You are on page 1of 22

Bull Earthquake Eng

https://doi.org/10.1007/s10518-017-0268-6

ORIGINAL RESEARCH PAPER

Seismic assessment of a benchmark based isolated


ordinary building with soil structure interaction

Davide Forcellini1

Received: 6 April 2017 / Accepted: 10 November 2017


Ó Springer Science+Business Media B.V., part of Springer Nature 2017

Abstract Base isolation (BI) has been applied all over the world as a well-known tech-
nique in order to reduce the destroying effects of earthquakes. Even if many researches
have been published on this issue, few contributions have been focused on the effects that
soil-structure interaction (SSI) can have on isolated buildings. In this regard, the paper
aims at simulating the SSI effects on a residential structure by performing 3D numerical
simulations. The soil is described with non-linear hysteretic materials and advanced
plasticity models. The paper applies the open-source computational interface OpenSeesPL,
implemented within the finite element code OpenSees. The interface performs the 3D
spatial soil domain, boundary conditions and input seismic excitation with convenient post-
processing and graphical visualization of results (including deformed ground response time
histories). The proposed approach enables to drive the assessment of isolation technique
with evaluation of soil non-linear response into a unique twist. Therefore, the paper aims at
assessing the cases where BI becomes detrimental. In particular, the model of the structure
allows us to assess the structural performance, calculating accelerations and displacements
at various heights. Consequently, this study can be considered one of the few attempts to
propose new design considerations for engineers and consultants.

Keywords Soil structure interaction  Base isolation  Ordinary building  Non


linearity  Numerical simulations  OpenSees  Seismic engineering

1 Background

Base isolation (BI) is one of the most convincing solutions in order to protect structures
from the destroying effects of earthquakes. This technique allows to decouple the structure
from the ground by intentionally concentrating seismic energy dissipation in a unique

& Davide Forcellini


davide.forcellini@unirsm.sm
1
University of San Marino, 47890 San Marino, Republic of San Marino

123
Bull Earthquake Eng

element (with low horizontal stiffness) between the foundation and the structure. In par-
ticular, structural responses are directly connected with the mutual dynamic characteristics
(natural frequencies) of the soil. In case of rigid soils, these characteristics depend on
structural mass and stiffness only. When the foundation soil is deformable, soil shear
velocity and layer depth (Kramer 1996) can affect the response of the entire system. In this
regard, extensive research has been conducted in the past 30 years regarding the effects of
SSI on the seismic response of civil engineering structures. In case of strategic structures
such as nuclear plans, international codes require to consider SSI effects in analysis and
design, as investigated by Bolisetti and Whittaker (2015) and Coleman et al. (2016). For
other kinds of structures, such as residential buildings, many codes (such as ATC-3-06 and
NEHRP-97) suggest simply to neglect SSI. However, some contributions, such as Luco
(1982) and Renzi et al. (2013), show that SSI can be non-conservative for safety and cost
reduction. The majority of these publications study SSI simply by introducing springs,
dashpots and artificial masses in the interface between the structure and the soil. This
approach is generally accepted, since modelling SSI is a challenging problem for
numerical simulations. However, it could be insufficiently detailed in order to take into
account all the existing phenomena. In this regard, this study aims at overcoming the
previous simplifications considering 3D numerical simulations of a soil-structure system
applied to a shear-type building isolated at the base. Moreover, many contributions such as
Novak and Henderson (1989), Vlassis and Spyrakos (2001), Tongaonkar and Jangid
(2003), Ucak and Tsopelas (2008) and Forcellini (2017) show that beneficial effects of BI
can be strongly affected by soil deformability and energy dissipation in the ground. In
particular, Constantinou and Kneifati (1988) and Sidiqui and Constantinou (1989) per-
formed analyses of multistory shear-type base-isolated structures on deforming subsoils. In
these contributions, the base-isolated (BI) structure is replaced by an equivalent single
degree of freedom (SDOF) and the results were extended to multi degrees of freedom
structures. In addition, Haiyang et al. (2014) showed that SSI effects weakens the isolation
efficiency by reproducing extensive experimental studies based on a structural scheme with
several degree of freedom. This paper aims at verifying such previous findings with a 3D
numerical approach that performs multi degrees of freedom structures, enabling to consider
the effects of higher modes. It considers several soil conditions in order to study the effect
of deformability. In particular, the paper performs the soil with non-linear hysteretic
materials and advanced plasticity models. Thanks to these assumptions, it was possible to
assess the mutual effects of BI technique and soil non-linearity at the same time. The
ultimate goal is the assessment of cases where BI becomes detrimental. In particular, the
study shows that considering the structure fixed at the base is non-conservative and it
underestimates dynamic effects. The study aims at considering the effects of SSI on
residential buildings by taking into account the previous contributions such as Figini and
Paolucci (2016), Pecker et al. (2013), Renzi et al. (2013), Sáez et al. (2013), Rayhani and
Naggar (2012), Saouma et al. (2011), Pitilakis et al. (2008), Mylonakis and Gazetas (2000)
and Iida (1998). This study considers a 3D soil-structure system and applies the FE
computational interface Opensees PL (Lu et al. (2011)), implemented in OpenSees
(Mazzoni et al. 2009). This platform is able to apply non-linear hysteretic materials with
plasticity models for the soil and a detailed description of the structural system (building
and foundation). In this regard, the proposed model can reproduce soil hysteretic elasto-
plastic shear response (including permanent deformations), damping foundation impe-
dances and realistic boundary conditions at the base of the soil and in correspondence with
the interface between the soil and the structure (as shown in Elgamal et al. 2009; Forcellini
and Gobbi 2015; Forcellini et al. 2016; Forcellini 2017). In particular, the study performs

123
Bull Earthquake Eng

the behavior of a realistic base isolated building consisting of a 4-story shear-type concrete
structure. It assesses the response at the base of the structure (foundation level) and along
its height.

2 Case study

The paper performs a case study of a benchmark BI concrete building (Figs. 1, 2) on


different deformable soil conditions.
The study performs the soil with non-linear hysteretic materials and advanced plasticity
models, in order to reproduce soil hysteretic elasto-plastic shear response (including per-
manent deformations) and damping foundation impedances. It applies the open-source
computational interface OpenSeesPL (Lu et al. 2011) implemented within the FE code
OpenSees (Mazzoni et al. 2009). The study has been divided into several steps. The first
step aims at calibrating the mesh and it consists of performing eigenvalue analyses. It
applies a hard soil (named soil A, shear wave velocity equal to 1600 m/s) in order to
reproduce rigid base conditions (see Sect. 4) and neglect SSI effects. In the second step,
soil deformability has been introduced and the dynamic characteristics (in terms of natural
periods) of the system (soil and structure) have been calculated. The third step consists of
performing dynamic analyses. Seven input motions (shown in Fig. 3 and Table 1) were
selected (on the base of Eurocode prescriptions) in order to significantly affect the dynamic
characteristics of the structure and applied along the x-axis (longitudinal direction). The
analyses assess the system performance by considering structural parameters, such as
accelerations, base shears and displacements. For all analyses, the Newmark transient
integrator is used with c = 0.6 and = 0.3. Stiffness and mass proportional damping is
added with 2% equivalent viscous damping at 1 and 6 Hz. The following sections describe
these steps.

Fig. 1 Opensees 3D system mesh—3D view

123
Bull Earthquake Eng

Fig. 2 Opensees 3D system mesh—vertical and plan view (structure)

3 Input n.1

Input n.2
2.5
Input n.3

Input n.4
2
Input n.5
Sa (g)

1.5 Input n.6

Input n.7
1

0.5

0
0.00 0.50 1.00 1.50 2.00 2.50 3.00 3.50 4.00
T (s)

Fig. 3 Input motions Spectra

3 Computational model

The finite element model (FEM) has been built with OpenSees. It allows high levels of
advanced capabilities for modelling and analyzing non-linear responses of systems, using a
wide range of material models, elements and solution algorithms. This platform consists of

123
Bull Earthquake Eng

Table 1 Input motions (characteristics)


Input Station PGA (g) PGV (cm/s) PGD (cm)

n. 1 Landers (1992) Lucerne Valley 0.72 147.45 265.14


n. 2 Northridge (1994) Rinaldi Receiving 0.89 185.08 60.07
n. 3 Northridge (1994) Sylmar Converter 0.70 135.82 58.20
n. 4 Northridge (1994) Sylmar Hospital 0.87 139.54 50.37
n. 5 Hyogo-Ken (1995) Takatori 0.74 155.44 44.95
n. 6 Erzincan (1992) Erzincan 0.44 125.80 53.30
n. 7 El Centro (1940) CALTECH IIA001 0.35 38.47 82.44

a framework for saturated soil response as a two-phase material following the u-p (where u
is displacement of the soil skeleton and p is pore pressure) formulation. Its interface had
been originally calibrated for pile analyses. Here it has been modified in order to consider
the presence of the system structure—foundation. The Rayleight damping was assigned to
the whole model and calculated (for all soil conditions) as: 2% at 1 and 6 Hz frequency.

3.1 Structure

The structure (Figs. 1, 2) consists of a four-story (floor height: 3.4 m, total height: 13.6 m)
reinforced concrete benchmark building with a 4 9 2 columns scheme (2 in the longitu-
dinal direction (15 m spaced) and 4 in the transversal direction (10 m spaced)). Each
column was discretized in sixteen elasticBeamColumn elements. The total height of the
building is 13.6 m. Seismic masses were calculated considering the structural components
(such as beams, columns, slabs, balconies, stairs) and were concentrated at each floor.
Thanks to its plan and vertical regularity, the building was performed with a shear-type
behavior. In addition, the isolation devices are considered to perform their function.
Therefore, the structure was assumed to be capacity-designed (it is able to respond in the
elastic range). Isolation devices at the base of the building have been modelled by applying
the so-called ElastomericBearing element (Mazzoni et al. 2009). The structural model
overcomes the previous publications, where the structure was modelled with a single
degree of freedom: Forcellini and Gobbi (2015) and Koutsourelakis et al. (2002a, b). The
foundation has been modelled with a 28.4 m 9 34.4 m, 0.5 m thick concrete slab and it
was connected to the soil in order to simulate a rigid shallow foundation. The slab is
surrounded by a weak soil (named in the following with soil W), representing the infill
layer of soil surrounding the foundation. Both the foundation and W soil have been
modelled with the implemented material named Pressure Independent Multiyield (Lu et al.
2011; Mazzoni et al. 2009). Table 2 shows the values used for these two materials.

Table 2 Material models


Concrete Soil W

Mass density (kN/m3) 24 17


Reference shear modul (kPa) 1.25 9 107 5.50 9 104
Reference bulk modul (kPa) 1.67 9 107 1.50 9 105

123
Bull Earthquake Eng

3.2 Soil model

The 3D mesh (Fig. 2) aims at performing 3D SSI analyses, by applying OpenSees


potentialities. Based on previous studies (Elgamal et al. 2009; Forcellini and Gobbi 2015),
the soil has been considered a one-layer homogenous cohesive material with 20 m depth.
In order to adequately reproduce the mechanisms of ground deformability, a high level of
spatial resolution and mesh dimension have been necessarily performed. The study of the
final mesh has been calibrated with a convergence study between four different meshes
with increasing dimensions and number of elements. The 3D FE mesh
(68.4 m 9 74.4 m 9 20.5 m) was built up with 8228 non-linear solid brick elements
called ‘‘Bbar brick’’ (Mazzoni et al. 2009) and 9660 nodes. Vertically the mesh was
discretized in ten layers (0.5 m thickness for the first and 20 m for the others). The model
base boundaries were set at 20.5 m depth from ground surface. Mesh dimensions have
been determined between 0.125 and 0.027 times of the Rayleigh wavelength, in according
with the suggestions indicated in Attewell and Farmer (1973) and Jesmani et al. (2012).
Discretization is built up with relatively small elements around the structure and gradually
larger toward the outer mesh boundaries. The model applies hysteretic elasto-plastic
materials in order to take into account the realistic soil behaviour, modified by stiffness
degradation and energy dissipation (Caballero and Farahmand-Razavi 2008). In particular,
the damping is not predefined by the user with a value, as many numerical platforms do. It
is directly computed by the implemented materials (including permanent deformations and
damping foundation impedances). Moreover, base and lateral boundaries have been
modelled to be impervious, as to represent a small section of a presumably infinite (or at
least very large) soil domain and allowing the seismic energy to be removed from the site
itself. In order to dissipate the arriving scattered waves, lateral boundaries need to move in
pure shear (and thus simulating free-field conditions), as reported in Coleman et al. (2016).
Therefore, they have been located as far as possible from the structure as to decrease their
effects on the response. In particular, at any special location, symmetry conditions can be
adopted and periodic boundaries (Law and Lam 2001) have been considered. Displacement
degrees of freedom of the left and right boundary nodes have been tied together both
longitudinally and vertically using the penalty method. For more details, see Elgamal et al.
(2009), Forcellini and Gobbi (2015) and Forcellini (2017). Soil damping has been mod-
elled by considering a non-linear material (Parra 1996; Yang et al. 2003), allowing to take
into account the dynamic nature of the phenomena (such as hysteretic response and
radiation damping). In particular, a Von Mises multi-surface (Iwan 1967; Mroz 1967)
kinematic plasticity model has been applied in order to reproduce the soil hysteretic elasto-
plastic shear response (including permanent deformations). The model is developed within
the framework of multi-yield-surface plasticity (Prevost 1985) and focusing on controlling
the magnitude of cycle-by-cycle permanent shear strain accumulation (see Elgamal et al.
2003; Yang et al. 2003). Plasticity is exhibited only in the deviatoric stress–strain response.
The volumetric stress–strain response is linear-elastic and it is independent of the devia-
toric response. Plasticity is formulated based on the multi-surface (nested surfaces) con-
cept, with an appropriate non-associative flow rule (Prevost 1985; Dafalias 1986;
Bousshine et al. 2001; Nemat-Nasser and Zhang 2002; Radi et al. 2002). The non-linear
shear stress strain back-bone curve is represented by the hyperbolic relation (Kondner
1963), defined by the two material constants (the low-strain shear modulus and the ultimate
shear strength). The soils have been modelled with four clay materials called Pressure
Independent Multiyield (Mazzoni et al. 2009) and built up with representative parameters

123
Bull Earthquake Eng

(Table 2). Characteristic site periods have been calculated by assuming a uniform and
damped soil (Kramer 1996) and the linear formulation: T = 4 H/Vs. H is the height of the
soil layer and Vs the shear wave velocity of each layer. In order to introduce these
parameters inside the code, several backbone curves have been implemented. Figure 4
shows the applied curves for the considered soils (named Soil A, Soil B, Soil C and Soil
D).

3.3 Interface

The interface between the columns and the foundation has been modelled following the
previous research studies (Elgamal et al. 2009; Forcellini, 2017). In particular, rigid beam-
column links (normal to the column longitudinal axis) were used to represent the geometric
space occupied by the column. The soil domain 3D brick elements are connected to the
column at the outer nodes of these rigid links using the equalDOF constraint (for trans-
lations only), implemented in Opensees. This link connects two separate points (one
belonging to the structure and the second to the soil) and it imposes the displacements to be
the same between the structure and the soil nodes (Lu et al. 2011; Mazzoni et al. 2009).
Therefore, the system is able to capture the interaction between the columns and the
foundation, including the potential settlement into the surrounding soil.

4 Calibration

The calibration aims at testing Opensees potentialities in reproducing the problem. The
original FIX model (without isolation) has been considered here. Two steps are here
investigated.
Following Coleman et al. (2016) and Forcellini (2017), the first step aims at calibrating
the fixed-base structural model. A modal analysis of the system based on a hard soil
(named Soil A, with shear velocity equal to 1600 m/s, SSI neglected) was simulated. In
order to verify the fixed conditions, acceleration time histories at the base of the mesh were
compared with the ones at the top of the soil (and thus propagating at the base of the
columns). The two time histories were found identical. One eigenvalue analysis has been

1100
1000
900 SOIL A
Shear Stress (kPa)

800 SOIL B
700 SOIL C
SOIL D
600
500
400
300
200
100
0
0 0.5 1 1.5 2 2.5 3 3.5
Shear Strain(%)

Fig. 4 Backbone curves

123
Bull Earthquake Eng

performed in order to calculate the first four shape modes (Fig. 5). These results were
compared with those obtained by a SAP2000 3D numerical simulation and a theoretical
simplified 1D scheme (Table 4). It was shown that OpenSees assumptions are able to
represent the dynamic characteristics of the building in case of fixed conditions.
The second step aims at verifying the soil domain model. As shown in Coleman et al.
(2016) and Forcellini (2017), this task can be performed by comparing the response of the
free-field soil domain model (without the structure) with known cases. In particular, the
three selected soils were performed in order to represent typical clay materials (Table 3).
Figure 6 shows transfer functions between the input motion at the base of the mesh and the
output at the top of the soil layer for a selected motion (along longitudinal component) for
all the considered soil conditions (here input n. 2 is shown, but the other 6 input motions
showed similar results). The peak values are shown to be close to the characteristic site
periods (named with ‘‘linear’’ and represented with vertical lines in the figure) calculated
by the linear formulation T = 4 H/Vs (Kramer 1996). H is the height of the soil layer and
Vs the shear wave velocity of each layer. The comparison verifies the reproduction of the
vibrational characteristics of the soil layer for each homogeneous soils examined in the
paper (Table 4). Table 5 shows the effect that soil deformability plays on the first period of
the system (soil and structure coupled). It is possible to assess the period elongation (well-
known in literature for example in NIST 2012). The increases in the fundamental period of
the system (if compared with soil A) are: 3.20, 7.08 and 19.7% respectively for soil B, soil
C and soil D.

5 Base isolation implementation

The study performs the isolator in the longitudinal direction only. The bearings have been
modelled with a high stiffness (10.5 9 107 N/m2) in the other directions (vertical and
transversal). Table 6 shows the behavior of the isolation devices adopted in the study. They
were modeled by applying the so-called ElastomericBearing element (Mazzoni et al.
2009).
This element needs the calibration of 3 parameters: the initial elastic stiffness (named
K), the yield strength (named Fy) and the post-yield stiffness ratio (alpha). In order to

Fig. 5 Shape modes (Modes 1–4, soil A conditions)

123
Bull Earthquake Eng

Table 3 Soil models


Soil A Soil B Soil C Soil D

Mass density (kN/m3) 22 21 20.5 18


Reference shear modul (kPa) 5.60 9 106 6.10 9 105 1.72 9 105 4.05 9 104
Reference bulk modul (kPa) 7.50 9 106 1.30 9 106 5.17 9 105 1.89 9 105
Cohesion (kPa) 1000 500 160 40
Shear Wave Velocity (m/s) 1600 540 290 150
Characteristic site period (s) 0.05 0.148 0.276 0.533

Fig. 6 Local effects assessment—transfer functions: comparison between the performed soils with the
characteristic site periods (linear, Kramer 1996)

Table 4 Natural periods (longi-


Model T1 (s) T2 (s) T3 (s) T4 (s)
tudinal direction)
3D (SAP 2000) 0.679 0.201 0.103 0.065
1D (theory) 0.660 0.213 0.133 0.106
Opensees 0.672 0.229 0.148 0.121

Table 5 Natural periods in cor-


T1 (s) T2 (s) T3 (s) T4 (s)
respondence with different soil
conditions
Soil A 0.678 0.236 0.150 0.121
Soil B 0.700 0.240 0.153 0.121
Soil C 0.726 0.282 0.238 0.151
Soil D 0.812 0.508 0.360 0.248

assess the behavior of the isolator, the real height (38.4 cm) was assigned. Two different
models were considered: a linear (BI-L) and a bilinear (BI-NL) model (Fig. 7).
For the linear case (BI-L), alpha was set up as 1. The bearing elements do not contribute
to the Rayleigh damping of the system (soil and structure), as specified in Sect. 3. The
isolators were connected with the foundation that is set to be rigid. In order to model a
common basemat, the nodes at the base of the column were linked together with equaldof

123
Bull Earthquake Eng

Table 6 Characteristics of isolation system implemented in the model


D H G Te Smax Kr (kN/ Klead (kN/ Keff (kN/ Fy Alpha
(mm) (mm) (MPa) (mm) (mm) mm) mm) mm) (kN)

BI-L 700 384 0.9 210 440 – – 2.46 – 1


BI- 700 384 0.9 210 440 1.47 25.72 2.46 439 0.057
NL

Fig. 7 Models of the isolators

(Mazzoni et al. 2009). The isolation devices have been designed for soil A case. The
fundamental period was increased from 0.678 to 3.35 s in order to reduce the seismic
demand in terms of pseudo-accelerations (compare Fig. 3).
In order to calibrate the isolated models, an eigenvalue analysis was carried out with all
the considered soils. Figure 8 shows BI shape modes in case of soil A. Table 7 shows the
periods of the structure with several soil conditions. The fundamental period depends on
the isolation device, as shown in Kelly (1993) and Naeim and Kelly (1999) and it increases
for all the soil conditions to 3.35 s. The results show that soil deformability does not affect

Fig. 8 Shape modes (BI model, hard soil)

123
Bull Earthquake Eng

Table 7 Natural periods of BI


Model T1(iso) (s) T2(iso) (s) T3(iso) (s) T4(iso) (s)
model with different soil
conditions
Soil A 3.349 0.297 0.158 0.122
Soil B 3.352 0.297 0.158 0.122
Soil C 3.356 0.297 0.158 0.122
Soil D 3.370 0.297 0.167 0.158

the periods when isolation is applied. Moreover, they demonstrate that isolation performs
its function by decoupling the structure from the soil. Comparing with non-isolated cases
(Table 5), the increases are more significant for soil A than in case of soil D. This
demonstrates that the efficiency of base isolation technique increases with rigid soils, as
shown in Forcellini (2017).

6 Dynamic analyses

In this paragraph dynamic analyses are discussed. Base values (displacements, accelera-
tions and base shears) are calculated in correspondence of the first node of the first
elements of each columns in the case of fixed structure (FIX). In base isolated cases (BI-L
and BI-NL), the base values are calculated in correspondence with the top nodes of the
elastomeric elements. The top values are those in correspondence with the last nodes of the
last elements of each columns. Table 8 shows the effects of soil deformability on maxi-
mum values of base displacements and base accelerations for the considered soil condi-
tions. It is possible to see that soil deformability induces a general increase of maximum
displacements and a decrease of the maximum accelerations.
Figure 9 shows base shear time histories at the base of structure (for FIX case) and at
the top of the isolator (for BI models) in correspondence with input 3 (chosen because the
most severe for the BI structure). It is possible to understand that the isolation systems
perform their function to decouple the structure from the soil in correspondence with soil A
and soil B. For soil C, BI-L and BI-NL models do not show big improvements.
Table 9 shows the maximum values of base shear for FIX, BI-L and BI-NL for all the
input and soil conditions. Column 4 and column 5 show the ratio between the isolated
models (BI-L and BI-NL) and the FIX case. Comparing these two columns, it is possible to
assess the importance of model the isolator with a non-linear behaviour. In correspondence
of soil D, the effect of the isolator depends strongly on the considered input. If FIX model
values are compared for different soils, it is possible to see that base shear values reduce
with soil deformability (and consequently with soil damping). Therefore, for isolated
models, the reductions due to base isolation is not as significant as the ones due to soil
deformability. This happens in correspondence with soil D (Fig. 9), where the base shear
values are bigger for BI models than those for FIX model. In case of high deformable soil
(soil D), the shear strain induced by the earthquake becomes higher than the soil shear
capacity. In these cases, the bearings absorb the strain transmitted to the structure, per-
forming their function. Therefore, the base shears in correspondence with the bearings are
bigger than the values in correspondence with the base of the columns in the fixed con-
figurations. This behavior, where soft soils are shown to act as base isolation systems
themselves, is known in the literature by Forcellini (2017).

123
Bull Earthquake Eng

Table 8 Base displacements and base accelerations (maximum values) for FIX, BI-L and BI-NL for all
input motions (soil A, B, C and D)
Soil A Output accelerations (g) Output displacements (mm)

FIX-A BI-L-A BI-NL-A FIX-A BI-L-A BI-NL-A

Input n. 1 0.96 0.93 0.94 1.16 1.02 0.98


Input n. 2 1.43 1.38 1.44 1.67 0.61 0.73
Input n. 3 0.75 0.75 0.75 1.43 0.66 0.57
Input n. 4 0.94 0.93 0.95 0.94 0.80 0.78
Input n. 5 0.74 0.55 0.77 1.37 0.53 0.59
Input n. 6 0.46 0.46 0.46 0.94 0.30 0.34
Input n. 7 0.46 0.46 0.46 0.55 0.30 0.30

Soil B Output accelerations (g) Output displacements (mm)

FIX-B BI-L-B BI-NL-B FIX-B BI-L-B BI-NL-B

Input n. 1 1.06 1.15 1.14 8.42 10.99 9.71


Input n. 2 0.94 0.97 0.96 13.28 18.36 19.61
Input n. 3 0.84 0.85 0.85 14.06 9.27 8.49
Input n. 4 1.14 1.13 1.14 10.70 18.02 18.61
Input n. 5 0.95 0.94 0.94 27.55 11.11 11.29
Input n. 6 0.67 0.71 0.72 5.73 6.96 7.67
Input n. 7 0.60 0.64 0.63 4.41 5.31 5.20

Soil C Output accelerations (g) Output displacements (mm)

FIX-C BI-L-C BI-NL-C FIX-C BI-L-C BI-NL-C

Input n. 1 0.79 0.71 0.75 62.11 42.38 37.54


Input n. 2 0.71 0.68 0.64 429.72 334.14 341.34
Input n. 3 0.74 0.58 0.57 210.44 185.72 172.69
Input n. 4 0.73 0.78 0.73 196.18 129.75 134.74
Input n. 5 0.87 0.57 0.61 521.88 302.68 307.63
Input n. 6 0.68 0.48 0.57 160.50 69.05 76.97
Input n. 7 0.44 0.48 0.48 24.14 25.31 25.31

Soil D Output accelerations (g) Output displacements (mm)

FIX-D BI-L-D BI-NL-D FIX-D BI-L-D BI-NL-D

Input n. l 0.33 0.42 0.3 8 569.23 549.75 508.77


Input n. 2 0.33 0.35 037 438.59 409.75 426.05
Input n. 3 0.35 2.38 2.23 194.69 226.76 214.3I
Input n. 4 0.41 1.94 1.81 422.57 149.38 140.40
Input n. 5 0.43 1 53 1.55 760.41 125.61 124.77
Input n. 6 0.29 1 02 0.95 408.96 81.46 86.06
Input n. 7 0.19 1.51 1.37 79.97 114.20 106.19

123
Bull Earthquake Eng

Fig. 9 Base shear Input-3 of FIX, BI-L and BI-NL (soil A, B, C and D)

Figures 10 and 11 show a comparison between time histories obtained with FIX and BI
models, in terms of accelerations and displacements on the top of the structure for input
motion 2 (chosen here because the most severe the FIX structure).
Figure 10 shows a reduction of top accelerations in correspondence with soil D for FIX
models. This is due to the effects of soil deformability. Table 10 shows the maximum top
accelerations and top displacements. Columns 4 and 5 show the ratio between the isolated
models (BI-L and BI-NL) and the FIX case for all the soil conditions. In terms of accel-
erations, it is possible to see that the effect of base isolation is significant for soil A
(maximum reductions: BI-L: - 90%, BI-NL: - 86.7%), soil B (maximum reductions: BI-

123
Bull Earthquake Eng

Table 9 Base shear (maximum values) for FIX, BI-L and BI-NL for all input motions (soil A, B, C and D)
Soil A Base shear (kN) BI-L-A BI-NL-A

FTX-A BI-L-A BI-NL-A FTX-A (%) FIX-A (%)

Input n. l 14848 5886 3752 - 60.4 - 74.7


Input n. 2 31660 4029 2808 - 87.3 - 91.1
Input n. 3 34370 8639 3720 - 74.9 - 89.2
Input n. 4 19241 4895 3267 - 74.6 - 83.0
Input n. 5 28457 6069 3797 - 78.7 - 86.7
Input n. 6 18453 5165 3466 - 72.0 - 81.2
Input n. 7 11788 1662 1718 - 85.9 - 85.4
Soil B Base shear (kN) BI-L-B BI-NL-B

FIX-B BI-L-B BI-NL-B FIX-B (%) FIX-B (%)

Input n. l 10268 5798 3762 - 43.5 - 63.4


Input n. 2 24820 4044 3276 - 83.7 - 86.8
Input n. 3 19583 8241 3701 - 57.9 - 81.1
Input n. 4 13035 4657 3167 - 64.3 - 75.7
Input n. 5 22529 5900 3753 - 73.8 - 83.3
Input n. 6 16243 4976 3438 - 69.4 - 78.8
Input n. 7 6504 1616 1691 - 75.2 - 74.0

Soil C Base shear (kN) BI-L-C BI-NL-C

FIX-C BI-L-C BI-NL-C FIX-C (%) FIX-C (%)

Input n. l 10134 5798 3870 - 42.8 - 61.8


Input n. 2 15997 4539 3491 - 71.6 - 78.2
Input n. 3 16560 8009 3916 - 51.6 - 76.4
Input n. 4 12448 4725 3199 - 62.0 - 74.3
Input n. 5 17201 5371 3786 - 68.8 - 78.0
Input n. 6 15597 5151 3655 - 67.0 - 76.6
Input n. 7 6581 1755 1755 - 73.3 - 73.3

Soil D Base shear (kN) BI-L-D BI-NL-D

FIX-D BI-L-D BI-NL-D FIX-D (%) FIX-D (%)

Input n. l 3813 5383 3568 41.2 - 6.4


Input n. 2 4291 2535 2094 - 40.9 - 51.2
Input n. 3 3433 8620 4247 151.1 23.7
Input n. 4 4527 5000 3667 10.5 - 19.0
Input n. 5 4833 6156 3730 27.4 - 22.8
Input n. 6 4270 5086 3752 19.1 - 12.1
Input n. 7 3935 1753 2140 - 55.5 - 45.6

L: - 85.1%, BI-NL: - 80.9%) and soil C (maximum reductions: BI-L: - 76.8%, BI-NL:
- 68.7%). For soil D, soil deformability is so high that the effect of the isolators is reduced
and it becomes detrimental in correspondence with some input motions.

123
Bull Earthquake Eng

Fig. 10 Acceleration time histories at the top of the structure (z = 13.6 m, Input n. 2): FIX, BI-L and BI-
NL models (soil A, B, C and D)

123
Bull Earthquake Eng

Figure 11 shows that the displacements increase in correspondence with soil C, where
the site fundamental period is near to second structural period (T2). In particular, for the
FIX structure, T2 plays a significant role (modal participation mass ratio around 10%).
When isolation is applied (BI), the mass of the structure is activated completely by T1,
which has been shifted away from the period affected by the majority of ground motions.
In terms of maximum top displacements, increases are observed for all the models, due to
soil deformability. Table 10 (last column) shows the ratio between the maximum values in
correspondence with BI-NL model and those in correspondence with BI-L model (maxi-
mum reductions: soil A: - 58.1%, soil B: - 55.4%, soil C: - 45.5%, soil D: - 51.3%). In
particular, it is possible to assess that when isolators are modelled with linear behaviors,
top displacements are overestimated. This happens particularly for rigid soils (soil A and
soil B).
Figures 12 and 13 show the maximum displacements in correspondence with the var-
ious floors of the structure and in correspondence of the isolator for input n.3. Figure 12
compares the different configurations (FIX, BI-L, BI-NL) for the various soil conditions. It
is possible to see that the behavior in correspondence of soil A and B are similar and soil
displacements are small. In correspondence with soil C, floor displacements are still due to
the presence of the isolators. For soil D, soil deformability is so high that reduces the
effects of the isolators. The displacements of FIX configurations are bigger than those
calculated for BI-NL model. In addition, for any soil conditions, the maximum displace-
ments for the BI-L configurations are shown to be not acceptable because they overpass the
prescribed limits of the isolators. In cases of BI-NL models, these displacements are less
than half those calculated in correspondence with the linear isolation. Figure 13 compares
the different soil conditions for each configuration. It shows that for FIX configurations,
floor displacements increase with soil deformability. In particular, the horizontal maximum
displacements concentrate in the first floor of the structure. Configuration BI-L and BI-NL
are almost insensitive to soil conditions thanks to the effects of BI, as expected. Isolators
reduce the displacements in correspondence with the first floor. In addition, it is possible to
assess the benefit of non-linearity in reducing the displacements of the structure.

7 Conclusions

The study conducted in this paper may be viewed as an original contribution to seismic
assessment of a residential benchmark BI building with SSI effects. The FE computational
model of the soil is based on an advanced constitutive formulation that enables to drive the
assessment of structural behavior with evaluation of soil non-linear response into a unique
twist. In particular, the paper assesses the importance of both non-linear behavior of
isolators and soil deformability in evaluating the efficiency of base isolation technique.
The results confirmed that SSI effects become significant when soil deformability
increases. Such effect consists in period elongation, base shear reduction and displace-
ments increase. The presence of the isolation allows to decouple the building from the soil.
Moreover, BI technique was shown to be less effective when soil deformability increases.
In addition, the results show the importance of considering non-linearity both in the soil
and in the isolation model. In particular, neglecting soil non-linearity can be non-con-
servative and can underestimate the dynamic effects of SSI. Moreover, when isolation is
modeled with a non-linear behavior, the system was found to be more effective in reducing
displacements and base shears.

123
Bull Earthquake Eng

Fig. 11 Displacement time histories at the top of the structure (z = 13.6 m, Input n. 2): FIX, BI-L and BI-
NL models (soil A, B, C and D)

123
Bull Earthquake Eng

Table 10 Acceleration and displacement maximum values at the top of the structure (z = 13.6 m) for FIX,
BI-L and BI-NL for all input motions (soil A, B, C and D)
Soil A Top acceleration (g) BI-L-A BI-NL-A Top displacements (mm) BI-NL-A

FIX- BI-L- BI-NL- FIX-A FIX-A FIX- BI-L-A BI-NL- BI-L-A


A A A (%) (%) A A (%)

Input n. 1.15 0.31 0.24 - 73.0 - 79.1 129.61 839.12 594.26 - 29.2
1
Input n. 1.70 0.24 0.30 - 86.1 - 82.6 357.30 420.59 236.37 - 43.8
2
Input n. 1.99 0.46 0.26 - 77.1 - 86.7 289.10 1014.47 424.68 - 58.1
3
Input n. 1.52 0.27 0.31 - 82.3 - 79.7 191.22 698.13 483.94 - 30.7
4
Input n. 1.09 0.27 0.29 - 75.6 - 73.7 273.98 865.93 615.29 - 28.9
5
Input n. 1.09 0.27 0.29 - 75.6 - 73.7 163.05 737.31 530.56 - 28.0
6
Input n. 0.93 0.09 0.24 - 90.0 - 74.7 102.73 237.09 43.93 - 81.5
7

Soil B Top acceleration (g) BI-L-B BI-NL-B Top displacements (mm) BI-NL-B

FIX-B BI-L-B BI-NL-B FIX-B FIX-B FIX-B BI-L-B BI-NL-B BI-L-B


(%) (%) (%)

Input n. 1 0.90 0.32 0.29 - 63.9 - 67.8 137.47 852.62 609.59 - 28.5
Input n. 2 1.62 0.24 0.31 - 84.8 - 80.9 462.60 424.87 234.99 - 44.7
Input n. 3 1.12 0.44 0.27 - 60.5 - 76.2 364.50 955.16 425.55 - 55.4
Input n. 4 1.13 0.28 0.37 - 75.3 - 67.6 228.52 675.96 478.88 - 29.2
Input n. 5 1.50 0.32 0.49 - 78.9 - 67.6 373.21 864.62 616.52 - 28.7
Input n. 6 1.01 0.27 0.29 - 73.6 - 71.8 247.74 731.05 538.46 - 26.3
Input n. 7 0.66 0.10 0.24 - 85.1 - 63.2 82.06 233.62 43.99 - 81.2

Soil C Top acceleration (g) BI-L-C BI-NL-C Top displacements (mm) BI-NL-C

FIX-C BI-L-C BI-NL-C FIX-C FIX-C FIX-C BI-L-C BI-NL-C BI-L-C (%)
(%) (%)

Input n. 1 0.59 0.33 0.26 - 44.5 - 55.5 209.68 894.51 662.65 - 25.9
Input n. 2 1.05 0.24 0.33 - 76.8 - 68.7 711.85 844.04 563.05 - 33.3
Input n. 3 0.94 0.40 0.31 - 56.9 - 66.7 415.46 948.57 517.30 - 45.5
Input n. 4 0.81 0.27 0.29 - 67.1 - 64.8 392.53 780.96 547.97 - 29.8
Input n. 5 1.06 0.29 0.49 - 72.8 - 53.5 782.44 737.41 760.78 3.2
Input n. 6 0.90 0.28 0.27 - 68.9 - 70.4 398.37 803.74 626.97 - 22.0
Input n. 7 0.53 0.29 0.29 - 44.6 - 44.6 100.22 140.15 44.53 - 68.2

Soil D Top acceleration (g) BI-L-D BI-NL-D Top displacements (mm) BI-NL-D

FIX-D BI-L-D BI-NL-D FIX-D FIX-D (%) FIX-D BI-L-D BI-NL-D BI-L-D
(%) (%)

Input n. 1 0.27 0.29 0.21 4.4 - 22.4 602.45 1336.35 1073.51 - 19.7
Input n. 2 0.28 0.11 0.20 - 59.8 - 27.5 245.39 505.57 246.30 - 51.3
Input n. 3 0.22 0.51 0.46 135.5 113.6 602.45 963.67 551.33 - 42.8
Input n. 4 0.34 0.29 0.52 - 15.2 51.2 468.13 690.25 477.01 - 30.9

123
Bull Earthquake Eng

Table 10 continued

Soil D Top acceleration (g) BI-L-D BI-NL-D Top displacements (mm) BI-NL-D

FIX-D BI-L-D BI-NL-D FIX-D FIX-D (%) FIX-D BI-L-D BI-NL-D BI-L-D
(%) (%)

Input n. 5 0.31 0.34 0.46 8.7 49.2 830.21 898.96 528.58 - 41.2
Input n. 6 0.28 0.35 0.35 25.7 25.7 457.20 748.07 573.17 - 23.4
Input n. 7 0.27 0.11 0.42 - 60.2 56.5 130.91 230.05 129.46 - 43.7

soil A soil B
13.6 13.6

10.2 10.2 FIX


FIX H build. (m) BI-L
H build. (m)

BI-L
6.8 BI-NL
6.8 BI-NL

3.4 3.4

0 0
0 200 400 600 800 1000 1200 0 200 400 600 800 1000 1200

Horiz. Displ. (mm) Horiz. Displ. (mm)

soil C soil D
13.6 13.6

10.2 10.2
H build. (m)

FIX
H build. (m)

FIX
BI-L 6.8 BI-L
6.8
BI-NL BI-NL

3.4 3.4

0 0
0 200 400 600 800 1000 1200 0 200 400 600 800 1000 1200

Horiz. Displ. (mm) Horiz. Displ. (mm)

Fig. 12 Horizontal displacements for FIX, BI-L and BI-NL (soil A, B, C and D)-input n.3

In conclusion, this study can be considered a first step in order to propose new design
considerations for engineers and consultants. Further analysis will aim at reproducing more
input motions and implementing more accurate models to simulate more typologies of
isolators.

123
Bull Earthquake Eng

Fig. 13 Horizontal FIX


displacements for FIX, BI-L and
13.6
BI-NL (soil A, B, C and D)-input
n.3

10.2
soil A

H build. (m)
soil B
6.8 soil C
soil D
3.4

0
0 200 400 600 800 1000 1200
Horiz. Displ. (mm)

BI-L
13.6
soil A
soil B
10.2
soil C
H build. (m)

soil D
6.8

3.4

0
0 200 400 600 800 1000 1200
Horiz. Displ. (mm)
BI-NL
13.6

10.2
H build. (m)

6.8
soil A
soil B
soil C
3.4
soil D

0
0 200 400 600 800 1000 1200
Horiz. Displ. (mm)

123
Bull Earthquake Eng

Acknowledgements The author wants to acknowledge Daniele Mina and Andrea Canini, who helped to
perform the interface and adapt it to the case study. Dr. Marcie Brooks is also thanked for her help in editing
the paper.

References
Attewell P, Farmer IW (1973) Attenuation of ground vibrations from pile driving. Ground Eng 6(4):26–29
Bolisetti C, Whittaker AS (2015) Site response, soil-structure interaction and structure-soil-structure
interaction for performance assessment of buildings and nuclear structures. Technical report, MCEER
15-0002, University of Buffalo
Bousshine L, Chaaba A, De Saxcé G (2001) Softening in stress-strain curve for Drucker–Prager nonasso-
ciated plasticity. Int J Plast 17(1):21–46
Coleman JL, Bolisetti C, Whittaker AS (2016) Time-domain soil-structure interaction analysis of nuclear
facilities. Nucl Eng Des 298(2016):264–270
Constantinou MC, Kneifati M (1988) Dynamics of soil-base-isolation structure systems. J Struct Eng ASCE
114(l):211–221
Dafalias YF (1986) Bounding surface plasticity I: mathematical formulation and hypoplasticity. J Eng Mech
ASCE 112(9):966–987
Elgamal A, Yang Z, Parra E, Ragheb A (2003) Modeling of cyclic mobility in saturated cohesionless soils.
Int J Plast 9(6):883–905
Elgamal A, Lu J, Forcellini D (2009) Mitigation of liquefaction-induced lateral deformation in sloping
stratum: three-dimensional numerical simulation. J Geotech Geoenviron Eng 135(11):1672–1682
Figini R, Paolucci R (2016) Integrated foundation-structure seismic assessment through non-linear dynamic
analyses. Earthq Eng Struct Dynam 56:1–6. https://doi.org/10.1002/eqe.2790
Forcellini D (2017) Cost assessment of isolation technique applied to a benchmark bridge with soil structure
interaction. Bull Earthq Eng. https://doi.org/10.1007/s10518-016-9953-0
Forcellini D, Gobbi S (2015) Soil structure interaction assessment with advanced numerical simulations. In:
Computational methods in structural dynamics and earthquake engineering (COMPDYN) conference,
Crete Island, Greece, 25–27 May 2015
Forcellini D, Gobbi S, Mina D (2016) Numerical simulations of ordinary buildings with soil-structure
interaction. In: VI international conference on structural engineering, mechanics and computation
(SEMC). Capetown, South Africa, 05-07 September 2016, ISBN 978-1-138-02927-9
Haiyang Z, Xu Y, Chao Z, Dandan Y (2014) Shaking table tests for the seismic response of a base-isolated
structure with the SSI effect. Soil Dyn Earthq Eng 67(2014):208–218
Iida M (1998) Three-dimensional non-linear soil building interaction analysis in the lakebed zone of Mexico
city during the hypothetical Guerrero earthquake. Earthq Eng Struct Dyn 27:1483–1502
Iwan WD (1967) On a class of models for the yielding behavior of continuous and composite systems.
J Appl Mech ASME 34:612–617
Jesmani M, Fallahi AM, Kashani HF (2012) Effects of geometrical properties of rectangular trenches
intended for passive isolation is sandy soils. Earth Sci Res 1(2):137–151
Kelly JM (1993) Earthquake-resistant design with rubber. Springer, London
Kondner RL (1963) Hyperbolic stress-strain response: cohesive soils. J Soil Mech Found Div
89(SM1):115–143
Koutsourelakis S, Prevost JH, Deodatis G (2002a) Risk assessment of an interacting structure–soil system
due to liquefaction. Earthq Eng Struct Dyn 31:851–879. https://doi.org/10.1002/eqe.125
Koutsourelakis S, Prevost JH, Deodatis G (2002b) Risk assessment of an interacting structure–soil system
due to liquefaction. Earthq Eng Struct Dyn 31:851–879. https://doi.org/10.1002/eqe.125
Kramer SL (1996) Geotechnical earthquake engeneering. In: Hall WJ (ed) Prentice-Hall, international series
in civil engineering and engineering mechanics. Prentice-hall International, Upper Saddle River
Law HK, Lam IP (2001) Application of periodic boundary for large pile group. J Geotech Geoenviron Eng
127–10:889–892
Lu J, Elgamal A, Yang Z (2011) OpenSeesPL: 3D Lateral Pile-Ground Interaction, User Manual, Beta 1.0
(http://soilquake.net/openseespl/)
Luco JE (1982) Linear soil-structure interaction: a review. Earthq Ground Motion Eff Struct 53:41–57
Mazzoni S, McKenna F, Scott MH, Fenves GL (2009) Open system for earthquake engineering simulation,
user command-language manual. (http://opensees.berkeley.edu/OpenSees/manuals/usermanual). Paci-
fic Earthquake Engineering Research Center, University of California, Berkeley, OpenSees version 2.0
Mroz Z (1967) On the description of anisotropic work hardening. J Mech Phys Solids 15:163–175

123
Bull Earthquake Eng

Mylonakis G, Gazetas G (2000) Seismic soil-structure interaction: beneficial or detrimental? J Earthq Eng
4:277–301
Naeim F, Kelly JM (1999) Design of seismic isolated structure—from theory to pratice. Jonh Wiley & Sons
Inc, New York
Nemat-Nasser S, Zhang J (2002) Constitutive relations for cohesionless frictional granular materials. Int J
Plast 18(4):531–547
NIST GCR 12-917-21 (2012) Soil-structure interaction for building structures. U.S. Department of Com-
merce National Institute of Standards and Technology Engineering Laboratory Gaithersburg, MD
20899, September 2012
Novak M, Henderson P (1989) Base-isolated building with soil–structure interaction. Earthq Eng Struct Dyn
18(6):751–765
Parra E (1996) Numerical modeling of liquefaction and lateral ground deformation including cyclic mobility
and dilation response in soil systems. Ph.D. thesis, Rensselaer Polytechnic Institute, Troy, NY
Pecker A, Paolucci R, Chatzigogos CT, Correia AA, Figini R (2013) The role of non-linear dynamic soil-
foundation interaction on the seismic response of structures. Bull Earthq Eng 12(3):1157–1176. https://
doi.org/10.1007/s10518-013-9457-0
Pitilakis D, Dietz M, Wood DM, Clouteau D, Modaressi A (2008) Numerical simulation of dynamic soil-
structure interaction in shaking table testing. Soil Dyn Earthq Eng 28(6):453–467
Prevost JH (1985) A simple plasticity theory or frictional cohesionless soils. Soil Dyn Earthq Eng 4(1):9–17
Radi E, Bigoni D, Loret B (2002) Steady crack growth in elastic-plastic fluid-saturated porous media. Int J
Plast 18(3):345–358
Rayhani MT, El Naggar MH (2012) Physical and numerical modeling of seismic soil-structure interaction in
layered soils. Geotechn Geolog Eng J 30(2):331–342
Renzi S, Madiai C, Vannucchi G (2013) A simplified empirical method for assessing seismic soil-structure
interaction effects on ordinary shear-type buildings. Soil Dyn Earthq Eng 55:100–107
Sáez E, Lopez-Caballero F, Modaressi-Farahmand-Razavi A (2013) Inelastic dynamic soil-structure inter-
action effects on moment-resisting frame buildings. Eng Struct 51:166–177. https://doi.org/10.1016/j.
engstruct.2013.01.020
Saouma V, Miura F, Lebon G, Yagome Y (2011) A simplifyed 3D model for soil-structure interaction with
radiation damping and freefield input. Bull Earthq Eng 9(5):1387–1402
Siddiqui FMA, Constantinou MC (1989) Simplified analysis method for multistory halfspace base-isolated
structures on viscoelastic halfspace. Earthq Eng Struct Dynam 18:63–77
Tongaonkar N, Jandid R (2003) Seismic response of isolated bridges with soil-structure interaction. Soil
Dyn and Earthq Eng 23(4):287–302
Ucak A, Tsopelas P, A.M. ASCE (2008) Effect of soil-structure interaction on seismic isolated bridges.
J Struct Eng 134(7):1154
Vlassis A, Spyrakos C (2001) Seismically isolated bridge piers on shallow soil stratum with soil-structure
interaction. Comput Struct 79:2847–2861
Yang Z, Elgamal A, Parra E (2003) A computational model for cyclic mobility and associated shear
deformation. J Geotechn Geoenviron Eng (ASCE) 129(12):1119–1127

123

You might also like