
Iu Iφ

0.80
0.06 L s
0.60 Proposed model
s/d = 3
s/d = 5 0.04
0.40 s/d = 10 d
Fan et al.
s/d = 3 0.02
0.20 s/d = 5
s/d = 10
0 0
0 0.1 0.2 0.3 a0 0.5 0 0.1 0.2 0.3 a0 0.5

1.20 0.10
L/d = 20 Ep/Es = 10000
Iu Iφ

0.80
0.06 L s
0.60 Proposed model
s/d = 3
s/d = 5 0.04
0.40 s/d = 10
d
Fan et al.
0.20
s/d = 3 0.02
s/d = 5
s/d = 10
0 0
0 0.1 0.2 0.3 a0 0.5 0 0.1 0.2 0.3 a0 0.5

Figure 14. Kinematic response factors Iu and I for a 2×2 pile group in a homogeneous soil.

Finally Figure 14 shows the results obtained from a 2×2 pile group system. Here again the
proposed model behaves well and comparison reveals a slightly lower level of accuracy only for
values of a0 greater than 0.3.

Copyright q 2009 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2009; 38:1281–1305
DOI: 10.1002/eqe
A MODEL FOR THE 3D KINEMATIC INTERACTION ANALYSIS 1301

4. APPLICATION TO A REALISTIC FOUNDATION

To demonstrate the capacity of the model, the procedure is applied to a realistic foundation
consisting of a 4×4 pile group (Figure 15). The soil profile consists of a deformable soil layer
overlying a very dense sand deposit.
A full three-dimensional model is used to represent the geometry of the foundation system.
The piles are fully embedded in the soil, socketed into the sand deposit and connected at the
heads by a rigid massless cap. The concrete piles have a Young modulus of 3×107 kPa and a
density of 2.5 Mg/m3 . They are characterized by a length of 24 m, with circular cross sections
of 1 m diameter and a spacing of three diameters (center to center). The deformable deposit has
a depth of 18 m, a shear wave velocity Vs1 = 100 m/s and a density s1 = 1.5 Mg/m3 . The dense
sand deposit is characterized by shear wave velocity Vs2 = 800 m/s and density s2 = 2.5 Mg/m3 .
Poisson’s ratio is considered to be = 0.4 and material hysteretic damping = 10%, which is
compatible with the design level of strain in the soil. The soil elastodynamic Green’s functions,
as in the previous applications, were developed starting from the dynamic impedances and the
attenuation laws proposed by Dobry and Gazetas [4] and Makris and Gazetas [21, 28].
The foundation is modelled by discretizing each pile into 1 m long finite elements. A master
node is introduced at the centroid of the pile group at pile head level where the origin of the
reference system is also placed.
The seismic action defined at the outcropping bedrock consists of an artificial accelerogram
matching the EC8 Type 1 elastic response spectrum for ground type A. Seismic excitation is
assumed to act along the direction of the axis of symmetry x 1 .
The free-field displacements are evaluated by one-dimensional local response analysis. The
artificial accelerogram was linearly deconvoluted at bedrock level and then propagated through the

x2 0.25
3 x1 a [g]
1 2 1m
-0.25
0 10 t [s] 20
0 0.7
EC8 Type 1, ground A
Vs1 = 100 m/s Sa [g] Artificial accelerogram
z ρs1 = 1.5 Mg/m³ 0.5
ν = 0.4
= 10% 0.4
0.3
18 m
Vs2 = 800 m/s
ρs2 = 2.5 Mg/m³ 0.2
ν = 0.4
0.1
= 10%
25 m
0
Vsb > 800 m/s 0 1 2 T [s] 4
(a) (b)

Figure 15. (a) Pile group analyzed and (b) accelerogram and response spectrum at the outcropping bedrock.

Copyright q 2009 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2009; 38:1281–1305
DOI: 10.1002/eqe
1302 F. DEZI, S. CARBONARI AND G. LEONI

-1200 Mx2 [kNm] 1200 -400 Vx1 [kN] 400 -300 N [kN] 300
0
1
2
3

12

18

[m]

24
(a)
-1200 Mx2 [kNm] 1200 -400 Vx1 [kN] 400 -300 N [kN] 300
0

12

18

[m]

24
(b)

Figure 16. Stress resultants in the piles: (a) envelops of non-zero stress resultants due to
kinematic interaction and (b) non-zero stress resultants at the time step corresponding
to the maximum kinematic effect.

soil profile. Once the solution is obtained the displacements at the generic depth are computed in
the time domain by applying the Inverse Fourier Transform for each pile of the group.
Figure 16(a) shows the envelopes of the stress resultants due to kinematic interaction along three
piles which most represent the group behaviour, whereas Figure 16(b) shows diagrams relevant to
the time at which the maximum bending moment is attained. Considering the pile group geometry
and the master node chosen, the response of the system is uncoupled along x 1 , x2 and z. It is
worth mentioning that, even if no vertical action is applied to the system, axial forces arise in the
piles as a consequence of the rocking-translation coupling. In other words, axial forces develop to
balance the rigid cap subjected to the pile head bending moments.

Copyright q 2009 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2009; 38:1281–1305
DOI: 10.1002/eqe
A MODEL FOR THE 3D KINEMATIC INTERACTION ANALYSIS 1303

0 10 20 30 f [Hz] 50 0 10 20 30 f [Hz] 50
40⋅106 40⋅106
[kN/m] [kN/m]

30⋅106
30⋅106

Kv - ωCv
Kh - ωCh

25⋅106 25⋅106

20⋅106
20⋅106
15⋅106 Kv 15⋅106

10⋅106 ωCh ωCv 10⋅106


5⋅106
5⋅106
0 0
Kh
-5⋅10 6
-5⋅106
40⋅106 40⋅106
[kNm/rad] [kNm/rad]

30⋅10 6
30⋅106
Kr - ωCr

25⋅10 6
25⋅106

Kt - ωCt
Kr
20⋅106 20⋅106
ωCt
15⋅10 6
15⋅106

10⋅10 6 ωCr 10⋅106


6
5⋅106 5⋅10
Kt
0 0

-5⋅106 -5⋅106
12.0⋅106 1.2
1.0
[kN/rad]
0.8
90⋅105 0.6

Iu
0.4
Khr - ωChr

5
75⋅10 0.2
60⋅105 0

45⋅105 ωChr 0.05


0.04
30⋅105


Khr 0.03
0.02
15⋅105
0.01
0 0
0 10 20 30 f [Hz] 50 0 10 20 30 f [Hz] 50

Figure 17. Dynamic impedances of the soil–pile system and kinematic response factors.

The results reported in Figure 16 show that the corner piles are subjected to a higher level
of axial force, shear and bending moments. Pile-to-pile interactions are responsible for the slight
differences of shear forces and bending moments at the interface between the different soil layers.
More significant differences are evident in the axial force diagrams. This is due to the cap rocking
that produces forces proportional to the distance between the pile and the axis of rotation x 2 .
Finally, Figure 17 shows the non-zero components of the dynamic impedance matrix (46) and
the kinematic response factors of the foundation. Owing to the symmetry of the problem only five
different components are non-zero.

Copyright q 2009 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2009; 38:1281–1305
DOI: 10.1002/eqe
1304 F. DEZI, S. CARBONARI AND G. LEONI

5. CONCLUSIONS

A numerical procedure to perform the spatial analysis of the pile–soil–pile kinematic interaction
for pile group foundations is presented. The problem was formulated in the frequency domain by
considering the linear behaviour of the piles and the soil. This approach is valid for any pile group
geometry and soil profile. It may be used in the soil–structure interaction analysis of complex
structures according to the sub-domain method.
The convergence of the proposed numerical model was checked with respect to impedances,
kinematic response factors and pile stress resultants by varying the mesh size and considering
different geometry and soil conditions. The model was validated by comparing results with those
available in the literature obtained with more sophisticated methods.
The main features of the model are summarized below.
• The pile–soil–pile interaction and the radiation problem are accounted for by considering the
elastodynamic Green’s functions that permit the mutual interactions between all the piles of
the group consistently and simultaneously.
• The Green’s functions obtained by different formulations may be implemented.
• The model may be used to process generic free-field motions; this also permits studying cases
where particular 2-D or 3-D local response analyses are required to account for specific site
and topographic effects (e.g. basin effect).
• Rigid caps are modelled by constraining the displacements of the pile heads to those of a
rigid body (master node). The foundation impedances and the foundation input motion for
the inertial interaction analysis of generic spatial superstructures may easily be obtained by
condensing the problem.
• The model allows accurately evaluating the dynamic response of the soil–foundation system
and calculate the stress resultants due to the ground motion in each pile.

REFERENCES
1. Dezi F. Soil–structure interaction modelling. Ph.D. Thesis, Università Politecnica delle Marche, 2006.
2. Dezi F, Scarpelli G. Influence of soil–structure interaction on the seismic response of bridge piers. V Congresso
Nazionale dei Ricercatori di Ingegneria Geotecnica, Bari, 2006; in Italian.
3. Dezi F, Scarpelli G. Effect of soil–pile–structure interaction on the seismic design of the fixed abutment of a
railway bridge, V Congresso Nazionale dei Ricercatori di Ingegneria Geotecnica, Bari, 2006; in Italian.
4. Dobry R, Gazetas G. Simple method for dynamic stiffness and damping of floating pile groups. Geotechnique
1988; 38(4):557–574.
5. Kaynia AM, Kausel E. Dynamic stiffness and seismic response of sleeved piles. Report R80-12, MIT, Cambridge,
MA, 1982.
6. Mylonakis G. Contributions to static and seismic analysis of piles and pile-supported bridge piers. Ph.D. Thesis,
Faculty of the Graduate School of State University of New York, 1995.
7. Wolf JP. Dynamic Soil–structure Interaction. Prentice-Hall: Englewood Cliffs, NJ, 1987.
8. Wolf JP. Soil–Structure Interaction Analysis in Time Domain. Prentice-Hall: Englewood Cliffs, NJ, 1988.
9. Yegian M, Wright S. Lateral soil resistance–displacement relationships for pile foundations in soft clays.
Proceedings of the 5th Offshore Technology Conference, OTC 1893, Houston, vol. 2, 1973; 663–676.
10. Angelides DC, Roesset JM. Nonlinear dynamic stiffness of piles. Research Report R80-13, Department of Civil
Engineering, MIT, Cambridge, MA, 1980.
11. Randolph MF. The response of flexible piles to lateral loading. Geotechnique 1981; 31:247–259.
12. Faruque MO, Desai CS. 3-D material and geometric nonlinear analysis of piles. Proceedings of the Second
International Conference on Numerical Methods in Offshore Pilling, University of Texas at Austin, TX, 1982;
553–575.

Copyright q 2009 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2009; 38:1281–1305
DOI: 10.1002/eqe
A MODEL FOR THE 3D KINEMATIC INTERACTION ANALYSIS 1305

13. Trochanis A, Bielak J, Christiano P. A three-dimensional nonlinear study of piles leading to the development
of a simplified model. Report R-88-176, Department of Civil Engineering, Carnegie Institute of Technology,
December, 1988.
14. Trochanis A, Bielak J, Christiano P. Three dimensional nonlinear study of piles. Journal of Geotechnical
Engineering (ASCE) 1991; 117(3):429–447.
15. Wu G, Finn W. Dynamic elastic analysis of pile foundations using finite element method in the time domain.
Canadian Geotechnical Journal 1997; 34(1):44–52.
16. Bentley KJ, El Naggar MH. Numerical analysis of kinematic response of single piles. Canadian Geotechnical
Journal 2000; 37(6):1368–1382.
17. Sanchez-Salinero I, Roesset J. Static and dynamic stiffness of single piles. Report GR82-31, Civil Engineering
Department, University of Texas, Austin, 1982.
18. Sen R, Davis TG, Barnerjee PK. Dynamic analysis of piles and pile groups embedded in homogeneous soils.
International Journal for Earthquake Engineering and Structural Dynamics 1985; 13:53–65.
19. Novak M. Dynamic stiffness and damping of piles. Canadian Geotechnical Journal 1974; 11:574–597.
20. Flores-Berrones R, Whitman RV. Seismic response of end-bearing piles. Journal of the Geotechnical Engineering
Division (ASCE) 1982; 108(4):554–569.
21. Makris N, Gazetas G. Dynamic pile–soil–pile interaction—part II: lateral and seismic response. Earthquake
Engineering Structural Dynamics 1992; 21(2):145–162.
22. Kavvadas M, Gazetas G. Kinematic seismic response and bending of free-head piles in layered soil. Geotechnique
1993; 43(2):207–222.
23. Dobry R, O’Rourke MJ. Discussion on seismic response of end-bearing piles by Flores-Berrones R and
Whitman RV. Journal of the Geotechnical Engineering Division (ASCE) 1983; 109:778–781.
24. Nikolaou S, Mylonakis G, Gazetas G, Tazoh T. Kinematic pile bending during earthquakes: analysis and field
measurements. Géotechnique 2001; 51(5):425–440.
25. Mylonakis G. Simplified model for seismic pile bending at soil layer interfaces. Soils and Foundations 2001;
41(4):47–58.
26. Wheeler LT, Sternberg E. Some theorems in classical elastodynamics. Archive for Rational Mechanics and
Analysis 1968; 31(1):51–90.
27. Makris N, Badoni D, Delis E, Gazetas G. Prediction of observed bridge response with soil–pile–structure
interaction. Journal of Structural Engineering (ASCE) 1994; 120(10):2992–3011.
28. Makris N, Gazetas G. Displacement phase differences in a harmonically oscillating pile. Geotechnique 1993;
43(1):135–150.
29. Wolf JP. Foundation Vibration Analysis using Simple Physical Models. Prentice-Hall: Englewood Cliffs, NJ, 1994.
30. Mylonakis G, Gazetas G. Vertical vibrations and additional distress of grouped piles in layered soil. Soils and
Foundations 1998; 38(1):1–14.
31. Gazetas G, Dobry R. Horizontal response of piles in layered soils. Journal of Geotechnical Engineering 1984;
110:20–40.
32. Gazetas G, Dobry R. Single radiation damping model for piles and footings. Journal of Engineering Mechanics
1984; 110:937–956.
33. Fan K, Gazetas G, Kaynia A, Kausel E, Ahmad S. Kinematic seismic response of single piles and pile groups.
Journal of Geotechnical Engineering 1991; 117(12):1860–1879.
34. Bathe KJ, Wilson EL. Numerical Methods in Finite Element Analysis. Prentice-Hall Inc.: Englewood Cliffs, NJ,
1976.

Copyright q 2009 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2009; 38:1281–1305
DOI: 10.1002/eqe

Anda mungkin juga menyukai