Next Article in Journal
Bidirectional, Analog Current Source Benchmarked with Gray Molasses-Assisted Stray Magnetic Field Compensation
Next Article in Special Issue
The Role of Soil Structure Interaction on the Seismic Resilience of Isolated Structures
Previous Article in Journal
Bottom-Up Model of Random Daily Electrical Load Curve for Office Building
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A Novel Framework to Assess Soil Structure Interaction (SSI) Effects with Equivalent Fixed-Based Models

by
Davide Forcellini
Department of Civil Engineering, University of Auckland, 20 Symonds St., Auckland 1010, New Zealand
Submission received: 7 October 2021 / Revised: 18 October 2021 / Accepted: 5 November 2021 / Published: 8 November 2021
(This article belongs to the Special Issue Soil Structure Interaction of Resilient Systems)

Abstract

:
Soil structure interaction (SSI) effects have been extensively studied with advanced numerical simulations even if these approaches are time consuming and require much effort to perform. In particular, when SSI models are compared with fixed based ones, two main effects need to be considered: period elongation and damping increase. The paper proposes numerical models to build fixed based models calibrated on these two parameters and perform complex SSI analyses. A new framework that may be used to assess SSI with equivalent fixed-based models is herein presented and validated with non-linear dynamic numerical simulations. Opensees was performed to reproduce non-linear numerical simulations by considering hysteretic materials and advanced soil models.

1. Background

The fundamental period and the damping are properties that significantly affect the dynamic behaviour of structural systems in terms of base shear, deformations and thus seismic forces. In this regard, soil–structure interaction (SSI) effects are particularly connected with the mutual relationship between the fundamental periods of the soil and of the structure. In particular, the comparison between the fixed-based structures (SSI neglected) and the flexible-based ones (where the effects of soil deformability are considered) shows that the fundamental period of the latter configuration elongates due to SSI effects. In addition, the overall damping of the system (soil + structure) also increases. These aspects have been extensively considered in literature both theoretically [1] and numerically [2,3].
In particular, the first studies (such as [4]) demonstrated that period elongation results from inertial interaction that affects the displacement demand of the structure. In particular, Refs. [5,6,7] proposed the so-called “structure to soil stiffness ratio” (that depends on the height and the shear velocity of the flexible layer) and it defines the elongation of the structural period due to SSI. Later, Ref. [8] extensively investigated the formulas in ASCE 7-05 for several buildings by considering over 800 fundamental periods. Ref. [9] proposed an empirical formula for assessing the fundamental period of reinforced concrete structures by performing 3D numerical simulations of SSI effects.
Other approaches were applied, such as [4,10] that used system identification to evaluate the effects of SSI on 57 buildings subjected to strong motions. Moreover, Refs. [11,12,13,14] proposed analytical surveys on the effects of SSI in terms of fundamental period and damping ratio. Experiments were also conducted, such as [15] that studied the effects of SSI on an instrumented 16-story, reinforced concrete building in Moldova, and Ref. [16] examined the experimental results of forced-vibration tests on a school building in Taiwan. In addition, Ref. [17] studied the response of a 14-story reinforced concrete building in Srpska to 20 recorded earthquakes, while [18] assessed the frequency and damping variation of low-rise masonry buildings. Moreover, Ref. [19] proposed to use Dunkerley’s formula to assess the variation of natural frequencies demonstrating high accuracy from experimental and numerical results. In addition, Ref. [20] investigated the effects of SSI with series of free-vibration experiments on a 1/4-scale steel-frame structure by reproducing 34 scenarios under fixed-base and flexible-base conditions. Moreover, Ref. [21] investigated the effects of SSI on several residential RC buildings with different capacity design principles founded on four different soil conditions and applying 20 acceleration records. Furthermore, Ref. [22] proposed a comparison between FEM simulations and dynamic tests with large scale soil-foundation-structure models, Ref. [23] studied the effects of soil–foundation interaction on the seismic response of a cooling tower by applying 3D-FEM analysis and [24] proposed a theoretical lumped-mass model calibrated by experimental results to perform an in-depth parametric analysis.
The main novelties of the paper are: (1) it proposes a methodology that calculates the period elongation and the damping increase with an equivalent 2-DOF (degree of freedom) model. Then, (2) several equivalent MDOF (multi-degree of freedom) systems fixed at the base (named EQ) were produced to represent more complex MDOF systems that account for SSI effects. In particular, EQ models are built up with the two parameters that are calculated in point 1. (3) Finally, in the paper, the models were validated with numerical case studies performed with Opensees PL to consider realistic response of the soil in terms of plastic and hysteretic mechanisms.
Section 2 describes the adopted 2-DOF model, the formulations for the period elongation and damping increase. Section 3 shows the numerical models, while parametric studies that were applied to validate the 2-DOF model are shown in Section 4. Section 5 presents the equivalent fixed-based structure (EQ model). The results are discussed in Section 6.

2. Two Degrees of Freedom (2-DOF) Model

The first step of the proposed methodology consists of modelling the soil–foundation structure system with a simplified 2-DOF model (Figure 1), whose cinematics may be described in terms of:
  • ub (absolute) bedrock displacement;
  • ug ground displacement;
  • us structural displacement.
The preliminary assumption consists of considering the longitudinal direction of the motion, neglecting the rotations.
Applying the second Newton’s law to both masses (ms: structural mass and mg: ground mass), the equation of motion may be written as:
m g u g ¨ + c g u g ˙ + k g u g k s ( u s u g ) = m g u b ¨
m s u s ¨ + c s u s ˙ + k s ( u s u g ) = m s u b ¨
conveniently, it is possible to re-write (1) and (2) by merging the kinematic terms together: accelerations ( u s ¨ ,   u g ¨ ) , velocities ( u s ˙ ,   u g ) ˙ , displacements ( u s , u g ) :
m g u g ¨ + c g u g ˙ + ( k g + k s ) u g k s u s = m g u b ¨
m s u s ¨ + c s u s ˙ k s u g + k s u s = m s u b ¨
and thus, in the matrix form, the equations of motion of the 2-DOF system may written as:
M u ¨ + C u ˙ + K u = M r u b ¨
where:
M = ( m g 0 0 m s ) is the mass matrix;
C = ( c g 0 0 c s ) is the damping matrix;
K = ( k g + k s k s k s k s ) is the stiffness matrix;
u = ( u g u s ) is the displacement vector;
r = ( 1 0 ) is the base vector

2.1. Period Elongation

The natural frequencies of the system can be found by solving the eigenvalue problem:
| K ω 2 M | = 0
ω 4 · m s · m g ω 2 · m g · k s ω 2 · m s · ( k g + k s ) + k s · k g = 0
The equation may be easily solved by putting:
ω 2 = λ
α = (   ω g ω s ) 2
and:
β = m s m g
and by dividing (7) with m s · m g :
λ 2 λ ω s 2 ( 1 + α + β ) + α · ω s 4 = 0
The solution of (11) leads two roots, given by:
λ =   ω s 2 ( 1 + α + β ) ± (   ω s 4 ( 1 + α + β ) ) 2 4 α · ω s 4 2
where the term under the square root is the discriminant that may be rewritten as:
= ω s 4 [ ( 1 α ) 2 + 2 β ( 1 + α ) + β 2 ]
Since soil structure interaction effects are significant when the structural frequency is larger than that of the soil ( m g > > m s ) , β ~ 0 (13) may be rewritten as:
= ω s 4 [ ( 1 α ) 2 ]
Substituting (14) inside (12), leads to the two natural frequencies of the system:
ω 1 2 = ω s 2   1 + α + β + 1 α 2
ω 2 2 = ω s 2   1 + α + β 1 + α 2
(15) and (16) may be rewritten as:
ω 1 = ω s   2 + β 2
ω 2 = ω s   2 α + β 2
and consequently, the fundamental periods of the system are:
T 1 = T S   2 2 + β ~ T S
T 2 = T S   2 2 α + β
It is worth noting that, since α and β are small values, T2 is bigger than T1 and the coefficient 2 2 α + β describes the period elongation that is demonstrated to depend both on the ratio of the masses and on the ratio between the frequencies of the structure mass and of soils. T1 is the first natural period of the system and depends on the structural characteristics. For small β T1 can be approximated with TS (Equation (19)).

2.2. Damping

The 2-DOF model considers the damping as the second effect of SSI. Assuming that the damping is low enough to maintain the orthogonality properties of the mode shapes, it is possible to write:
u 1 ¨ + 2 ξ 1 ω 1 u 1 ˙ + ω 1 2 x 1 = α 1 u b ¨
u 2 ¨ + 2 ξ 2 ω 2 u 2 ˙ + ω 2 2 x 2 = α 2 u b ¨
where:
ξ 1 and ξ 2 are the damping ratios, respectively: ξ 1 = C 1 2 ω 1 M 1 and ξ 2 = C 2 2 ω 2 M 2
α 1 and α 2 are the modal participation factors for the two mode shapes of the system and M1 and M2 are the masses in the first and second mode:
M 1 = ( A ( 1 ) ) T M · ( A ( 1 ) ) and M 2 = ( A ( 2 ) ) T M · ( A ( 2 ) )
Finally, A ( 1 ) and A ( 2 ) are the mode shapes, that can be calculated by introducing (17) and (18) inside:
( K ω i 2 M ) ( A ) = ( 0 )
Solving by setting A 1 ( i ) = 1 :
The first mode shape ( A ( 1 ) )   is given (by setting A 1 ( 1 ) = 1 ) :
( A ( 1 ) ) = ( A 1 ( 1 ) A 2 ( 1 ) ) = ( 1 2 α 2 β β ( 2 + β ) )
and the second mode shape ( A ( 2 ) )   is given (by setting A 1 ( 2 ) = 1 ) :
( A ( 2 ) ) = ( A 1 ( 2 ) A 2 ( 2 ) ) = ( 1 1 2 α + β )
Therefore:
ξ 1 = ( A ( 1 ) ) T C · ( A ( 1 ) ) 2 ω 1 ( A ( 1 ) ) T M · ( A ( 1 ) ) = ξ g β + 2 2 α ( 1 + ( 2 α 2 β ) 2 β ( β + 2 ) 2 ) + ξ s ( 2 α 2 β ) 2 β 2 ( β + 2 ) 2 β + 2 2 ( 1 β + ( 2 α 2 β ) 2 β 2 ( β + 2 ) 2 )
ξ 2 = ( A ( 2 ) ) T C · ( A ( 2 ) ) 2 ω 2 ( A ( 2 ) ) T M · ( A ( 2 ) ) = ξ g 2 α + β 2 α ( 1 + β ( 2 α + β ) 2 ) + ξ s 1 ( 2 α + β ) 2 2 α + β 2 ( 1 β + 1 ( 2 α + β ) 2 )
Equation (27) may be rewritten as:
ξ 2 = ξ g 2 α 2 α + β ( 1 + β ( 2 α + β ) 2 ) + ξ s 1 ( 2 α + β ) 2 2 2 α + β ( 1 β + 1 ( 2 α + β ) 2 )
Equation (28) shows that the damping of the system consists of two contributions: the first is proportional to the soil damping ( ξ g ) and the second term depends on the contribution of the structural damping ( ξ s ) . In particular, Equation (28) indicates also that the damping of the equivalent system is larger than the damping ratio of the structure itself, confirming previous findings, such as [1].

2.3. Soil Parameters

The dependency of period elongation and damping increase have been shown to depend on the coefficient β and consequently on the soil mass that is relatively complex to calculate. In this regard, following the approach in [1], soil frequency can be calculated as:
ω g = π V s 2 H
Therefore, considering the soil as an equivalent 1-DOF system:
ω g = π V s 2 H = k g m g
where vs. is the shear wave velocity and H is the depth of a uniform layer of isotropic, linear elastic soil underlying rigid bedrock [1], it is possible to calculate the soil mass as:
m g = 4 k g H 2 π 2 V s 2
Finally, the stiffness of the soil can be defined [1] as:
k g = 8 G a 2 ν
It is worth noting that previous publications, such as [1], proposed that period elongation and damping increase depend on the ratio between the translational and rotational stiffness of the structure and of the system. Here instead α and β are defined by considering the physical parameters of both the soil (see Equations (31) and (32)) and the structure and thus avoiding to consider empirical definitions of the stiffness. In addition, Equations (20) and (28) confirmed that (1) SSI is relatively significant for cases of flexible structures on stiff soil deposits, as shown in [1]; (2) SSI may be significant for stiff structures founded on soft soils; (3) the fundamental period of soil-structure system is longer than that of fixed-base structure (period elongation); (4) damping of soil-structure system is higher than the damping of the structure alone.

3. Numerical Models

In this section, two series of numerical models are considered: with (1) fixed-base and (2) SSI. Several case studies are herein performed by considering several values of the height of the soil layer and the shear wave velocity. In particular, the case studies consider a 2-floor shear-type building (representing a low-rise residential structure) founded on shallow foundations. Sixteen finite-element models were performed in order to assess the mutual interaction between the soil and the structure. The first one consists of the fixed-based (SSI neglected) structure while the other 15 cases are based on soil layers with different characteristics: different depths (10, 20 and 30 m) and stiffness (100 m/s, 200 m/s, 300 m/s, 400 m/s and 500 m/s).
The 3D numerical models (soil + foundation + structure, Figure 2, Figure 3 and Figure 4) were performed with OpenSees PL, [25,26]. In order to reasonably maintain the computational costs some assumptions were considered. First of all, (1) the foundation slab was modelled as rigid adopting an equivalent linear elastic material with large stiffness, (2) deformations inside the foundations, soil and intermediate nodes along the structure were calculated but not memorized. (3) The floors were considered rigid by introducing rigid diaphragms instead of deformable beams.
FIXED model: the nodes at the base of the structure were restrained in all six directions (longitudinal, transversal and vertical) to fix the structure at the base by neglecting the effects of SSI. This model was used to calculate the periods (0.13 s and 0.04 s) and verify that the first one is representing the dynamic behaviour of the structure (participation mass: 89%). The building is regular in plan (30 m × 20 m rectangular base) and regular along its height (3.3 m, each floor, total height: 6.6 m, above ground level). The seismic-resistant system consists of RC columns (three in longitudinal direction, 6 m spaced, and three in transversal direction, 6 m spaced) that were modelled with elastic beam column elements (11 elements for each floor, parameters: Young’s modulus 3.5 × 107 kN/m2 and shear modulus: 1.73 × 107 kN/m2). The seismic masses were concentrated in correspondence with the connections between the beams and the columns and the floor were tied together to maintain a shear-type behavior, by following the approach already proposed in [27,28].
SSI models: the 15 numerical models were built up to realistically represent non-linear behaviours of the soil consisting of amplitude-dependent amplifications (or de-amplifications), plastic accumulations of ground deformation, and permanent movements (rotations, longitudinal displacements) of the entire system (soil + foundation + structure). Advanced material models were adopted in order to represent the high non-linear behaviour of a weak ground soil.
The foundation consists of a 0.50 m-thick RC rigid slab (28.4 × 34.4 m) designed by calculating the eccentricity loads (the ratio between the overturning bending moment and the vertical force) in the most detrimental way (minimum vertical loads and maximum bending moments) at foundation level. The bearing pressure of the building is: 195 kPa. EqualDOF [26] were used to link together the nodes at the base of the columns and the ones belonging to the foundation in order to model the slab as a rigid element. The nodes of the interface between the column and the foundation were connected with a horizontal rigid beam-column link to model the continuity of the link slab-column. Two solutions were applied to allow continuity of deformation between the foundation (equivalent concrete material) and the soil. (1) Foundations were modelled with pressure independent multi-yield (PIMY), the same model as the soil and consisting of a non-linear hysteretic material with a Von Mises multi-surface kinematic plasticity (Table 1), allowing the continuity among the two models and (2) the interface between the nodes belonging to the foundations and those of the soil were realized with zero length elements to guarantee the balance at each step and avoid convergence problems. Therefore, the models allow us to represent uplifts and gapping between the foundation and the soil, as well as other non-linear effects due to SSI.
The surrounding soil was also modeled with PIMY (Table 2, Figure 5) in order to ensure the connectivity among the two models. In addition, zero length elements were used to represent the interface between the foundation slab and the nodes belonging to the surrounding soil.
Three models of the soil were built with 3D meshes (Figure 2, Figure 3 and Figure 4) with different depth (10, 20 and 30 m, named respectively Model1, Model2 and Model3) and thus with different characteristics in terms of dimensions and number of nodes and BrickUP elements (See Table 3). The soil was modeled with a PIMY model, whose non-linear behavior is described by hyperbolic backbone curve that models the relationship between the shear strains and the shear stresses. A parametric study was carried on two parameters: soil density (1.5, 1.8 and 2.1 kN/m2, respectively for S1, S2 and S3) and the shear wave velocity that was varied between 100 m/s to 500 m/s. Figure 6, Figure 7 and Figure 8 show the non-linear backbone curves, represented by hyperbolic relations defined by the strain shear modulus and ultimate shear strength, while the representative parameters are shown in Table 4, Table 5 and Table 6. The mesh was defined by following the approach included in [27,28]. Every element is built with 20 nodes with three translational DOF, recorded using OpenSees Node Recorder at the corresponding integration points [23]. The number of elements was defined on the basis of a maximum frequency (15 Hz) above which the spectral content of the input may be considered negligible and the element dimensions were increased with the distance to the center of the piles. Transmitting boundaries were considered at the base to dissipate the radiating waves and to accurately model the damping by preventing the reflection of the seismic waves back into the soil medium after being incident on the far-off boundaries. Vertical direction (described by the third DOF) was constrained. Lateral boundaries were modelled by adopting the penalty method (tolerance: 10-4) to avoid problems associated with the equations system conditions. Vertical directions were restrained while longitudinal and transversal directions were set free to allow soil shear deformations.

4. Validation Results

In this section the comparison between the numerical simulations (totally 45 models) and Equation (20) are compared in order to validate and discuss the proposed models.
Figure 9 shows the comparison between the results from Finite Element Model (FEM) and the results from Equation (20) for the 45 performed models.
First or all, it is worth noting that the fundamental periods increase with the depth of the soil layer, as expected, since the system period depends on both α and β that are proportional to H−2 (thickness of soil layer), and thus T is proportional to H.
In addition, for all the models, the smallest values of vs. (100 m/s) give the maximum values of fundamental period, meaning that soil deformability is important in elongating the periods. This is also expected by Equation (20), since α is proportional to the shear wave velocity and thus T is inversely proportional to Vs.
Moreover, the results show that the values provided by the proposed formulation are close to those derived from FEM models. The maximum difference occurs for Model 1 (H = 10 m), for the most rigid soils (Vs = 500 m/s) with +6.71%, 6.92% and 7.34% for S1, S2 and S3, respectively. Therefore, it is possible to see that (1) the proposed formulation is validated by the numerical simulations even if realistic soil mechanisms such as non-linear deformations and plasticity reduce period elongation. (2) The proposed formulation gives small discrepancies of the periods (not necessarily conservative) for the system.

5. Equivalent Fixed-Based Structure (EQ) Model

The proposed formulation was herein applied to perform the equivalent fixed-based structure (EQ model) that represents Model 2 (H = 20 m): S21 (density = 1.8 kN/m2 and Vs = 100 m/s). Parameters α and β were calculated by (9)–(10) and resulted as 0.03 and 0.001469, respectively. Two assumptions were considered: (1) structural masses were proportionally increased to obtain 0.80 s as the fundamental period (compare with Figure 9b) and (2) the parameter η (η = 0.778, [29]) was applied to account for the increased damping of the system (11.5%, instead of the conventional 5%).
Ten input motions were selected from the NGA database on the basis of the Eurocode prescriptions to be compatible with the code-specified spectrum for the life-safety limit state (corresponding to an earthquake with a return period of 475 years, lat.: 42.333 N, 14.246 E., S = 1.50714, Tb = 0.257 s, Tc = 0.770 s, Td = 2.538, Cc = 2.02847, NTC). Figure 10 shows the selected elastic response spectra (pseudo-spectral acceleration vs. period) and the mean spectral ordinates. Note that the mean spectrum (calculated among the selected ones) is included between the −10% and +30% of the code-based spectral shape, as prescribed in [30]. Table 7 shows the characteristics of the selected input motions in terms of peak ground acceleration (PGA), peak ground velocity (PGV) and peak ground displacement (PGD). Note that five of the selected input motions were taken from near field (NF) motions and five from far field (FF) ones. In addition, the set was selected to have significantly dispersion of intensities to significantly affect the dynamic characteristics of the system.
The non-linear dynamic analyses were performed with the longitudinal component of the input motions that was applied at the base of the model. As in [31], the large dimension of the 3D model needs to consider several steps in the analyses. In the first step the soil (without the structure and foundation) was loaded and linear properties (weight, shear and bulk modulus) were considered. In step 2, the soil properties were changed from elastic to plastic (in 25 load steps). In step 3 structural loads were applied and finally dynamic analyses were performed in step 4. The NewtonLineSearch algorithm was used to increase the speed of the solution.

6. Results

Figure 11 shows the results in terms of top acceleration spectra for the selected 10 input motions, for the two models (SSI and EQ model). SSI is the finite element model described in Section 3, while the EQ model is the equivalent fixed-based model with fundamental period 0.80 s and 11.5% damping. It is worth noting that the peak values occur at the fundamental period of the system, demonstrating that both models work properly. Moreover, SSI model results are conservatively smaller than those calculated for the EQ model with a range of 84.0% for BOL and 95.5% for MUL. In particular, the shape of the spectra for many input motions is close between the two models, especially for FF input motions. It is important to notice that the application of the damping (11.5%) and thus the reduction of the input motion by the parameter η allows us to obtain realistic values of the spectra. In only two cases (PER and YER) the values are larger than those calculated by the SSI model. In particular, for these input motions, the differences occur for PGA > 1 g. In one case (BRN), for some values EQ results are not conservative, if compared with SSI cases, and this may be due to the effects of soil non-linear mechanisms that cannot be included in the equivalent fixed model. However, the general trends of the EQ spectra are similar and demonstrate that the proposed methodology is able to represent the amplifications due to site and basin effects. In this regard, five input motions (ERZ, ATM; BOL, CHY and YER) show amplifications at higher periods that are conservatively represented by the EQ spectra in a close way. For RRS, the site effects are represented but with a delay: the peak value for the SSI model (4.97 m/s2) occurs for 1.7 s, while the peak one for EQ is 6.40 m/s2, occurring at 1.9 s, with a delay of 0.2 s. Moreover, it is worth noting that for FF cases, the EQ model seems to be more representative than for the cases of NF, where the differences between the two models are more significant. Overall, the proposed EQ model that implements Equations (20) and (28), is shown to be conservatively correct to represent the complex mechanism due to the SSI between the structure and the soil.

7. Conclusions

The paper assesses the effects of SSI with an equivalent fixed based model (named EQ model) that consider the effects of SSI by applying the period elongation and the damping increase. These parameters were firstly calculated with simplified 2-DOF dynamic models and thus applied to represent the EQ models. Numerical simulations were carried out to verify the accuracy of the models in terms of period elongation, that were compared and shown to be appropriately similar. Moreover, a case study was performed with several input motions (both near and far field) to assess the potentialities of the approach to conservatively represent the mechanisms of soil deformability. The overall outcome is that the proposed framework performs correctly and may realistically represent the complex SSI effects, allowing us to save time and reduce the complexities of more realistic three-dimensional highly non-linear models. In particular, this approach may be used when a preliminary design of complete (soil + foundation + structure) systems is required. In this regard, more comparative studies will be object of future work.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Conflicts of Interest

The author declares no conflict of interest.

References

  1. Kramer, S.L. Geotechnical Earthquake Engineering, Prentice Hall, Upper Saddle River. New Jersey 1996, 07458. [Google Scholar]
  2. Bhattacharya, K.; Dutta, S. Assessing lateral period of building frames incorporating soil-flexibility. J. Sound Vib. 2004, 269, 795–821. [Google Scholar] [CrossRef]
  3. Khalil, L.; Sadek, M.; Shahrour, I. Influence of the soil–structure interaction on the fundamental period of buildings. Earthq. Eng. Struct. Dyn. 2007, 36, 2445–2453. [Google Scholar] [CrossRef]
  4. Stewart, J.P.; Seed, R.B.; Fenves, G.L. Seismic Soil-Structure Interaction in Buildings. II: Empirical Findings. J. Geotech. Geoenvironmental Eng. 1999, 125, 38–48. [Google Scholar] [CrossRef] [Green Version]
  5. Veletsos, A.S.; Meek, J.W. Dynamic behaviour of building-foundation systems. Earthq. Eng. Struct. Dyn. 1974, 3, 121–138. [Google Scholar] [CrossRef]
  6. Veletsos, A.S.; Nair, V.V.D. Seismic Interaction of Structures on Hysteretic Foundations. J. Struct. Div. 1975, 101, 109–129. [Google Scholar] [CrossRef]
  7. Bielak, J. Dynamic behaviour of structures with embedded foundations. Earthq. Eng. Struct. Dyn. 1974, 3, 259–274. [Google Scholar] [CrossRef]
  8. Kwon, O.-S.; Kim, E.S. Evaluation of building period formulas for seismic design. Earthq. Eng. Struct. Dyn. 2010, 39, 1569–1583. [Google Scholar] [CrossRef]
  9. Hatzigeorgiou, G.D.; Kanapitsas, G. Evaluation of fundamental period of low-rise and mid-rise reinforced concrete buildings. Earthq. Eng. Struct. Dyn. 2013, 42, 1599–1616. [Google Scholar] [CrossRef]
  10. Stewart, J.P.; Fenves, G.L.; Seed, R.B. Seismic Soil-Structure Interaction in Buildings. I: Analytical Methods. J. Geotech. Geoenviron. Eng. 1999, 125, 26–37. [Google Scholar] [CrossRef]
  11. Avilés, J.; Pérez-Rocha, L.E. Diagrams of Effective Periods and Dampings of Soil-Structure Systems. J. Geotech. Geoenviron. Eng. 1999, 125, 711–715. [Google Scholar] [CrossRef]
  12. Avilés, J.; Pérez-Rocha, L.E. Soil-structure interaction in yielding systems. Earthq. Eng. Struct. Dyn. 2003, 32, 1749–1771. [Google Scholar] [CrossRef]
  13. Avilés, J.; Pérez-Rocha, L.E. Influence of Foundation Flexibility on Rμ and Cμ Factors. J. Struct. Eng. 2005, 131, 221–230. [Google Scholar] [CrossRef]
  14. Avilés, J.; Suárez, M. Effective periods and dampings of building-foundation systems including seismic wave effects. Eng. Struct. 2002, 24, 553–562. [Google Scholar] [CrossRef]
  15. Zaicenco, A.; Alkaz, V. Soil-structure interaction effects on an instrumented building. Bull. Earthq. Eng. 2007, 5, 533–547. [Google Scholar] [CrossRef]
  16. Ko, Y.-Y.; Chen, C.-H. Soil-structure interaction effects observed in thein situforced vibration and pushover tests of school buildings in Taiwan and their modeling considering the foundation flexibility. Earthq. Eng. Struct. Dyn. 2009, 39, 945–966. [Google Scholar] [CrossRef]
  17. Trifunac, M.D.; Todorovska, M.I.; Manić, M.I.; Bulajić, B.Đ. Variability of the fixed-base and soil-structure system frequencies of a building-The case of Borik-2 building. Struct. Control. Heal. Monit. 2008, 17, 120–151. [Google Scholar] [CrossRef]
  18. Michel, C.; Zapico, B.; Lestuzzi, P.; Molina, F.J.; Weber, F. Quantification of fundamental frequency drop for unreinforced masonry buildings from dynamic tests. Earthq. Eng. Struct. Dyn. 2010, 40, 1283–1296. [Google Scholar] [CrossRef]
  19. Luco, J.E. Bounds for natural frequencies, Dunkerley’s formula and application to soil–structure interaction. Soil Dyn. Earthq. Eng. 2013, 47, 32–37. [Google Scholar] [CrossRef]
  20. Xiong, W.; Jiang, L.-Z.; Li, Y.-Z. Influence of soil–structure interaction (structure-to-soil relative stiffness and mass ratio) on the fundamental period of buildings: Experimental observation and analytical verification. Bull. Earthq. Eng. 2016, 14, 139–160. [Google Scholar] [CrossRef]
  21. Oz, I.; Senel, S.M.; Palanci, M.; Kalkan, A. Effect of Soil-Structure Interaction on the Seismic Response of Existing Low and Mid-Rise RC Buildings. Appl. Sci. 2020, 10, 8357. [Google Scholar] [CrossRef]
  22. Abate, G.; Gatto, M.; Massimino, M.R.; Pitilakis, D.; Papadrakakis, M.; Fragiadakis, M. large scale soil-foundation-structure model in greece: Dynamic tests vs fem simulation. In Proceedings of the 6th International Conference on Computational Methods in Structural Dynamics and Earthquake Engineering (COMPDYN 2015); Institute of Structural Analysis and Antiseismic Research School of Civil Engineering National Technical University of Athens (NTUA), Rhodes Island, Greece, 15–17 June 2017; pp. 1347–1359. [Google Scholar]
  23. Castelli, F.; Grasso, S.; Lentini, V.; Sannito, M.S.V. Effects of Soil-Foundation-Interaction on the Seismic Response of a Cooling Tower by 3D-FEM Analysis. Geosciences 2021, 11, 200. [Google Scholar] [CrossRef]
  24. Gatto, M.P.A.; Montrasio, L.; Zavatto, L. Experimental Analysis and Theoretical Modelling of Polyurethane Effects on 1D Wave Propagation through Sand-Polyurethane Specimens. J. Earthq. Eng. 2021, 1–24, 1–24. [Google Scholar] [CrossRef]
  25. Lu, J.; Elgamal, A.; Yang, Z. OpenSeesPL: 3D Lateral Pile-Ground Interaction, User Manual, Beta 1.0. 2011. Available online: http://soilquake.net/openseespl/ (accessed on 6 November 2021).
  26. Forcellini, D. Soil-structure interaction analyses of shallow-founded structures on a potential-liquefiable soil deposit. Soil Dyn. Earthq. Eng. 2020, 133, 106108. [Google Scholar] [CrossRef]
  27. Forcellini, D. Seismic assessment of a benchmark based isolated ordinary building with soil structure interaction. Bull. Earthq. Eng. 2018, 16, 2021–2042. [Google Scholar] [CrossRef]
  28. Mazzoni, S.; McKenna, F.; Scott, M.H.; Fenves, G.L. Open System for Earthquake Engineering Simulation, User Command-Language Manual, OpenSees version 2.0.; Pacific Earthquake Engi-neering Research Center, University of California: Berkeley, CA, USA, 2009; Available online: http://opensees.berkeley.edu/OpenSees/manuals/usermanual (accessed on 6 November 2021).
  29. CEN. European Committee for Standardisation TC250/SC8/ Eurocode 8: Design Provisions for Earthquake Resistance of Structures. In Part 1.1: General Rules, Seismic Actions and Rules for Buildings, PrEN1998-1; CEN: Brussels, Belgium, 2003. [Google Scholar]
  30. Iervolino, I.; Maddaloni, G.; Cosenza, E. Eurocode 8 Compliant Real Record Sets for Seismic Analysis of Structures. J. Earthq. Eng. 2008, 12, 54–90. [Google Scholar] [CrossRef]
  31. Forcellini, D. Numerical simulations of liquefaction on an ordinary building during Italian (20 May 2012) earthquake. Bull. Earthq. Eng. 2019, 17, 4797–4823. [Google Scholar] [CrossRef]
Figure 1. Schematic two degrees of freedom (2-DOF) model.
Figure 1. Schematic two degrees of freedom (2-DOF) model.
Applsci 11 10472 g001
Figure 2. Three-dimensional (3D) mesh (Model 1).
Figure 2. Three-dimensional (3D) mesh (Model 1).
Applsci 11 10472 g002
Figure 3. Three-dimensional (3D) mesh (Model 2).
Figure 3. Three-dimensional (3D) mesh (Model 2).
Applsci 11 10472 g003
Figure 4. Three-dimensional (3D) mesh (Model 3).
Figure 4. Three-dimensional (3D) mesh (Model 3).
Applsci 11 10472 g004
Figure 5. Backbone curve (Infill soil).
Figure 5. Backbone curve (Infill soil).
Applsci 11 10472 g005
Figure 6. Backbone curve (S1 soil).
Figure 6. Backbone curve (S1 soil).
Applsci 11 10472 g006
Figure 7. Backbone curve (S2 soil).
Figure 7. Backbone curve (S2 soil).
Applsci 11 10472 g007
Figure 8. Backbone curve (S3 soil).
Figure 8. Backbone curve (S3 soil).
Applsci 11 10472 g008
Figure 9. (a). Soil S1: comparison (Model1, Model2 and Model3). (b). Soil S2: comparison (Model1, Model2 and Model3). (c). Soil S3: comparison (Model1, Model2 and Model3).
Figure 9. (a). Soil S1: comparison (Model1, Model2 and Model3). (b). Soil S2: comparison (Model1, Model2 and Model3). (c). Soil S3: comparison (Model1, Model2 and Model3).
Applsci 11 10472 g009
Figure 10. Selected input motions.
Figure 10. Selected input motions.
Applsci 11 10472 g010
Figure 11. (a) Results (near-field (NF) motions): comparison between top spectra: SSI model and EQ model. (b) Results (far-field (FF) motions): comparison between top spectra: SSI model and EQ model.
Figure 11. (a) Results (near-field (NF) motions): comparison between top spectra: SSI model and EQ model. (b) Results (far-field (FF) motions): comparison between top spectra: SSI model and EQ model.
Applsci 11 10472 g011aApplsci 11 10472 g011b
Table 1. Foundation parameters.
Table 1. Foundation parameters.
ParameterConcrete
Mass Density (kN/m3)24
Reference Shear Modulus (kPa)1.25 × 107
Reference Bulk Modulus (kPa)1.67 × 107
Table 2. Infill soil parameters.
Table 2. Infill soil parameters.
ParameterConcrete
Mass Density (kN/m3)17
Reference Shear Modulus (kPa)3.83 × 104
Reference Bulk Modulus (kPa)1.50 × 105
Table 3. Mesh characteristics.
Table 3. Mesh characteristics.
3D MeshDepth (m)Number of ElementsNumber of Nodes
Model 110.047,91653,868
Model 220.069,69676,313
Model 330.091,47698,758
Table 4. S1 soil characteristics.
Table 4. S1 soil characteristics.
ParameterS11S12S13S14S15
Mass Density (kN/m3)15.015.015.015.015.0
Reference Shear Modulus (kPa)1.50 × 1046.00 × 1041.35 × 1052.40 × 1053.75 × 105
Reference Bulk Modulus (kPa)7.00 × 1042.80 × 1056.30 × 1051.12 × 1061.75 × 106
Shear Wave velocity (m/s)100200300400500
Table 5. S2 soil characteristics.
Table 5. S2 soil characteristics.
ParameterS21S22S23S24S25
Mass Density (kN/m3)18.018.018.018.018.0
Reference Shear Modulus (kPa)1.80 × 1047.20 × 1041.62 × 1052.88 × 1054.50 × 105
Reference Bulk Modulus (kPa)8.40 × 1043.36 × 1057.56 × 1051.34 × 1062.10 × 106
Shear Wave velocity (m/s)100200300400500
Table 6. S3 soil characteristics.
Table 6. S3 soil characteristics.
ParameterS31S32S33S34S35
Mass Density (kN/m3)21.021.021.021.021.0
Reference Shear Modulus (kPa)2.10 × 1048.40 × 1041.89 × 1053.36 × 1055.25 × 105
Reference Bulk Modulus (kPa)9.80 × 1043.92 × 1058.82 × 1051.57 × 1062.45 × 106
Shear Wave Velocity (m/s)100200300400500
Table 7. Input motion characteristics.
Table 7. Input motion characteristics.
Input MotionsNamePGA (g)PGV (cm/s)PGD (cm)
NFSuperstition HillsBPT0.45112.0552.49
NFLoma PrietaBRN0.4855.7811.69
NFErzincanERZ0.5284.0027.68
NFCape Mendocino–PetroliaPET0.5948.3521.99
NFNorthridgeRRS0.83160.4029.72
FFFriuli TolmezzoATM0.4530.825.08
FFDuzce–BoluBOL0.4656.5323.09
FFChy ChyCHY0.18115.0968.81
FFNorthridge BHMUL0.2758.9913.16
FFLanders–YermoYER0.2451.4843.83
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Forcellini, D. A Novel Framework to Assess Soil Structure Interaction (SSI) Effects with Equivalent Fixed-Based Models. Appl. Sci. 2021, 11, 10472. https://0-doi-org.brum.beds.ac.uk/10.3390/app112110472

AMA Style

Forcellini D. A Novel Framework to Assess Soil Structure Interaction (SSI) Effects with Equivalent Fixed-Based Models. Applied Sciences. 2021; 11(21):10472. https://0-doi-org.brum.beds.ac.uk/10.3390/app112110472

Chicago/Turabian Style

Forcellini, Davide. 2021. "A Novel Framework to Assess Soil Structure Interaction (SSI) Effects with Equivalent Fixed-Based Models" Applied Sciences 11, no. 21: 10472. https://0-doi-org.brum.beds.ac.uk/10.3390/app112110472

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop