Anda di halaman 1dari 15

Soil Dynamics and Earthquake Engineering 31 (2011) 891905

Contents lists available at ScienceDirect

Soil Dynamics and Earthquake Engineering


journal homepage: www.elsevier.com/locate/soildyn

Transient kinematic pile bending in two-layer soil


Stefania Sica a, George Mylonakis b, Armando Lucio Simonelli a,n
a b

 degli Studi del Sannio, Piazza Roma 21, 82100 Benevento, Italy Dipartimento di Ingegneria, Universita Department of Civil Engineering, University of Patras, University Campus, Rio 26500, Greece

a r t i c l e i n f o
Article history: Received 16 September 2010 Received in revised form 25 January 2011 Accepted 2 February 2011 Available online 31 March 2011 Keywords: Piles Soilstructure interaction (SSI) Kinematic interaction Numerical modelling

abstract
The dynamic response of piles to seismic loading is explored by means of an extensive parametric study based on a properly calibrated Beam-on-Dynamic-Winkler-Foundation (BDWF) model. The investigated problem consists of a single vertical cylindrical pile, modelled as an EulerBernoulli beam, embedded in a subsoil consisting of two homogeneous viscoelastic layers of sharply different stiffness resting on a rigid stratum. The system is subjected to vertically propagating seismic S waves, in the form of a transient motion imposed on rock outcrop. Several accelerograms recorded in Italy are employed as input motions in the numerical analyses. The paper highlights the severity of kinematic pile bending in the vicinity of the interface separating the two soil layers. In addition to factors already investigated such as layer stiffness contrast, relative soilpile stiffness, interface depth and intensity of ground excitation, the paper focuses on additional important factors, notably soil material damping, stiffness of Winkler springs and frequency content of earthquake excitation. Existing predictive equations for assessing kinematic pile bending at soil layer interfaces are revisited and new regression analyses are performed. A synthesis of ndings in terms of a set of simple equations is provided. The use of these equations is discussed through examples. & 2011 Elsevier Ltd. All rights reserved.

1. Introduction During the passage of seismic waves through soft deposits, embedded piles tend to deform in a different manner with respect to the surrounding soil. The difference between soil and pile motion depends on several factors, notably soil layering, pilesoil stiffness contrast, excitation frequency and kinematic constraints at the pile head and tip [14]. As curvatures are imposed to the pile body by the vibrating soil, bending and shearing will develop even in the absence of a superstructure. The associated pile bending moments are, thereby, referred to as kinematic, to be distinguished from those generated by loads acting at the pile head due to the dynamic response of the superstructure (so-called inertial moments). Kinematic and inertial bending moments constitute complementary aspects of a unique phenomenon known as soilpilestructure interaction (SPSI). Reviews of the subject have been published, among others, by Novak [5], Pender [6] and Gazetas and Mylonakis [7]. Evidence from case histories as documented by Mizuno [8] and other Japanese researchers [912] or from recent experimental investigations on physical pile models in centrifuge and 1g earthquake simulators [1318] have elucidated the important

role of kinematic interaction in seismic response of pile foundations. Kinematic bending is signicant (as compared to its inertial counterpart) particularly in correspondence to stiff pile caps and soil layer interfaces. The latter may explain the concentration of seismic demand at depths where inertial effects are negligible. The accumulated evidence has generated signicant interest in exploring theoretical and analytical aspects of the phenomenon and developing seismic regulations to incorporate it into design procedures [1921]. Following the early work by Margason and Halloway [22], theoretical investigations of the problem began in the 1980s [1,2327] and continued into the 1990s and beyond [24,2831]. In 2005, a systematic research effort was initiated in Italy under the auspicious of the ReLUIS project (University Network of Seismic Engineering Laboratories), which has lead to a number of publications [3243]. The main goal of the project was to produce engineering provisions to be incorporated into the new national seismic code [21], which is compulsory in Italy since July 2009. 1.1. Unresolved issues At present, it appears that many aspects of kinematic pile bending are well understood, whilst others require further research and remain unresolved. First, most of the published results concentrate on exible piles (i.e., piles whose lengths are greater than the so-called active pile length [44], embedded in

Corresponding author. E-mail addresses: stefsica@unisannio.it (S. Sica), mylo@upatras.gr (G. Mylonakis), alsimone@unisannio.it (A.L. Simonelli). 0267-7261/$ - see front matter & 2011 Elsevier Ltd. All rights reserved. doi:10.1016/j.soildyn.2011.02.001

892

S. Sica et al. / Soil Dynamics and Earthquake Engineering 31 (2011) 891905

two-layer soils with the interface placed deep beneath the surface [24,31]. Less research has been carried out on short piles and/or interfaces located close to the pile head. For such systems the interplay between pile head and interface moments, as well as between inertial and kinematic effects, becomes signicant and, thereby, a proper summation rule is essential for combining the contributions of the individual response components. Second, whereas a number of simplied procedures for estimating kinematic pile bending moments at the interface of two soil layers are available [24,4,3], there is a lack of simple formulations for assessing corresponding moments at the pile head regardless of layer thickness and pile length. Likewise, kinematic pile head moments have not been addressed in cases deviating signicantly from that of a homogeneous soil. Third, the effect of the transient nature of input motion on the development of kinematic bending moments along the pile requires further investigation. Available predictive equations are either pure frequency-domain approaches [3], or mixed frequencytime domain formulations based on a limited number of accelerograms [4,35]. Key parameters such as number of earthquake cycles and frequency content are currently accounted for in an approximate way, or even neglected altogether. Fourth, the effect of certain modelling assumptions involving soil damping, stiffness of Winkler springs and wave propagation on kinematic bending moments needs to be better quantied. Fifth, the effect of material nonlinearity in the soil (and the pile itself) needs to be explored in a more systematic way, despite recent progress [33,40,4550]. Last but not least, the effect of material plasticity for both soil and pile occurring under stronger earthquakes needs to be claried. The work at hand focuses only on some of the aforementioned issues. Specically it aims at reporting the results of an extensive parametric investigation carried out on single piles in layered, viscoelastic soil accounting for different material properties, geometric factors and earthquake excitations. The analyses are performed using an extension of the numerical tool developed by Mylonakis et al. [29] and Mylonakis [62], which implements a Beam-on-Dynamic-Winkler-Foundation (BDWF) formulation encompassing nite-element-based springs and dashpots distributed along the pile axis. Although simplied, this type of modelling is known to provide satisfactory accuracy as compared to rigorous numerical schemes, often allowing closed-form solutions to be obtained [1,3,4,24,43]. With reference to kinematic pile moments at soil layer interfaces, the paper presents: (i) a validation of the BDWF procedure against other available solutions; (ii) a sensitivity study of pile bending as function of spring stiffness adopted in Winkler models; (iii) a sensitivity analysis of the results as function of soil damping employed in the analysis of free-eld response; (iv) a comprehensive parametric study on a two-layer soil prole as function of stiffness contrast between the upper and the lower layer, the depth of the upper layer and the earthquake waveform. On the basis of these results, new regression analyses are carried out to correlate peak pile bending moments with the above parameters, with emphasis on frequency content of ground motion. Finally, a set of conclusions and recommendations are produced for implementing the ndings in routine engineering calculations.

Mylonakis et al. [29] and Mylonakis [62]. As the fundamental aspects of the method are well-known, only a brief description is given here. Soilpile interaction is modelled through a set of continuously distributed springs and dashpots, the parameters of which, k k(o) and c c(o), have been calibrated against the results of nite-element and boundary-element analyses. Such springs and dashpots connect the pile to the free-eld soil; the wave-induced motion of the latter serves as the support excitation for the pilesoil system. Mylonakis et al. [29] and Mylonakis [62] reformulated and extended earlier frequency-domain solutions [2] to the time domain using a Discrete Fourier Transform (DFT) approach as described by Veletsos and Ventura [51]. This can accommodate precisely the frequency dependence of the spring and dashpot moduli, contrary to methods that require frequency-independent parameters to obtain the response directly in the time domain. The approach has been validated in Mylonakis et al. [29], Mylonakis [3,62] and Nikolaou et al. [4,71] by comparing pile bending moments to those obtained by a variety of continuum approaches solved by FEM and BEM procedures. 2.1. Validation Recently, the research groups working as part of the Italian ReLUIS project conducted a comparative study of predictions provided by different numerical tools on conceptual prototypes consisting of single piles embedded in two-layer proles under the assumption of vertically propagating S waves (Fig. 1). The parameter analyses were carried out by varying the dimensionless interface depth (h1/d) while keeping the overall soil depth constant. The conguration adopted for the pile and the soil is shown in Fig. 1. The analyses were performed by adopting a suite of Italian accelerograms (Table 1) selected from the Italian database SISMA [52]; the recordings were scaled at a peak acceleration of 0.35 g, which is consistent with a zone of high seismicity according to the Italian seismic zonation of 2003 [53]. For the purposes of this validation the accelerograms were directly applied at the base of the soil column, as done by the other ReLUIS research groups, without de-convolution or consideration for rock outcrop effects. In Table 1, the frequency content of the accelerograms adopted as input motion is quantied through the predominant frequency, fp, corresponding to the

Fixed head

h1 Vs1, 1 D1, 1 EpI, p L

Vs2, 2 D2, 2 h2 d

2. Analysis method and validation The response to vertical S-wave excitation of a single vertical cylindrical solid pile embedded in layered soil is investigated through a Beam-on-Dynamic-Winkler-foundation (BDWF) formulation, solved using the hybrid analyticalnumerical algorithm of

Fig. 1. System considered.

S. Sica et al. / Soil Dynamics and Earthquake Engineering 31 (2011) 891905

893

Table 1 Ground motions employed in the analysis (all records are scaled to PGA 0.35 g). # Record label Station name Earthquake Date (d/m/yr) Magnitude (Mw) 6.5 6.5 6.9 6.9 6 5.5 5.5 5.1 5.6 5.6 5.5 5.7 5.7 5.3 5.3 5.3 5.1 5.1 Epicentral distance (km) 23 23 32 32 21 10 10 11 12 12 20 23 23 8 8 8 10 10 PGA (g) fp (Hz)
b

fm (Hz)

Soil type

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18
a b c

A-TMZ270 A-TMZ000 A-STU270 A-STU000 A-AAL018 E-NCB090 E.NCB000 R-NCB090 J-BCT000 J-BCT090 E-AAL108 B-BCT000 B-BCT090 TRT000 C-NCB000 C-NCB090 R-NC2090 R-NC2000

Tolmezzo-Diga Ambiesta Tolmezzo-Diga Ambiesta Sturno Sturno Assisi-Stallone Nocera Umbra-Biscontini Nocera Umbra-Biscontini Nocera Umbra-Biscontini Borgo-Cerreto Torre Borgo-Cerreto Torre Assisi-Stallone Borgo-Cerreto Torre Borgo-Cerreto Torre Tarcento Nocera Umbra-Biscontini Nocera Umbra-Biscontini Nocera Umbra 2 Nocera Umbra 2

Friuli Friuli Campano Lucano Campano Lucano Umbria Marche Umbria Marche a Umbria Marche a Umbria Marche a Umbria Marche a Umbria Marche a Umbria Marche a Umbria Marche Umbria Marche Friuli a Umbria Marche a Umbria Marche a Umbria Marche a Umbria Marche a

06/05/76 06/05/76 23/11/80 23/11/80 26/09/97 06/10/97 06/10/97 03/04/98 14/10/97 14/10/97 06/10/97 26/09/97 26/09/97 11/09/76 03/10/97 03/10/97 03/04/98 03/04/98

0.32 0.36 0.32 0.23 0.19 0.38 0.26 0.31 0.34 0.33 0.19 0.18 0.19 0.21 0.19 0.27 0.31 0.38

1.6 3.8 5.0 2.6 3.1 8.3 7.1 5.6 10.0 6.3 4.5 12.5 8.3 10.0 25.0 8.3 5.6 6.3

2.0 2.5 1.2 1.5 3.0 5.8 6.1 5.6 6.0 4.8 4.1 6.5 5.1 4.7 7.8 6.5 5.4 6.6

A A A A A A A A A A A A A A A A A A

Aftershock. Predominant frequency from response spectra. Average frequency according to Rathje et al. [54].

2.0

1.5 p x 10-3

BEM (Cairo & Dente, [32]) FEM (Di Laora, [56]) BDWF (Cairo et al., [33]) BDWF (Mylonakis et al., [29]) BDWF (Dezi et al., [42])

1.0

0.5

0.0 2.0

1.5 p x 10-3

1.0

0.5

0.0 0 5 10 15 h1/d
Fig. 2. Peak kinematic pile bending strain at interface level as function of depth for A-TMZ270 (top) and A-TMZ000 (bottom) Tolmezzo recordings; L 20 m, d 0.60 m, n1 n2 0.4, r1 r2 1.9 Mg/m3, D1 D2 10%, Vs1 100 m/s, Vs2 400 m/s, Ep 2.5 107 kPa and rp 2.5 Mg/m3.

20

25

30

35

maximum value of the acceleration response spectrum for 5% damping, and through the mean frequency, fm, as dened in [54] on the basis of frequency content. In Fig. 2 the results obtained by the BDWF method employed in this study are represented in terms of maximum pile bending strain: M d ep max Ep I 2 1

Mmax being the maximum kinematic bending moment at the interface; EpI the bending stiffness of the pile and d/2 the distance from the pile centerline to the outer bre of the pile cross section. ep is plotted as function of the dimensionless embedment factor h1/d. The use of bending strain over other normalisation schemes is desirable in such problems as [3]: (i) it is dimensionless; (ii) it is directly measurable; (iii) it can be used to quantify damage, as yield amplitudes do not vary signicantly among common structural materials (being of the order of 10 3); (iv) it can generate bi-dimensionless transfer functions, when normalised by pertinent strain parameters such as soil shear strain or site amplication functions. The results obtained with the selected approach are compared to the kinematic bending moments provided by two different BDWF formulations and two continuum solutions employed in the ReLUIS project. The rst BDWF approach has been developed by Dezi et al. [42] and was validated by detailed 3D nite-element analyses [43] using the ABAQUS platform [70]; other two approaches one continuum and another BDWF were developed by Cairo and Dente [32] and Cairo et al. [33]. The continuum approach [32] is implemented through a frequency-domain BEM technique that makes use of the soil stiffness matrices derived by Kausel and Roesset [55] to simulate the response of a horizontally layered deposit. The last approach [56] is implemented by means of the commercial nite-element code ANSYS [57] using very ne discretization. From Fig. 2 a satisfactory comparison is noted among the results provided by the aforementioned methods. Particularly satisfactory is the comparison between the BDWF approaches of Mylonakis et al. [29], Cairo et al. [33], Dezi et al. [42] and the two continuum approaches (FEM or BEM). Some discrepancies appearing in Fig. 2a in correspondence to the ratios h1/d 25 and 32 may be attributed to the selected values of Winkler springs, different discretizations of the pile, different damping schemes (e.g., Rayleigh versus linearly hysteretic), as well as different boundary conditions at the pile tip. Some of these factors are discussed in the ensuing. 2.2. Sensitivity of BDWF analyses on selection of Winkler springs In the BDWF approach the stiffness k of the springs connecting the pile to the soil is generally dened as k d Es 2

894

S. Sica et al. / Soil Dynamics and Earthquake Engineering 31 (2011) 891905

where Es is Youngs modulus of the soil and d is a dimensionless coefcient. Following the early work of Novak [58], different formulations have been proposed over the years for evaluating the d factor. For instance, using the results from the nite-element formulation of Blaney et al. [59], Roesset [60] recommended the value:

d 1:2

[42,61,62]. Improvements over the original formula have been presented, among others, by Dobry et al. [63] and Syngros [64]. With reference to kinematic pile bending in a two-layer deposit, Kavvadas and Gazetas [2] related d to pile and soil properties according to the equation:  1=8  1=8  1=12  1=30 3 Ep L h1 G2 d 4 d h2 G1 1n2 Es1 where n is the Poisson ratio of the soil material (common in both layers), L the pile length, h1 and h2 the thickness of the upper and lower layers, respectively, Ep the pile Youngs modulus, Es1 the Young modulus of the upper soil layers and G1, G2 the shear moduli of the upper and lower layers, respectively. Note that for a hollow cylindrical pile, Youngs modulus should be taken as Ep[1 (1 2t/d)4], t being the tube thickness. Mylonakis [3] simplied the above equation for the case of long piles in two-layer soil as follows:  1=8 E d6 p 5 Es1 The effect of Winkler parameter d on kinematic pile bending moments was investigated by means of an extensive parametric analysis. Fig. 3 presents results referring to accelerogram A-TMZ270 recorded during the 1976 Friuli earthquake. It can be observed that d has a negligible effect on kinematic bending everywhere, except for the vicinity of the layer interface. In that region an increase in d from 1.2 to 2.78 increases kinematic bending by 20% or so. A wider set of results referring to the same pilesoil conguration is presented in Fig. 4. There, kinematic bending moments computed for different ds are normalised with the results obtained for d 1.2, which was selected as a reference case. Naturally kinematic bending moment increases with increasing d, the increase being of the aforementioned order for d approaching 2.8. Based on these results the following approximate relation was derived:  1=40 Ep M d 0:97 0:17d 6 M1:2 Es1 This equation is in agreement with the solution of Dobry and ORourke [24] as to the dependence of pile bending on pilesoil stiffness contrast. Note that for soft piles (Ep/Es1 r 500) the effect

regardless of other problem parameters. This simple proposal was later adopted by several investigators, including the authors,

p x 10-3 0 0 = 1.20 (Roesset, [60]) 60 2.30 (Kavvadas & Gazetas, [2]) 5 2.78 2.78(Mylonakis, (Mylonakis,[3]) [ ]) 3 0.5 1 1.5 2

10

15 z/d 20

25

interface

30

35
Fig. 3. Effect of Winkler spring factor d on kinematic bending moments computed along the pile for the system of Fig. 1, subjected to A-TMZ270 input motion (case S1-6 in Table 2).

1.4
A-AAL018 A-STU000 A-STU270 ATMZ000 ATMZ270 BBCT000 B-BCT090 C-NCB000 C-NCB090 E-AAL108 ENCB000 ENCB090 J-BCT000 J-BCT090 R-NC2000 R-NC2090 R-NCB090 TRT000

Eq. 6 M () / M (1.2) 1.2 (Ep/Es1 = 500)


Kavvadas & Gazetas, [2]

0.8 0.8 1.2 1.6 2 2.4 2.8

Roesset, [60]

Mylonakis, [3]

3.2

Fig. 4. Effect of Winkler spring factor d on kinematic bending moments at a soil layer interface for pilesoil conguration S1-6 subjected to the suite of records in Table 1.

S. Sica et al. / Soil Dynamics and Earthquake Engineering 31 (2011) 891905

895

of d becomes more signicant. This is evident in Fig. 5, where results are presented for the geometry studied in the previous gures, subjected to A-STU000 and A-TMZ270 records.

3. Parametric investigation An extensive parameter study has been performed by the aforementioned BDWF approach for both free- and xed-head piles in a two-layer subsoil subjected to vertically propagating seismic waves. The soil prole consists of a soft surface soil layer of thickness h1 and shear wave velocity Vs1, followed by a stiffer stratum of thickness h2 and shear wave velocity Vs2 (Fig. 1). Elastic bedrock conditions have been assumed at the bottom of the lower layer. All cases considered are listed in Table 2. For each geometry, shear wave velocities of the upper and lower layers were selected in such a way that subsoil proles correspond to class C or D of the EC8 classication on the basis of the equivalent shear wave velocity Vs,30 in the upper 30 m of the subsoil. Three different values of soil damping were employed: 2%, 10% and 20%, which cover several cases of practical interest. Eighteen runs were carried out for each parametric case based on the input motions of Table 1. The accelerograms were chosen in such a way that their original peak ground acceleration is as close as possible to the reference maximum peak acceleration on soil type A of a seismic zone according to the Italian seismic zonation of 2003 [53]. For comparison purposes, the selected accelerograms have been scaled in amplitude to peak acceleration of 0.35 g, linearly de-convoluted to bedrock level and then propagated upward in the soil to provide the excitation motion of the embedded pile. In Table 2 the ratio f1/fp between the fundamental frequency of the subsoil f1 and the predominant frequency of input motion fp is provided. In all analyses the following parameters

1.4 Eq. 6 from Mylonakis, [3]

1.2

M () /M (1.2)

Eq. 6 from Kavvadas & Gazetas, [2] 1 from Roesset, [60] A-TMZ270 A-STU000 0.8 0 100 200 300 400 500 600 Ep/Es1 700 800 900 1000 1100 1200

Fig. 5. Effect of Winkler stiffness factor d and Ep/Es1 ratio on interface kinematic bending moments, for pilesoil conguration S1-6 subjected to A-TMZ270 and A-STU000 input motions.

Table 2 Problem parameters considered. In all cases n1 n2 0.4, L/d 33, Vrock 1000 m/s. Scheme h1 (m) h2 (m) Vs1 (m/s) Vs2/Vs1 2 3 4 2 3 4 2 2.7 4 2 3 4 2 2.7 4 2 3 4 2 2.7 4 2 3 4 2 3 4 2 3 4 ID parameter case S1-1 S1-2 S1-3 S1-4 S1-5 S1-6 S1-7 S1-8 S1-9 S2-10 S2-11 S2-12 S2-13 S2-14 S2-15 S3-16 S3-17 S3-18 S3-19 S3-20 S3-21 S4-22 S4-23 S4-24 S5-25 S5-26 S5-27 S6-28 S6-29 S6-30 Vs, 66 75 80 133 150 160 200 218 240 133 150 160 200 218 240 160 169 174 235 245 255 171 225 267 150 180 200 122 132 138
30

(m/s)

Soil type (EC8) D D D D D D C C C D D D C C C D D D C C C D C C D C/D C D D D

Ep/Es1

f1 (Hz) 0.7 0.8 0.8 1.3 1.5 1.6 2.0 2.2 2.4 1.1 1.3 1.5

f1/fp, range 0.030.42 0.030.49 0.030.50 0.050.86 0.060.99 0.061.03 0.081.28 0.091.42 0.101.54 0.040.67 0.050.86 0.060.97 0.061.00 0.081.22 0.091.47 0.060.99 0.071.05 0.071.08 0.091.47 0.101.53 0.101.58 0.061.03 0.101.57 0.132.05 0.061.02 0.081.28 0.091.45 0.050.74 0.050.80 0.050.84

50

1880

S1

15

15

100

470

150

209

100 S2 15 30 150

470

209

2% 10% 20%

1.6 1.9 2.3 1.5 1.6 1.7 2.3 2.4 2.5 1.6 2.5 3.2 1.6 2.0 2.3 1.2 1.3 1.3

100 S3 15 6 150

470

209

S4

25

100

S5

10

20

100

470

S6

19

11

100

896

S. Sica et al. / Soil Dynamics and Earthquake Engineering 31 (2011) 891905

M (kN*m) 0 0 VS1/ VS2 = 1/2 5 12 24 (1.9 % A) 12 30 (3.0 % A) 200 400 600 800 1000

10

8 16 (0.6 % A)

z (m)

p = 0.8 %o

20 0

p = 0.5 %o

15

p = 1.1 %o

VS1/VS2 = 1/3 5
A-AAL018 A-STU270 A-STU000 ATMZ000 B-BCT000 C-NCB000 E-AAL108 E-NCB090 J-BCT090 R-NC2090 TRT000

z (m)

ATMZ270

10

B-BCT090 C-NCB090 E-NCB000

15

J-BCT000 R-NC2000 R-NCB090

20 0 VS1/VS2 = 1/4 5

z (m)

10

15

20
Fig. 6. Effect of shear wave velocity contrast on kinematic bending moments for pilesoil conguration S1. From top to bottom: parameter cases S1-4, S1-5 and S1-6 (subsoil D) of Table 2. Grey bands dene yield resistance of pile cross sections for different amounts of longitudinal reinforcement.

were kept constant: pile length (L 20 m), pile diameter (d 0.60 m), soil Poissons ratio (n1 n2 0.4), soil mass density (rs 1.9 Mg/m3), pile Youngs modulus Ep 2.5 107 kPa and pile density rp 2.5 Mg/m3. The shear wave velocity of the elastic bedrock, Vrock, was taken equal to 1000 m/s. 3.1. Effect of geometry and soil impedance contrast Typical results from the parameter study are shown in Figs. 68, where envelopes of maximum kinematic bending moments along the pile are plotted for each of the 18 signals. In all gures, three

zones shown in grey colour are indicated, corresponding to the range of yield moments for typical reinforcements of the concrete pile cross section (8F16, 24F12 and 12F30) and variable normal loads at the pile head. MN interaction diagrams were computed assuming concrete class C20/25 with Rck 25 N/mm2 and fck 20 N/ mm2 and steel rebars with fyk 375 N/mm2 and ftk 450 N/mm2, corresponding to the Italian steel class FeB38K. For each reinforcement conguration, the lower limit of the grey zone represents the cross sectional yielding moment corresponding to zero normal load while the higher one to a typical normal load N 1200 kN [normalised axial load nd N/(fckA) between 0 and 0.32].

S. Sica et al. / Soil Dynamics and Earthquake Engineering 31 (2011) 891905

897

M (kN*m) 0 0 VS1/ VS2 = 1/2 5 12 24 (1.9 % A) 12 30 (3.0 % A) 200 400 600 800 1000

10

8 16 (0.6 % A)

z (m)

p = 0.8 %o

20 0

p = 0.5 %o

15

p = 1.1 %o

VS1/VS2 = 1/3 5
A-AAL018 A-STU270 A-STU000 ATMZ000 B-BCT000 C-NCB000 E-AAL108 E-NCB090 J-BCT090 R-NC2090 TRT000

z (m)

ATMZ270

10

B-BCT090 C-NCB090 E-NCB000

15

J-BCT000 R-NC2000 R-NCB090

20 0 VS1/VS2 = 1/4 5

z (m)

10

15

20
Fig. 7. Effect of shear wave velocity contrast on kinematic bending moments for pilesoil conguration S1. From top to bottom: parameter cases S1-7, S1-8 and S1-9 (subsoil C) of Table 2. Grey bands dene yield resistance of pile cross sections for different amounts of longitudinal reinforcement.

The following observations may be drawn from these gures:

 Interface bending moments increase dramatically when the


shear wave velocity contrast between the upper and lower layers Vs1/Vs2 decreases from 1/2 to 1/4. This occurs for both subsoil types D (Fig. 6) and C (Fig. 7), in agreement with earlier research. In the latter case, however, peak moments at interface and pile head are much smaller than those computed for subsoil type D. Maximum bending moment develops close to the interface when the active pile length la is smaller than the thickness h1

of the rst layer. This length can be determined from the following empirical formula as  0:25 Ep 7 a wd Es1 where w is a dimensionless constant varying between 1.75 and 2.5 [24,29,44,62,64]. Conversely, when the active pile length a is equal or smaller than h1, the kinematic bending moment at the pile head may exceed that of the interface (Fig. 8). For subsoil proles corresponding to class D (Fig. 6) the computed kinematic bending moments may be well above

898

S. Sica et al. / Soil Dynamics and Earthquake Engineering 31 (2011) 891905

M (kN*m) 0 0 VS1/ VS2 = 1/2 5 12 24 (1.9 % A) 12 30 (3.0 % A) 200 400 600 800 1000

10

8 16 (0.6 % A)

z (m)

p = 0.8 %o

20 0

p = 0.5 %o

15

p = 1.1 %o

VS1/VS2 = 1/3 5
A-AAL018 A-STU270 A-STU000 ATMZ000 B-BCT000 C-NCB000 E-AAL108 E-NCB090 J-BCT090 R-NC2090 TRT000

z (m)

ATMZ270

10

B-BCT090 C-NCB090 E-NCB000

15

J-BCT000 R-NC2000 R-NCB090

20 0 VS1/VS2 = 1/4 5

z (m)

10

15

20
Fig. 8. Effect of shear wave velocity contrast on kinematic bending moments for pilesoil conguration S4. From top to bottom: parameter cases S4-22 (subsoil D), S4-23 and S4-24 (subsoil C) in Table 2. Grey bands dene yield resistance of pile cross sections for different amounts of longitudinal reinforcement.

the yielding moments for typical concrete reinforcements and normal loads acting along the pile. As the analyses have been performed assuming linear elastic behaviour for the materials, the computed moments can be assumed valid until the yield limit is reached. For subsoil proles corresponding to soil class C and shear wave velocity contrasts Vs2/Vs1 4 2, kinematic bending moments may also be considerable, especially for low levels of pile reinforcement (Fig. 7).

3.2. Effect of frequency content of input motion For a better interpretation of the analytical results, the ratio f1/finput between the fundamental natural frequency of the subsoil, f1, and the predominant frequency of input motion, finput, has been established for all cases. finput has been determined on the basis of both response spectrum predominant frequency, fp, and average frequency, fm, as described in [54] (Table 1). In Fig. 9 results pertaining to case S1-6 (Table 2) are shown. It is evident that the

S. Sica et al. / Soil Dynamics and Earthquake Engineering 31 (2011) 891905

899

2 finput = fm 1.5 finput = fp

f1 / finput

possible resonance

0.5

no resonance

0 A-AAL018 E-AAL108 ATMZ000 ATMZ270 C-NCB090 E-NCB090 B-BCTT000 R-NCB090 A-STU000 A-STU270 B-BCT090 J-BCT000 J-BCT090 R-NC2000 C-NCB000 E-NCB000 R-NC2090 TRT0000

Input motion
Fig. 9. Ratio of natural frequency of soil prole, f1, to the predominant frequency of input motion, finput, for case S1-6 (Table 2) and the signals of Table 1.

accelerograms that induce the highest kinematic moments are characterised by values of f1/fp or f1/fm close to unity, a trend obviously related to the development of resonant phenomena in the soil. This is in accordance with the ndings in [2,4] by frequencytime domain analyses: maximum effects of kinematic bending in piles occur when the frequency of ground motion is near the fundamental period of the subsoil (f1/finput E 1). Conversely, all the accelerograms that induce low kinematic moments in the pile have f1/finput below 0.5. These results suggest that a threshold value of ratios f1/finput can be dened beyond which kinematic effects are important (Fig. 9). This nding may provide a practical way for selecting records from a database of seismic events for a particular region: if acceleration time-histories (in addition to peak ground accelerations) are available as part of the design process, the potential of developing signicant kinematic pile bending can be established on the basis of the following criterion: - if ratio f1/finput is lower than the threshold value, kinematic bending will typically be minor and only inertial interaction can be accounted for in design; - if ratio f1/finput is higher than the threshold value, the potential for kinematic bending will be high and kinematic interaction should be considered, in addition to inertial interaction.

M (D) / M (10%)

1.5

Eq. 8

0.5

0 0 5 10 D (%) 15 20

Fig. 10. Effect of soil material damping on kinematic pile bending moments. Results refer to case S1-6 and the input motions of Table 1.

1400 1200 M (kNm) 1000 800 600 400 200 0 0 1 f1 2 fc 3 4 finput (Hz) 5 6 7 Mstatic Mresonance

3.3. Effect of soil damping The effect of soil damping on the response is examined in Fig. 10, which provides a comparison among the kinematic bending moments computed at layer interface for different levels of material damping in the soil (2%, 10% and 20%). The reported results pertain to parameter case S1-6 (Table 2). It is observed that soil damping can signicantly affect the magnitude of kinematic bending moments, as it controls free-eld response and it should be carefully assigned when linear elastic analyses are carried out. Based on these results, the following regression formula was derived: M D 5 D1=4 M 10% 9 8

Fig. 11. Relation between transient and steady-state interface kinematic bending moments for conguration S1-6 and the input motions of Table 1 (finput fm). Note the critical frequency fc beyond which maximum moment is smaller than static.

which is reminiscent of corresponding expressions in seismic regulations for structural problems [19,20] and can be used to quantify the effect of damping in preliminary design calculations. Naturally, kinematic response tends to drop with increasing

900

S. Sica et al. / Soil Dynamics and Earthquake Engineering 31 (2011) 891905

= maximum possible moment Mresonance 1 =

M (t)max Mresonance M (t)max Mstatic

1.0 0.8 0.6 0.4 0.2 mean + 1 mean mean-1 = 0.68 finput f1

1.5

maximum transient moment for a given input motion M(t)max moment for = 0 Mstatic

0.0 resonance range 3.0


0 static resonance 1

2.5
finput f1

1 = 1.94

finput f1

1.3

2.0 1 1.5 1.0 0.5 mean + 1 mean mean-1

Fig. 12. Denition of response factors Z and F1 for a given pilesoil conguration.

material damping, the decrease being of the order of 20% for an increase from 5% to 10%. 3.4. Frequency versus time domain bending

0.0
In Fig. 11 interface kinematic bending moments determined in the frequency domain are compared to corresponding moments computed in the time domain for each selected accelerogram, referring to parameter case S1-6 (Table 2). Time-domain moments naturally lie below the value associated with resonance condition (Mresonance). Interestingly, for most signals time-domain moments are lower than the static value Mstatic obtained for zerofrequency excitation. This occurs because the input accelerograms are characterised by much higher dominant frequencies (48 HzTable 1), than the fundamental frequency of the subsoil at hand (f1 1.6 Hz). For this frequency range, de-amplication is observed and kinematic effects consequently diminish. An enlarged version of Fig. 11 is schematically represented in Fig. 12 to dene the following factors correlating frequencyand time-domain results:

10

11

12

finput / f1
Fig. 13. Factors Z (top graph) and F1 (bottom graph) versus frequency ratio (finput/f1) for different pilesoil congurations and input motions. Only cases corresponding to long piles are shown.

(mean+ s) and lower (mean s) estimates, s being the standard deviation. With reference to the mean value, the following simple regression relations were derived:  1:5 f Z 0:68 input 11 f1

M t max Mresonance Mt max Mstatic

F1 1:94
9

 1:3 finput f1

12

F1

10

which were rst employed in Nikolaou et al. [4,71] and Mylonakis [3], respectively. In these equations, parameter Z can be interpreted as a de-amplication factor, bringing the resonance bending moment, Mresonance, to match the peak time-domain moment, M(t)max. [As an example, spectral amplication of 2.5 in conventional design spectra versus a peak resonant amplication of 1/(2D) 10 for 5% damping, is equivalent to an Z factor of 2.5/10 0.25 for a simple oscillator.] On the other hand, parameter F1 may be interpreted either as an amplication or as a de-amplication factor to bring the bending moment computed for static conditions, Mstatic, to match the peak timedomain moment, M(t)max. In the ensuing and except if specically otherwise indicated, factors Z and F1 in Eqs. (9) and (10) are determined from a numerical BDWF analysisnot from approximate expressions such as those provided in the original publications. In Fig. 13, the mean values of a numerical regression analysis performed for Z and F1 are displayed with corresponding upper

which are valid for long piles and finput/f1 ratios greater than approximately 1.5. Eqs. (11) and (12) can be used to quantify the effect of transient nature of input motion on kinematic response of piles. Naturally, both parameters tend to drop with increasing finput/f1 that is for conditions far from resonance. For finput/f1 E 1 the results exhibit signicant dispersion. In that frequency range, response is known to depend on number of excitation cycles [4], an effect which is not captured by Eqs. (11) and (12). For this reason, dashed lines have been employed in Fig. 13 for 0.5 o finput/f1 o 1.5.

4. Simple formulas for transient pile bending at a layer interface Mylonakis [3] developed a simple formulation for predicting kinematic bending moments at a layer interface under lowfrequency excitation (o-0). The kinematic bending moment may be derived from a strain transmissibility parameter, ep/g1, which is simply the ratio between peak pile bending strain, ep, and free-eld soil shear strain at the interface, g1. The corresponding

S. Sica et al. / Soil Dynamics and Earthquake Engineering 31 (2011) 891905

901

analytical solution is )  1 ("  1=4   #   ep 1 h1 k1 h1 1 cc11 4 c2 c 1 3 g1 o 0 2c d Ep d 13 where c (G2/G1) is the layer stiffness contrast, (h1/d) the embedment ratio, k1 the Winkler spring modulus associated with the upper layer. As Eq. (13) has been obtained for static conditions, it does not account for the transient, dynamic nature of the phenomenon. To address this limitation, Mylonakis [3] introduced a correction function F of the form:     e ep F p 14

3.0 2.5 2.0 Vs1/Vs2 = 1/4 1/3

1.5 1.0 0.5 0.0 3.0 2.5 2.0 Vs1/Vs2 = 1/2

g1

dyn

g1

o0

which accounts for the effect of frequency on kinematic pile bending. It should be noticed that, although conceptually related, F in Eq. (14) is not identical to F1 in Eq. (10). Indeed, whereas Eq. (10) relates static to transient response accounting for all relevant dynamic phenomena, Eq. (14) refers exclusively to SSI effects. For relatively soft piles and low-frequency input (oinput d/Vs1 o 0.1), reference [3] recommends that parameter F lies in the range 11.25. To investigate frequency-to-time domain response relations, Maiorano et al. [35] adopted the kinematic bending moment M(t)max computed in time domain by the FEM code VERSATP3D [65] and the shear strain at the interface (g1)dyn provided by a one-dimensional EERA analysis [66]. For this scope they re-arranged Mylonakis [3] solution in the form:

1.5 1.0 0.5 0.0 0 1 2 3 4 5 6 7 8 9 10 11 12 finput / f1

F2

2Ep I ep g d g1 o 0 1 dyn

M t   max

15

for which symbol F2 is employed in this article to avoid confusion with the denitions provided earlier. Contrary to factor F1 in Eq. (10) F2 is associated with: (i) the peak dynamic strain (g1)dyn, as computed from a 1-D wave propagation analysis [67]; (ii) the approximate solution for (ep/g1) given by Eq. (13). Accordingly, F2 relates the exact transient moment with an approximate moment that is neither purely static nor purely dynamic, referring partially to dynamic and SSI effects. By linear regression analysis, Maiorano et al. [35] found that parameter F2 is approximately 1.30 for conditions far from resonance and 1.40 near resonance. Using nonlinear regression analysis based on the numerical data obtained in this study, the parameter F2 was determined to be approximately 1.20 for layer stiffness contrast Vs1/Vs2 in the range 1/41/3 (Fig. 14). This value is close to the generic recommendation by Mylonakis [3]. Higher values of 1.3 on average were obtained for lower stiffness contrast Vs1/Vs2 1/2 (Fig. 14). In this case, however, the kinematic bending moment at the layer interface is not expected to be important. It is noteworthy that F2 is practically independent of (finput/f1). This is understood given that both M(t)max and (g1)dyn, in Eq. (15), encompass the frequency dependence of the free-eld response, which, thereby, cancels out from the ratio. As an alternative to the mechanistic model by Mylonakis [3], Nikolaou et al. [4,71] introduced the following empirical equation to compute the pile kinematic bending moment at the interface in resonant steady-state conditions:  0:30  0:65  0:50 Ep L Vs2 Mresonance 0:042tc d3 16 d E1 Vs1 where tc is a characteristic shear stress at the interface:

Fig. 14. Factor F2 versus frequency ratio (finput/f1) for pilesoil congurations S1 and S5. Only cases corresponding to long pile are shown; Vs1/Vs2 1/41/3 (top graph) and Vs1/Vs2 1/2 (bottom graph).

transient kinematic bending moment with the value obtained for resonant conditions, from Eq. (16). A theoretical weakness of the above formula is that the predicted moment tends to increase without bound for very slender piles, large pilesoil stiffness contrasts and large velocity contrasts between the two soil layers, a trend that has not been observed. In the same spirit as before, Maiorano et al. [35] re-arranged the Nikolaou et al. [4,71] equation in the form:

M t max G1 b g1 dyn

18

where parameter b is given by b d3  0:3  0:65  0:5 Ep L Vs2 d Es1 Vs1 19

tc as r1 H1

17

in which as is the free eld acceleration at soil surface. The authors introduced a parameter Z (Eq. (9)), to correlate the

In this approach, b can be interpreted as an average dynamic coefcient that links the maximum transient bending moment to the transient peak soil shear strain at the interface, (g1)dyn. Maiorano et al. [35] report b to be about 0.07 on average. Recall that parameter b is not equivalent to parameter Z in Eq. (9), for the denominator in Eq. (18) is not an exact peak steady-state moment. In the present work the procedure followed by Maiorano et al. [35] was employed to estimate the parameters F2 and b by adopting the M(t)max provided by the BDWF approach of Mylonakis et al. [29]. Results are shown in Fig. 15. It was found that F2 is equal to approximately 1.15 whereas b is around 0.053. Both these estimates are lower than those reported by Maiorano et al. [35]. The reasons for this discrepancy are worth investigating, yet lie beyond the scope of this paper.

902

S. Sica et al. / Soil Dynamics and Earthquake Engineering 31 (2011) 891905

2.5

2.0 (Mmax/a) x 102

mean+1 mean

2) Dynamic approach based on F2 The procedure by Mylonakis [3] may be written in the alternative form as   2Ep I ep g F 22 M t max d g1 o 0 1 dyn 2 where the static shear strain (g1)o 0 has been replaced by its dynamic counterpart (g1)dyn; (ep/g1)o 0 is given by Eq. (13) and dynamic modier F2 is equal to about 1.201.25 (Fig. 14). This approach is also attractive as dynamic effects due to frequency content of the input motion are incorporated into parameter (g1)dyn, whereas F2 is a constant. A disadvantage is that a free-eld site response analysis is required to establish the value of (g1)dyn. 3) Dynamic approach based on Z The original formulation by Nikolaou et al. [4,71] can be rewritten in the form: M t max Z Mresonance 23

1.5

mean-1

1.0

0.5 Mmax / a = 1.15 (1)dyn 5.110-4 0.0 1.2 1.0 (Mmax/bG1) x 103 0.8 0.6 0.4 0.2 0.0 0 0.005 Mmax / b = 0.053 (1)dyn 2.8 10-5 0.01 0.015 (1) dyn
Fig. 15. Graphical representation of peak kinematic bending moments using two different normalisation schemes (modied from [35]).

0.02

0.025

where Z can be determined from Eq. (11) and Fig. 13; Mresonance may be obtained from Eqs. (16) and (17). To this end, knowledge of surface acceleration as at resonance is required. The latter can be determined by a harmonic site response analysis or by a fully analytical approach assuming harmonic wave propagation in a two-layer subsoil. Cairo et al. [69] suggested a simple method to obtain the peak steady-state acceleration at ground surface as, based on the peak acceleration ar on soil type A and the fundamental period T1 of the subsoil. In this way, no numerical site response analysis is necessary and the approach can be fully analytical. As in the case with parameter F2, Z is sensitive to variations of frequency ratio (finput/f1). Numerical examples are provided in Appendix I.

5. Synthesis: new proposed approaches As previously illustrated, different simple approaches may be followed for determining the transient kinematic pile bending at the interface of two soil layers. Three such approaches are suggested in this paper. One is purely static and, accordingly, does not require dynamic analysis; the other two are dynamic and, thereby, are more demanding from a computational viewpoint. 1) Static approach based on F1 This is the simplest procedure and can be implemented without carrying out a dynamic analysis. To this end, the procedure by Mylonakis [3] is reformulated as   2Ep I ep g F 20 M t max d g1 o 0 1 o 0 1 where (ep/g1)o 0 is provided by Eq. (13) and F1 is given by Eq. (12) and Fig. 13; (g1)o 0 is the static shear strain at the interface, which can be easily established by the following equation: g1 o 0 a h1 2 Vs 1 21

6. Conclusions Results from an extensive parametric analysis were reported, carried out on single vertical elastic solid piles in layered soil, accounting for different material properties, geometric factors and earthquake excitations. The analyses were performed using a numerical tool developed by Mylonakis et al. [29] and Mylonakis [62] based on a properly calibrated Beam-on-Dynamic-Winkler-Foundation (BDWF) model. The main conclusions of the study are: 1. A sensitivity analysis of pile bending moments as function of spring stiffness k adopted in Winkler models and damping ratios D employed in free eld response analyses was carried out; Eqs (6) and (8) were developed to account for these effects in a simple manner. 2. A comprehensive parametric study on a two-layer prole as function of: (i) stiffness contrast between upper and lower layer, (ii) depth of upper layer, (iii) waveform of input motion and (iv) soil damping adopted in the free-eld site response analysis, was reported. It was found that: a) Severe kinematic interaction develops at the interface of two soil layers having sharply different stiffness and pertaining to subsoil of class C and D according to EC8 classication. In these cases the computed kinematic bending moments at the interface may be well above the yielding moments of the pile cross section computed for typical reinforced concrete piles and axial loads. b) A threshold value of ratio (f1/finput) can be determined beyond which kinematic effects are considerable. This limit provides a possible criterion for selecting signicant

where a as ar is a uniform pseudostatic seismic acceleration in the prole. Note that Eq. (21) is equivalent to the familiar Seed and Idriss [68] procedure with the depth factor rd taken equal to 1. This approach is attractive for engineering applications as it does not require a free-eld site response analysis. A disadvantage is that parameter F1 is sensitive to variations of (finput/f1) ratio, which is often not known with sufcient accuracy in practice. Adjustments due to damping and spring stiffness can be made with the help of Eqs. (6) and (8).

S. Sica et al. / Soil Dynamics and Earthquake Engineering 31 (2011) 891905

903

records from a ground motion database. Therefore, if the acceleration time-histories are provided for a given seismic zone (in addition to conventional peak ground acceleration), the potential of the site with the associated waveforms to induce signicant kinematic bending in piles may be established. 3. On the basis of the above results, new regression analyses were carried out for computing the transient pile bending moments at the soil layer interface. Three alternative procedures were outlined to solve the problem in the realm of routine engineering calculations. Specically: a) The approach based on F1 is perhaps the most attractive for engineering purposes, as it does not require a free-eld site response analysis. A disadvantage is that parameter F1 is sensitive to variations of (finput/f1) ratio, which is often not known with sufcient accuracy in practice. b) The approach based on F2 is also attractive as dynamic effects due to frequency content of the input motion are incorporated into parameter (g1)dyn, whereas F2 is a constant. A potential disadvantage is that a free-eld site response analysis is required to establish the value of (g1)dyn. c) The approach based on Z requires knowledge of Mresonance, which can be realized by means of Eq. (16). As in the case of parameter F1, Z is sensitive to variations of frequency ratio (finput/f1). 4. Large differences often appear in the parameters representing the frequency content of input motion (fp or fm, Table 1). For the sake of conservatism, the authors suggest to apply the simplied formulas choosing between fp and fm (or any other rational estimate) the value closer to the fundamental frequency f1 of the soil prole. Acknowledgements The work herein described is part of the ReLUIS research Project Innovative methods for the design of geotechnical systems, promoted and funded by DPC (Civil Protection Department) of the Italian Government and coordinated by the AGI (Italian Geotechnical Association). The authors wish to thank ReLUIS and AGI research coordinators.

From Eq. (21) the static shear strain is likewise obtained: g1 o 0 0:35 9:81 15 1002 0:00515 A:2

To account for the transient nature of the phenomenon, frequency parameter F1 needs to be determined by means of Eq. (12). For the case in hand, the ratio (finput/f1) is equal to 2.4. Applying Eq. (12), parameter F1 is estimated at 0.622. Finally, the transient kinematic bending moment is calculated from Eq. (20):
Mt max 2 2:5 107 0:00636 0:115 0:00515 0:622 195 kNm 0:6

A:3 which is almost identical to the exact value of 194 kN m.

Dynamic approach based on F2 In this approach a free-eld site response analysis should rst be carried out to obtain (g1)dyn. For the problem at hand an EERA analysis provided the value (g1)dyn 0.0029, which, remarkably, is lower than the static estimate in Eq. (A.2). Assigning F2 the value 1.20, as suggested in this paper, Eq. (22) provides
M t max 2 2:5 107 0:00636 0:115 0:0029 1:20 220 kN m 0:6

A:4 which is about 10% higher than the exact value.

Dynamic approach based on Z To apply this approach, knowledge of surface acceleration as at resonance is required. This may be obtained from a free-eld site response analysis, or by a fully analytical approach [69]. For the case in hand, as is found to be equal to 1.5 g. From Eq. (17), we have

tc as r1 H1 1:5 9:81 1:9 15 419 kPa


From Eq. (16) the estimate of the resonant moment is Mresonance 0:042 419 !0:65    20 0:30 2:5 107 400 0:50 1187kN m 0:6 100 53 200

A:5

Appendix I. Numerical examples The vertical pile shown in Fig. 1 is embedded in soil prole S1-4, subjected to Italian records A-TMZ000 and A-AAL018 (Tables 1 and 2), scaled at a peak ground acceleration of 0.35 g. In the ensuing the following parameters are considered: L 20 m, n1 n2 0.4, rs 1.9 Mg/m3, Ep 2.5 107 kPa, d 0.60 m, 3 rp 2.5 Mg/m , D 10%. The shear wave velocity of the elastic bedrock, Vrock, was taken at 1000 m/s. At the layer interface a BDWF analysis according to reference [29] provides maximum kinematic bending moments in the time domain equal to 194 and 248 kN m for the A-TMZ000 and A-AAL018 records, respectively. These values were computed assuming a spring stiffness d equal to 1.2. The three simplied approaches proposed in the paper will be hereafter applied to the above case. Record A-TMZ000

0:6

A:6

For (finput/f1) 2.4, Eq. (11) yields Z 0.18. Finally, applying Eq. (23): Mt max ZMresonance 0:18 1187 216 kN m A:7

Following the Maiorano et al. [35] approach, with the parameter b 0.053 obtained in this study:
M tmax bbG1 g1 dyn 0:053 67:5 19000 0:0029 197kN m

A:8 Both these results are close to the exact value of 194 kN m. Additional adjustments to the above estimates can be made with reference to d and D values (Eqs. (6), (8)), if required.

Record A-AAL018 Static approach based on F1 The strain transmissibility is computed from Eq.(13) as   ep 0:115 A:1 Static approach based on F1 Repeating the above analysis yields the same strain transmissibility (ep/g1)o 0 and static shear strain (g1)o 0 as before. For a ratio (finput/f1) 1.9, F1 0.842 (Eq. (12)).

g1

o0

904

S. Sica et al. / Soil Dynamics and Earthquake Engineering 31 (2011) 891905

The transient kinematic bending moment is obtained from Eq. (20) as


Mt max 2 2:5 107 0:00636 0:115 0:00515 0:842 264 kN m 0:6

A:9

Dynamic approach based on F2 A free-eld site response analysis using A-AAL018 record as input provides (g1)dyn 0.0036. Assigning, as before, F2 to be equal to 1.20, Eq. (22) yields
M t max 2 2:5 107 0:00636 0:115 0:0036 1:20 263 kN m 0:6

A:10

Dynamic approach based on Z The surface acceleration as at resonance, tc and Mresonance are the same as before, due to the hypothesis of linearly viscoelastic soil behaviour. For (finput/f1) 1.9, Eq. (11) yields Z 0.26. Finally, applying Eq. (23): M t max ZMresonance 0:26 1187 308kN m A:11

Following the Maiorano et al. [35] approach, with the parameter b 0.053 obtained in this study yields the estimate: M t max 0:053 67:5 19 000 0:0036 244 kN m A:12

References
[1] Flores-Berrones R, Whitman RV. Seismic response of end-bearing piles. Journal of Geotechnical Engineering Division, ASCE 1982;108(4):55469. [2] Kavvadas M, Gazetas G. Kinematic seismic response and bending of free-head piles in layered soil. Geotechnique 1993;43(2):20722. [3] Mylonakis G. Simplied model for seismic pile bending at soil layer interfaces. Soils and Foundations 2001;41(4):4758. [4] Nikolaou S, Mylonakis G, Gazetas G, Tazoh T. Kinematic pile bending during earthquakes: analysis and eld measurements. Geotechnique 2001;51(5): 42540. [5] Novak M. Piles under dynamic loads: state of the art. In: Proceedings of the second international conference on recent advances in geotechnical earthquake engineering and soil dynamics, St. Louis, vol. 3, 1991. p. 243356. [6] Pender M. Seismic pile foundation design analysis. Bulletin of the New Zealand National Society for Earthquake Engineering 1993;26(1): 49160. [7] Gazetas G, Mylonakis G. Seismic soilstructure interaction: new evidence and emerging issues. In: Dakoulas P, Yegian MK, Holtz RD, editors. Geotechnical Earthquake Engineering and Soil Dynamics III, Geo-Institute ASCE Conference, Seattle, vol. II, 1998. p. 111974. [8] Mizuno H. Pile damage during earthquake in Japan (19231983). In: Dynamic response of pile foundations experiment, analysis and observations. Geotechnical Special Publication, vol. 11. ASCE, 1987. p. 5378. [9] Matsui T, Oda K. Foundation damage of structures. Special Issue Soils Foundation 1996:189200. [10] Tokimatsu K, Mizuno H, Kakurai M. Building damage associated to geotechnical problems. Special Issue Soils Foundation 1996:189200. [11] Horikoshi K, Tateishi A, Ohtsu H. Detailed investigation of piles damaged by Hyogoken Nambu earthquake. In Proceedings of the 12th world congress on earthquake engineering 2000, paper no. 2477 (in CD-ROM). [12] Luo X, Murono Y. Seismic analysis of pile foundations damaged in the January 17, 1995 South-Hyogo earthquake by using the seismic deformation method. In: Proceedings of the fourth international conference on recent advances in geotechnical earthquake engineering and soil dynamics, paper 6.18, 2001. [13] Mizuno H, Iiba M, Kitagawa Y. Shaking table testing of seismic buildingpile two layeredsoil interaction. In: Proceedings of the eighth world conference on earthquake engineering, San Francisco, vol. 3, 1984. p. 64956. [14] Meymand P. Shaking table scale model test of non linear soilpilesuperstructure interaction in soft clay. PhD dissertation, University of California, Berkeley, 1998. [15] Wei X, Fan L, Wu X. Shaking table tests of seismic pilesoilpierstructure interaction. In: Proceedings of the fourth international conference on recent advances in geotechnical earthquake engineering and soil dynamics, paper 9.18, 2001. [16] Chau KT, Shen CY, Guo X. Non linear seismic soilpilestructure interactions: shaking table tests and FEM analyses. Soil Dynamics and Earthquake Engineering 2009;29:30010.

[17] Tokimatsu K, Suzuki H. Seismic soilpilestructure interaction based on large shaking table tests. In: Proceedings of the international conference on performance-based design in earthquake geotechnical engineering. IS-Tokyo, 2009. [18] Moccia F. Seismic soil pile interaction: experimental evidence. PhD dissertation, Universita degli Studi di Napoli Federico II, Napoli, 2009. [19] NEHRP. Recommended provisions for seismic regulations for new buildings and other structures (FEMA 450). Building Seismic Safety Council, Washington, DC, 2003. [20] EN 1998-5. Design of structures for earthquake resistance: foundations, retaining structures and geotechnical aspects. CEN European Committee for Standardization, Bruxelles, Belgium, 2005. [21] NTC2008M.LL. Norme Tecniche per le Costruzioni (NTC2008). M.LL.PP. DM 14 gennaio 2008, Gazzetta Ufciale della Repubblica Italiana, 29, 2008. [22] Margason E, Halloway DM. Pile bending during earthquakes. In: Proceedings of the sixth world conference on earthquake engineering, Meerut, India, 1977. p. 16906. [23] Kagawa T, Kraft LM. Lateral loaddeection relationships of piles subjected to dynamic loads. Soils and Foundations 1980;20(4):1936. [24] Dobry R, ORourke MJ. Discussion on Seismic response of end-bearing piles by Flores-Berrones R. and Whitman R.V. Journal of Geotechnical Engineering Division, ASCE 1983;109:77881. [25] Tazoh T, Shimizu K, Wakahara T. Seismic observations and analysis of grouped piles. In: Dynamic response of pile foundations: experiment, analysis and observation. Geotechnical Special Publication, vol. 11, 1987. p. 120. [26] Conte E, Dente G. Il comportamento sismico del palo di fondazione in terreni eterogenei. In: Proceedings of the XVII Convegno Nazionale di Geotecnica, Taormina, vol. 1, 1989. p. 13745 [in Italian]. [27] Mineiro AJC. Simplied procedure for evaluating earthquake loading on piles. Lisbon: De Mello Volume; 1990. [28] Kaynia AM, Mahzooni S. Forces in pile foundations under seismic loading. Journal of Engineering Mechanics, ASCE 1996;122(1):4653. [29] Mylonakis G, Nikolaou A, Gazetas G. Soilpilebridge seismic interaction: kinematic and inertial effects. Part I: soft soil. Earthquake Engineering and Structural Dynamics 1997;26:33759. [30] Guin J, Banerjee PK. Coupled soilpilestructure interaction analysis under seismic excitation. Journal of Structural Engineering 1998;124:43444. [31] Saitoh M. Fixed-head pile bending by kinematic interaction and criteria for its minimization at optimal pile radius. Journal of Geotechnical and Geoenvironmental Engineering 2005;131(10):124351. [32] Cairo R, Dente G. Kinematic interaction analysis of piles in layered soils. In: Proceedings of the 14th European conference on soil mechanics and geotechnical engineering, ISSMGE-ERTC 12 Workshop Geotechnical  tron Editore; cd-rom, paper no. Aspects of EC8, Madrid, Spain. Bologna: Pa 13, 2007. [33] Cairo R, Conte E, Dente G. Nonlinear seismic response of single piles. In: Santini A, Moraci N, editors. Proceedings of the seismic engineering conference on commemorating the 1908 Messina and Reggio Calabria Earthquake (MERCEA08), Reggio Calabria, Italy, vol. 1. New York: American Institute of Physics, Melville; 2008. p. 6029. [34] Castelli F, Maugeri M. Simplied approach for the seismic response of a pile foundation. Journal of Geotechnical and Geoenvironmental Engineering 2009;135(10):144051. [35] Maiorano RMS, De Sanctis L, Aversa S, Mandolini A. Kinematic response analysis of piled foundations under seismic excitation. Canadian Geotechnical Journal 2009;46(5):57184. [36] De Sanctis L, Maiorano RMS, Aversa S. A method for assessing kinematic bending moments at the pile head. Earthquake Engineering and Structural Dynamics 2010;39:113354. [37] Di Laora R, Mandolini A. Some remarks on the kinematic vs. inertial interaction for piled foundations. In: Proceedings of the international conference on performance-based design in earthquake geotechnical engineering, IS-Tokyo, 2009. [38] Moccia F, Sica S, Simonelli AL. Kinematic interaction: a parametric study on single piles. In: Proceedings of the international conference on performancebased design in earthquake geotechnical engineering, IS-Tokyo, 2009. [39] Sica S, Mylonakis G., Simonelli AL. Kinematic bending of piles: analysis vs. code provisions. In: Proceedings of the fourth international conference on earthquake geotechnical engineering, Greece, paper 1674, 2007. [40] Sica S, Mylonakis G, Simonelli AL. Kinematic pile bending in layered soils: linear vs. equivalent-linear analysis. In: Proceedings of the international conference on performance-based design in earthquake geotechnical engineering, IS-Tokyo, 2009. [41] Simonelli AL, Sica S. Fondazioni profonde sotto azioni sismiche. Opere  tron Editore; 2008. Geotecniche in Condizioni Sismiche. Bologna, Italy: Pa p. 30934 [Chapter 11, in Italian]. [42] Dezi F, Carbonari S, Leoni G. A model for the 3D kinematic interaction analysis of pile groups in layered soils. Earthquake Engineering and Structural Dynamics 2009;38(11):1281305. [43] Dezi F, Carbonari S, Leoni G. Kinematic bending moments in pile foundations. Soil Dynamics and Earthquake Engineering 2010;30:11932. [44] Randolph MF. The response of exible piles to lateral loading. Geotechnique 1981;31:24759. [45] Wu G, Finn W. Dynamic Elastic Analysis of pile foundations using nite element method in the time domain. Canadian Geotechnial Journal 1997;34(1):4452.

S. Sica et al. / Soil Dynamics and Earthquake Engineering 31 (2011) 891905

905

[46] Boulanger RW, Curras CJ, Kutter BL, Wilson DW, Abghari A. Seismic soilpile structure interaction experiments and analyses. Journal of Geotechnical and Geoenvironmental Engineering, ASCE 1999;125(9):7509. [47] Kimura M, Zhang F. Seismic evaluation of pile foundations with three different methods based on three-dimensional elasto-plastic nite element analysis. Soils and Foundations 2000;40(5):11332. [48] El Naggar MH, Novak M. Nonlinear analysis of dynamic lateral pile response. Soil Dynamics and Earthquake Engineering 1996;15:23344. [49] Maheshwari BK, Truman KZ, El Naggar MH, Gould PL. 3D FEM nonlinear dynamic analysis of pile groups for lateral transient and seismic excitations. Canadian Geotechnical Journal 2004;41(1):11833. [50] Maiorano RMS, Aversa S, Wu G. Effects of soil non-linearity on bending moments in piles due to seismic kinematic interaction. In: Proceedings of the fourth international conference on earthquake geotechnical engineering, paper 1574, Thessaloniki, Greece, 2007. [51] Veletsos AS, Ventura CE. Efcient analysis of dynamic response of linear system. Earthquake Engineering and Structural Analysis 1984;12:52136. [52] Scasserra G, Lanzo G, Stewart JP, Delia B. SISMA (Site of Italian Strong-Motion Accelerograms): a web-database of ground motion recordings for engineering applications. In: Proceedings of the seismic engineering conference commemorating the 1908 Messina and Reggio Calabria Earthquake, Reggio Calabria, vol. 2. Melville, New York: AIP; 2008. p. 164956 /http://sisma.dsg. uniroma1.itS. [53] OPCM 3274. Primi elementi in materia di criteri generali per la classicazione sismica del territorio nazionale e di normative tecniche per le costruzioni in zona sismica. G.U. no. 105 8/5/2003, 2003 [in Italian]. [54] Rathje M, Abrahamson NA, Bray JD. Simplied frequency content estimates of earthquake ground motions. Journal of Geotechnical and Geoenvironmental Engineering 1998;124(2):1509. [55] Kausel E, Roesset JM. Stiffness matrices for layered soils. Bulletin of the Seismological Society of America 1981;71(6):174361. [56] Di Laora R. Seismic soilstructure interaction for pile supported system. PhD dissertation, Universita degli Studi di Napoli Federico II, Napoli, 2009. [57] ANSYS. Users manual. Houston, USA: SAS IP. [58] Novak M. Dynamic stiffness and damping of piles. Canadian Geotechnical Journal 1974;11:57491.

[59] Blaney GW, Kausel E, Roesset JM. Dynamic stiffness of piles. In: Proceedings of the second international conference on numerical methods in geomechanics, 19th ed., vol. 2, 1976. p. 100112. [60] Roesset JM. The use of simple models in soil-structure interaction. In: ASCE specialty conference, Knoxville, TN, Civil Engineering and Nuclear Power, vol. 2, 1980. [61] Gazetas G, Fan K, Tazoh T, Shimizu K, Kavvadas M, Makris N. Seismic pile groupstructure interaction. Piles under dynamic loads. Geotechnical Special Publication 1992;No. 34(ASCE):5693. [62] Mylonakis G. Contributions to static and seismic analysis of piles and pilesupported bridge piers. PhD dissertation, State University of New York at Buffalo, U.S.A., 1995. [63] Dobry R, ORourke MJ, Rosset JM, Vicente E. Horizontal stiffness and damping of single piles. Journal of Geotechnical Engineering Division, ASCE 1982;108(GT3):43959. [64] Syngros C. Seismic response of piles and pile-supported bridge piers evaluated through case histories. PhD dissertation, The City College and the Graduate Center of the City University of New York, New York, 2004. [65] Wu G. VERSAT-P3D: quasi-3D dynamic nite element analysis of single piles and pile groups, version 2006. Canada: Wutec Geotechnical International; 2006. [66] Bardet JP, Ichii K, Lin CH. EERAa computer program for equivalent-linear earthquake site response analyses of layered soil deposits. University of Southern California, 2000. [67] Schnabel B, Lysmer J, Seed H. SHAKE: a computer program for earthquake response analysis of horizontally layered sites. Report E.E.R.C. 70-10, Earthquake Engineering Research Center, University of California, Berkeley, 1972. [68] Seed HB, Idriss IM. Ground motions and soil liquefaction during earthquakes. Berkeley, CA: Earthquake Engineering Research Institute; 1982 134 pp. [69] Cairo R, Dente G, Sica S, Simonelli AL. Soil-pile kinematic interaction: from research to practice. Italian Geotechnical Journal, in press, doi:1-2/RIG 2011. [70] ABAQUS. Users manual. ABAQUS Inc. and DS. On line documentation /http://www.simulia.com/support/documentation.htmlS. [71] Nikolaou A, Mylonakis G, Gazetas G. Kinematic bending moments in seismically stressed piles. Report NCEER-95-0022, National Center for Earthquake Engineering Research, State University of New York, Buffalo, NY, 1995. 220 pp.

Anda mungkin juga menyukai