Next Article in Journal
Comprehensive Evaluation of Adhesive Compounds and Their Properties Involving Harrington’s Desirability Function
Previous Article in Journal
Investigation of Waste Steel Fiber Usage Rate and Length Change on Some Fresh State Properties of 3D Printable Concrete Mixtures
Previous Article in Special Issue
Comparison of Various Methods for Determining Dynamic Behavior of Voided Floor Slabs
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Numerical Investigation of Seismic Soil–Structure–Excavation Interaction in Sand

1
Structural Engineering Department, Zagazig University, Zagazig 44519, Egypt
2
Institute of Architectural Engineering, Faculty of Civil Engineering, Technical University of Košice, 04001 Kosice, Slovakia
3
Civil and Environmental Engineering Department, University of North Carolina, Charlotte, NC 28223, USA
*
Author to whom correspondence should be addressed.
Buildings 2025, 15(20), 3732; https://doi.org/10.3390/buildings15203732
Submission received: 9 September 2025 / Revised: 10 October 2025 / Accepted: 14 October 2025 / Published: 16 October 2025
(This article belongs to the Special Issue Constructions in Europe: Current Issues and Future Challenges)

Abstract

The dynamic loads affecting earth-retaining structures may increase in seismically active regions. Therefore, studying the soil–structure interaction among the soil, shoring systems, and adjacent structures is crucial. However, there is limited research on this important topic. This study investigates the seismic performance of a deep braced excavation and a nearby 10-story building in sandy soil formation. The main focus of this study is the consideration of the influence of varying foundation depths of adjacent structures on the seismic response of the shoring system and the performance of the shoring system and adjacent structure under different earthquake records. PLAXIS 2D software (Version 22.02) was used to carry out the numerical analysis. Sandy soil was modeled using the Hardening Soil with small-strain stiffness model (HS-small). Back analysis of observation data extracted from a real case study of a deep braced excavation in the central district of Kaohsiung City, adjacent to the O7 Station on the Orange Line of the Kaohsiung MRT system in Taiwan, was used to validate the numerical analysis. Beyond model validation, a parametric study was conducted to address the effect of the foundation level of the building adjacent to the excavation on both the seismic behavior of the shoring system and the structure itself, using the Loma-Prieta (1989) earthquake record. The parametric study was further extended to assess the responses of the shoring system and the adjacent structure under the influence of the earthquake records of Loma-Prieta (1989), Northridge (1994), and El-Centro (1940). The results show that the maximum lateral displacement of the diaphragm wall occurred at the top of the wall in all studied cases. The maximum dynamic bending moment in the retaining structure was more than three times the static one on average. In contrast, the dynamic shear force was more than 2.85 times the static one on average. In addition, the dynamic axial force of the first and second struts was 1.38 and 3.17 times the static forces, respectively. The results also reveal large differences in the behavior of the shoring system and the adjacent structure between the different earthquake records.

1. Introduction

A growing population and urban activities within large cities call for underground space to establish urban services, transportation infrastructure, parking areas, or other engineering works that necessitate deep vertical excavations near existing buildings. Retaining structures are essential for inhibiting large and unsafe soil displacement in the areas around excavation. These retaining structures are constructed and supported at different levels by horizontal struts and/or ground anchors, preserving excavation sides and limiting damage to the surrounding structures and utilities.
Braced excavations generate significant changes in stress and strain of the surrounding soil, resulting in permanent displacements to adjacent structures and/or infrastructure and causing severe damage in some cases [1,2,3,4,5]. In order to prevent damage to adjacent structures, it is essential to control the lateral deflections of the retaining structures as the excavation deepens [6,7]. Most studies have focused on the analysis of the behavior of deep excavations under static conditions using distinct methodologies. Empirical and semi-empirical approaches are commonly used, depending on extensive databases of excavation projects from various locations worldwide [1,2,8,9,10,11]. Additionally, numerical methods have been extensively employed to investigate various parameters affecting deep excavations and surrounding structures, including soil elasticity, creep, and the soil–wall interface [9,10,11,12,13,14,15,16,17,18,19,20,21,22,23,24]. Laboratory physical model tests were conducted by Chowdhury et al. [25] to investigate the static behavior of braced excavation.
Nevertheless, research has rarely been conducted on braced retaining structures subjected to seismic conditions. Callisto et al. [26] studied braced retaining walls numerically. Their results showed that propagation of the seismic waves through the soil causes quick mobilization of shear stress in the soil surrounding a propped cantilever wall. The wall experiences a significant increase in its lateral deflection and bending moment, along with an increase in the axial forces of the struts when compared to static conditions. Callisto and Soccodato [27] carried out numerical time history analysis, applying two earthquake time histories to cantilever retaining walls embedded in dry coarse-grained soil using FLAC 2D software (v. 5.0). It is suggested that critical horizontal acceleration can serve as a representation of the seismic resistance of a retaining wall’s permanent rigid body movement. This critical acceleration is calculated based on limit equilibrium analysis by an iterative method. A significant increase in bending moment during earthquakes is also observed owing to the instantaneous contact stress redistribution.
Konai et al. [28] investigated the seismic behavior of braced excavations with a single level of struts in dry sand using shaking table tests and numerical modeling using sinusoidal motions to simulate seismic loading. The results of the strut axial force, lateral displacement, and bending moment of the retaining structure from shaking table tests were compared to those of numerical models. The results demonstrated that in seismic conditions, wall stiffness has a larger influence on wall deflection and bending moments than strut forces.
Experimental studies have been conducted to interpret the behavior of different types of retaining structures under seismic loading [29,30,31,32,33,34]. Several centrifuge tests have been performed on cantilever and propped retaining walls in dry sand [35]. It has been observed that during earthquakes, no significant additional displacements would have occurred if the retaining wall had already been exposed to an earthquake that is more severe than the current one. Furthermore, it has been observed that the retaining wall undergoes nearly rigid permanent displacements even for maximum accelerations that are smaller than the critical limit equilibrium one.
Bahrami et al. [36] numerically studied the seismic performance of strutted diaphragm walls in dry sand using FLAC 2D software. The results indicated that under static loads, structures designed using the Peck method performed adequately. However, under seismic conditions, the wall bending moments and shear forces, as well as the struts’ axial forces, increased significantly, exceeding the permissible limits specified in ACI 318 [37] and AISCE [38]. Castaldo and De Iuliis [39] numerically studied the seismic vulnerability of existing buildings affected by adjacent deep excavations using PLAXIS 2D. By considering the non-linear dynamic properties of both soils and structures, it was found that the presence of adjacent excavation significantly increased the ductility demand and inter-story drift of nearby buildings. The use of Reduced Length Buckling-Restrained Braces (RLBRBs) as struts in deep excavations subjected to seismic loading was numerically investigated by Khademalrasoul et al. [40]. The results showed that incorporating RLBRBs in soft soils could enhance the overall stability of deep excavations under seismic loading. The effects of earthquakes on a 9.5 m deep anchored excavation with a diaphragm wall and an adjacent concrete building were analyzed using PLAXIS 2D software (Version 8) [41]. The results revealed that increasing the pre-stressing force of the anchor beyond a specific balanced value serves as a constant final anchor force after the earthquake. The SOIL-STRUCT tool was applied to assess a 10-story building adjacent to a deep excavation of 28 m depth and supported by a 0.7 m thick diaphragm wall [42]. The analysis results showed that roof displacements with smooth asymptotes are larger than those with concave downward asymptotes. Additionally, roof displacements are significantly influenced by an increase in the yield stress of steel, whereas this effect is less pronounced for concrete building frames. Chowdhury et al. [43] numerically investigated the behavior of a strutted retaining structure under seismic conditions using FLAC-2D (v.5.0). It was observed that during an earthquake, the wall’s lateral displacement is the most affected parameter, and the struts’ axial force is the least affected parameter. Yeganeh et al. [44] used the finite difference method in FLAC-2D (version 7.0) to investigate building–excavation interaction and structure models with a fixed base.
To better understand the excavation–structure interaction, this research investigates the performance of deep braced excavations in sand under seismic loads, considering the interaction with adjacent structures using a numerical approach. The hardening soil model with small-strain stiffness (HS-small) is employed to simulate the non-linear and hysteretic behavior of the soil in both static and dynamic conditions. Model validation is performed using a back analysis of a well-documented case study of a deep braced excavation in sandy soil, as cited in [45]. Following validation, a comprehensive parametric study is performed on deep braced excavation in sandy soil with an adjacent structure in static and dynamic cases. This study focused on the response of strutted diaphragm walls and surrounding structures under different earthquake records. The acceleration time histories of Loma-Prieta (1989), Northridge (1994), and El-Centro (1940) earthquakes are utilized in this study. Moreover, the foundation level is also investigated under the effect of the Loma-Prieta (1989) earthquake record. This study discusses the diaphragm wall lateral displacements, settlement trough behind the wall, straining actions of the diaphragm wall, axial forces of the struts, and maximum story displacement of the adjacent structure under static and seismic conditions to evaluate and assess the seismic performance of both the deep braced excavation and the adjacent structure.

2. Numerical Modeling

Plain strain analysis is used to model the soil behavior statically and dynamically. The HS-small strain model, which is implemented in PLAXIS 2D (Version 22.02) based on research of Benz [46], is used to model soil behavior to account for the stress dependency of stiffness and variance in the ratio of secant shear stiffness (Gs/G0) with increasing shear strain (γ). The HS-small strain model uses the hyperbolic approximation of the stress–strain curve proposed by Hardin [47], which was modified then by Santos et al. [48], given as:
τ = G s γ =   G 0   γ 1 + 0.385 γ γ 0.70
where τ is the soil shear stress, and Gs is the secant shear modulus related to the shear strain (γ). γ0.7 is the shear strain when Gs is reduced to 70% of G0. In the HS-small model, the stress dependency of the shear modulus G0 is considered with the power law, which matches the one used for the other stiffness parameters, as indicated in Equation (2). The threshold shear strain γ0.7 is taken independently of the mean stress, as follows:
G 0 =   G 0 r e f c   c o s   σ 3   s i n   c o s +   p r e f   s i n m
Moreover, the model uses Mohr–Coulomb strength parameters (cohesion c and friction angle ϕ), along with soil stiffness, which is defined by three input values: E50 (secant stiffness modulus at 50% deviatoric stress (qf)), given by Equation (3), Eoed (1D oedometer loading modulus), and Eur (unloading–reloading modulus).
E 50 = E 50 r e f c   c o s σ 3   s i n   c o s + p r e f   s i n m
where E 50 r e f is the secant modulus at the reference confining stress, pref (equal to 100 kPa), and σ 3 is the minor principal stress. Here, m is the exponent factor for stress dependency, and it is taken to be equal to 0.50 for sandy soil, as proposed by Benz [46]. The default values of Eoed = E 50 r e f and Eur = 3 E 50 r e f , as proposed by [49,50].
Despite using the HS-small constitutive model, which captures energy dissipation through hysteric behavior, a small Rayleigh damping of 2% is used in this study during the dynamic analysis to eliminate the high-frequency noise resulting from the numerical integration. In finite element modeling, the formulation of damping is often based on mass and stiffness and is defined as proposed by [51,52,53]:
C = α M + β [ K ]
where [M] is the mass matrix, [K] is the stiffness matrix, and α and β are the Rayleigh damping coefficients, given by:
α = 2 ς ω 1 ω 2 ω 1 + ω 2
β = 2 ς ω 1 + ω 2
where ς is the damping ratio, which is 2% in this study, and ω1 and ω2 are considered to be the target angular frequencies corresponding to the target circular frequencies f1 and f2, respectively. The first frequency, f1, is equal to the fundamental frequency of the whole soil layer, which is defined as the frequency at which the most significant amplification can be expected, and it is calculated by Equation (7) based on the formula proposed by Kramer [54]. The second frequency is the closest odd number given by the ratio of the fundamental frequency of the input signal at the bedrock and the fundamental frequency of the whole soil layer, as proposed by Hudson [55].
f n = V s 4 H

2.1. Meshing Considerations

The discretization of real models influences the responses of both linear and non-linear finite element models. Dynamic analysis requires that the maximum element dimension be limited to one-eighth to one-fifth of the shortest wavelength considered in the analysis, as suggested by other researchers [56,57]. In this research, 15-node triangular soil elements are used to model the soil continuum. The dimension of the elements is adjusted, and the local refinement of the meshes is implemented to obtain an appropriate value for the average length of the element dimension to increase the accuracy of the results and so that the seismic waves can propagate without numerical errors.

2.2. Boundary Conditions

During static analysis, the movement of side boundaries is restrained horizontally, while movements for the bottom boundary of the model are restrained horizontally and vertically. As for the dynamic analysis, a repeatable tied degree of freedom boundary condition, which was proposed by Zienkiewicz et al. [58], is utilized to simulate the side boundaries of the model to connect the nodes at the same elevation, causing the same displacements. On the other hand, a compliant base boundary condition, as proposed by Joyner and Chen [59], is applied at the bottom of the model to consider the absorption and application of the dynamic loads.

3. Verification of a Case Study on Deep Excavation in Sand [45]

A real case study of a deep braced excavation in Kaohsiung, Taiwan, presented by [45], as shown in Figure 1, is utilized to validate the numerical model before dynamic analysis to assess the model’s accuracy in capturing the behavior of real braced excavation. The excavation dimensions are 70 m in length and 20 m in width, with a 16.8 m depth. A 0.90 m diaphragm wall is used as a retaining structure, supported by four levels of steel struts spaced 5.50 m apart horizontally. The excavation is executed in medium-dense to dense silty–sandy soil with bands of low-plasticity clay (CL). Additionally, before excavation, the groundwater table is nearly 2.0 m below the ground surface and is lowered by 1.0 m before any excavation stage to provide a dry environment for the construction process. Details of the soil lithology, excavation levels, and strut levels have been provided in Hsiung and Dao [45].
PLAXIS 2D FE software (Version 22.02) developed by [60] is used to perform back analysis of the intended case study. A plane-strain FEM model is carried out by adopting a 15-node triangular element to simulate the deformations and stresses in the soil volume. Hardening soil with a small strain model (HS-small) is utilized to perform this study, as recommended by Hsiung and Dao [45], to represent the behavior of the sand layers, whereas the behavior of the clay bands is modeled using the elastic–perfectly plastic Mohr–Coulomb (MC) model. The soil properties analyzed in this study followed the recommendations of Hsiung and Dao [45]. The diaphragm wall and steel struts are modeled as linear elastic materials, adopting plate elements for the diaphragm wall and fixed-end anchors for the steel struts. A Poisson’s ratio of 0.2 is used for both the diaphragm wall and steel struts. The diaphragm wall elastic modulus is 24.8 × 106, while the elastic modulus of steel struts is 2.1 × 105 (MPa). The stiffness of the diaphragm wall and steel struts is decreased by 30% and 40%, respectively, from their nominal values to account for bending moment-induced cracks in the diaphragm wall, repeated use, and improper installation of steel struts, as suggested by Ou [61]. More detailed data about the soil properties, diaphragm wall, and steel struts have been provided by Hsiung and Dao [45].
The verification model’s dimension is extended to a depth of 60 m below the ground surface to comply with the site investigation program. The length of the side boundary of the model to the diaphragm wall is taken to be seven times the maximum excavation depth, as suggested by Khoiri and Ou [62] for deep excavations in sands, as shown in Figure 2a. The mesh discretization shown in Figure 2b is set to global medium, and local refinement is implemented to enhance the accuracy of the analysis.

Verification Results of the FEM Model

Figure 3 shows the comparison of the diaphragm wall lateral displacements resulting from the finite element simulation at the final stage of the excavation and the measured ones from the field data. The measured and predicted lateral displacements at the final stage of excavation showed good agreement. The measured lateral displacement for the first part of the wall, up to 4.0 m of the wall, is less than the predicted one by about 8.0 to 12.0%. However, the predicted and measured displacements seem to be identical up to the final excavation level at 16.80 m, with a difference not exceeding 1%. Moreover, the predicted displacement beyond the final excavation level is further increased by an increasing rate from 1.11% to about 11.40%. The results imply that the HS-small strain model achieved the best prediction of the wall lateral deflections and ground surface settlement in such simulations of these cases, and this is consistent with the findings of previous research [45,62,63]. The ground surface settlement adjacent to the diaphragm wall is also investigated and compared to that of field measurements, as shown in Figure 4. It is demonstrated that the HS-small strain constitutive soil model predicts the settlement in reasonable agreement with measured field data, especially beside the diaphragm wall. However, the predicted settlement seems to be larger than the measured ones as the excavation goes deeper to the final level, but the settlement trough still has the same pattern, as proposed by Hsieh and Ou [8] for braced excavations.

4. Dynamic Modeling of the Shoring System and Adjacent Structure

This research focuses on a braced excavation with 10.00 m depth that is modeled in a dense sandy soil with a 60.00 m deep overlying bedrock and a surrounding structure composed of a basement and nine replicated stories, each with two spans 5.00 m. Side boundaries of the model, as shown in Figure 5, which are chosen to be at 7B (where B is the width of excavation) from the left and right walls. According to ASCE 7-16 [64], the site classification is performed according to the SPT (N) values. For this study, the chosen N-value locates the soil in a seismic category (D), as indicated in ASCE 7-16 [64], and the shear wave velocity, Vs is chosen to be 200 m/s in this study. The soil properties used in this study are presented in Table 1. The shoring system and surrounding structure are modeled statically and dynamically using different earthquake records. Static and dynamic behavior is investigated.
In this analysis, linear elastic plate elements are used to model the diaphragm wall with an elasticity modulus of 26.00 GPa, which is calculated according to ACI-318 [37], and a Poisson’s ratio of 0.20. A Rayleigh damping of 5% is considered during seismic analysis.
Two levels of struts with horizontal spacing in the longitudinal direction equal to 5.00 m are used in this study. The first level is installed at 2.00 m from the ground surface, and the second one is installed at 6.00 m from the ground surface, as recommended by the design guide of Chowdhury et al. [15]. Linear elastic node-to-node anchors are used to model the struts with axial stiffness of 25 × 106 and 50 × 106 kN/m2, respectively.
A reinforced concrete structure composed of two bays, each with 5.00 m width, with one basement with 3.25 m height and nine typical stories with 3.00 m height each, is set adjacent to the shoring system. The equivalent dead and live loads are considered to be 10 kN/m’/storey. The structural elements are simulated as linear elastic plate elements with axial and flexural stiffness to model columns, walls, beams, and slabs of an equivalent thickness of 0.30 m, while the raft foundation has a 1.00 m thickness. Modal analysis using CSI-SAP2000 (Computers and Structures Inc., Berkeley, CA, USA) identified the first and second modes of vibration as dominant, representing approximately 90% of the structure’s dynamic mass participation. These modes have time periods of 2.47 s and 1.07 s, respectively, following the regulations stated in ASCE [64]. Using these time periods and a 5% damping ratio, the Rayleigh damping constants α and β are calculated using Equation (8) as 0.1984 and 0.008556, respectively, as presented in the methodology proposed in references [52,65].
ς i = 1 2 ( α ω i + β ω i )
where α is the mass-proportional damping constant, β is the stiffness-proportional damping constant, ςi is the damping ratio, and ω is the angular frequency, in rad/s. Interface elements are added to simulate soil–structure interaction between the shoring system, basement walls, the raft, and the surrounding soil. Similar results of recent research have been reported [66,67,68,69,70].

5. Parametric Study

The main objective of this study is to investigate the seismic performance of deep braced excavations in sand, considering the presence of a building adjacent to the excavation, to interpret the soil–structure interaction in such complex cases. All the studied parameters are summarized in Table 2.
First, static analysis is performed, considering that the diaphragm wall is wished-in-place and no deformation has accumulated from the adjacent wall construction. Then, the excavation starts from the pre-defined levels up to the final excavation level. After each excavation stage, the bracing system is also installed at the proposed locations. Eventually, dynamic analysis is carried out, imposing the seismic loads as a prescribed displacement in the horizontal direction (i.e., Global X-direction) using a dynamic multiplier that represents the time history of the seismic ground motions.
In the dynamic stage of construction, the dynamic boundary conditions and damping are executed before running the analysis, as explained in detail in Section 2.1 and Section 2.2. The implicit Newmark time integration scheme is employed to solve the dynamic equilibrium equations of the system. The dynamic time intervals are used as the total time of the earthquake record. The time steps are also adjusted to match the number of data points that represent the dynamic multiplier of the ground motion. To control the accuracy of the numerical integrations, the Newmark time integration coefficients α and β are chosen to achieve the following conditions.
β 1 2 , α 1 4 1 2 + β 2
The Newmark scheme with α = 0.25 and β = 0.5 (average acceleration method) is utilized, as recommended by PLAXIS [71].

Characteristics of the Input Ground Motions

Acceleration time histories of three well-known earthquakes of Loma-Prieta (1989), Northridge (1994), and El-Centro (1940), retrieved from the Pacific Earthquake Engineering Research Center (PEER) [72], are considered and applied on bedrock after the last stage of excavation, as shown in Figure 6, Figure 7 and Figure 8. The main characteristics of the earthquake records are given in Table 3.

6. Analysis Results and Discussion

The parametric study aims to assess the shoring system design parameters and the behavior of adjacent structures in static and seismic analyses.

6.1. Effect of the Adjacent Structure’s Foundation Level on the Shoring System–Structure Interaction

6.1.1. Displacement–Time History (DTH)

Figure 9a–c present the displacement–time history for the top of the diaphragm wall and the top roof of the adjacent structure, respectively, for different foundation levels of the structure adjacent to the shoring system. The free-field ground surface response (point B in Figure 5) for different foundation levels, as shown in Figure 9a, indicates that the response for all foundation levels is the same as the original ground motion at the base, with a time difference of 0.33 s and maximum amplification ratios of 232%, 209%, and 203% for 1.00 m, 3.00 m, and 5.00 m depths, respectively. It was also noticed that permanent lateral displacements of the ground surface occur in the same direction as the original ground motions of 62.92 mm, 16.71 mm, and 0.56 mm for the depths of 1.00 m, 3.00 m, and 5.00 m, respectively. The results indicate that as the building foundation level increased, the permanent displacement of ground surface decreased significantly by 73.40% as the foundation depth increased from 1.00 to 3.00 m and by 96.60% as the foundation depth increased from 3.00 to 5.00 m.
Figure 9b demonstrates the response of the top of the diaphragm wall (point C in Figure 5) for different foundation levels. The results indicate a time difference of 0.23 s at the top point of the diaphragm wall and an amplification of the incident waves by 331%, 385%, and 388% for foundation depths of 1.00, 3.00, and 5.00 m, respectively. At the end of the earthquake, the crest of the diaphragm wall experienced a permanent lateral movement of −87.89, −128.03, and −138.55 mm for foundation levels of 1.00, 3.00, and 5.00 m, respectively. The response of the top point of the diaphragm wall indicates that the wall displacement at the top increased as the structure’s foundation level increased. As the foundation level increased from 1.00 to 3.00 m, the permanent displacement increased by 45.70%, and it increased by 8.20% as the depth increased from 3.00 to 5.00 m. Deeper foundations caused an increase in lateral displacement at the top of the diaphragm wall because of an amplification effect. The large amplification of seismic waves led to large lateral displacements of the foundation raft towards the excavation area, and this, in turn, exposed the diaphragm wall to extra lateral displacements. Moreover, ponding between the basement floor and the diaphragm wall may have caused this large displacement of the diaphragm wall.
Figure 9c shows the response at the top roof of the structure adjacent to the shoring system (point D in Figure 5). A significant increase in the displacement at the top roof was noticed as the foundation level of the structure increased. The structure underwent large permanent lateral displacements at the top roof at the end of ground shaking. The lateral displacements were amplified by 41.22, 36.37, and 25.80 times compared to the original ground motion for foundation levels of 1.00, 3.00, and 5.00 m, respectively. The top roof exhibited a permanent lateral displacement of −348.32, −406.54, and −446.15 mm at the end of shaking for foundation levels of 1.00, 3.00, and 5.00 m, respectively. As the foundation level increased from 1.00 to 3.00 m, the permanent displacement increased by 16.70%, and a 9.70% increase in permanent displacement was noticed as the depth increased from 3.00 to 5.00 m.

6.1.2. Settlement Trough Beside the Shoring System

Figure 10a,b represent the results of the settlement trough in static and dynamic cases for different foundation levels of the structure adjacent to the shoring system, respectively. The static settlement trough in Figure 10a shows that as the foundation level of the structure increased, the settlement trough decreased significantly. Moreover, the extent of the settlement trough decreased as the foundation level increased, and for a structure’s foundation level of 1.00 m, the settlement trough extended to five times the ultimate excavation depth (5 H) until no significant changes were noticed. On the other hand, for foundation levels of 3.00 m and 5.00 m, the settlement trough extended to two times the ultimate excavation depth (2 H) until no significant changes were noticed. The results indicate that the structure was exposed to differential settlement due to excavation and that the differential settlement decreased as the foundation level increased.
On the contrary, the dynamic settlement trough, as shown in Figure 10b, behaves differently from the static one. The results indicate that the settlement trough and the differential settlement of the structure increase as the foundation level of the structure increases from 1.00 m to 3.00 m, while as the foundation level increased from 3.00 m to 5.00 m, the settlement trough and differential settlement decreased once more. The dynamic trough extends to three times the final excavation depth (3 H), and beyond this distance, no significant changes in the settlement were noticed. The results of the dynamic trough indicate that the ground surface was exposed to uniform permanent vertical settlement of about 5.0 cm for all foundation levels.
Furthermore, for the adjacent structure, the deflection and cracked conditions of the structure components have a direct association with particular parameters, such as differential settlement, relative rotation, and the deflection ratio [61]. Rankine [73] provided guidelines for how the maximum settlement and maximum slope of a building affect its potential damage. The results of maximum static vertical settlement of the building adjacent to the excavation indicate that the building is in a moderate-to-high-risk category based on the Rankine guidelines [73]. However, in the case of dynamic vertical settlement, the building falls under the high-risk category, in which structural damage to the building is expected.

6.1.3. Lateral Displacement of the Shoring System

Figure 11a,b show the static and dynamic lateral displacements of the diaphragm wall for different foundation levels of the adjacent structure, respectively. The static lateral displacements of the diaphragm wall presented in Figure 11a show that the lateral displacements decreased significantly as the foundation level of the structure increased. The lateral displacement at the crest of the diaphragm wall decreased by 60.70% as the foundation level increased from 1.00 m to 3.00 m and decreased by 18.10% as the foundation level increased from 3.00 m to 5.00 m, respectively. However, the lateral displacement of the wall in the dynamic analysis behaves differently, as shown in Figure 11b. The results of the dynamic analysis show that the maximum lateral displacement occurred at the top of the diaphragm wall in all the studied cases. Moreover, as the structure’s foundation level increased, the lateral displacement of the wall increased significantly. As the structure’s foundation level increased from 1.00 m to 3.00 m, the dynamic lateral displacement at the top of the diaphragm wall increased by 47.10%, while as the foundation level increased from 3.00 m to 5.00 m, the lateral displacement increased by 9.50%. On the other hand, as the foundation level increased from 3.00 m to 5.00 m, the dynamic lateral displacements beyond the bottom of the excavation level decreased by 8.40%.
Figure 12 presents a comparison of the dynamic and static lateral displacements at the top of the diaphragm wall and at the bottom of the excavation level for different foundation levels of the structure adjacent to the shoring system. The dynamic lateral displacements at the top of the wall are 1.15, 4.32, and 5.77 times the static ones for the structure’s foundation levels of 1.00 m, 3.00 m, and 5.00 m, respectively. Moreover, the dynamic lateral displacement at the bottom of the excavation level is 1.43, 3.37, and 4.15 times the static ones for the structure’s foundation levels of 1.00 m, 3.00 m, and 5.00 m, respectively.
The significant increase in the dynamic values is attributed to an amplification of the seismic waves while being transmitted from the base of the model to the ground surface. This behavior implies the vital role of foundation levels of structures adjacent to deep excavation projects in the behavior of the shoring system in dynamic cases. Moreover, modeling the structure adjacent to the excavation as a real structure, not only as a surcharge load, is very important and shows the significance of seismic soil–structure interaction (SSSI) in complex problems, such as deep excavations.
The damage degree of the diaphragm wall can be identified based on the residual lateral displacement of the diaphragm wall, as proposed by Gazetas [74]. The results show that the maximum dynamic lateral displacement at the top of the diaphragm wall falls between 0.087 and 0.14 m, indicating noticeable damage to the wall, according to the findings of Gazetas [74].

6.1.4. Straining Actions of the Diaphragm Wall and Its Supporting Struts

The foundation level of the structure adjacent to the deep excavation affects the straining actions of the diaphragm wall and axial forces of the supporting struts as well. Static and dynamic bending moments along the wall are presented in Figure 13a,b, respectively. The result pattern of the static bending moments shown in Figure 13a for the foundation level of 1.00 m is different from other foundation levels of 3.00 m and 5.00 m, especially at the second strut location. The bending moment at the second strut location decreases as the foundation level increases. The maximum static bending moment occurs at 87.5% H for all foundation levels. Moreover, it was also noticed that the maximum moment beyond the bottom of the excavation level occurred at depths of 15 to 25% D. On the other hand, the dynamic bending moment shown in Figure 13b reveals that the maximum bending moment takes place at the location of the second strut.
Furthermore, the static bending moment increased by 37.51% as the foundation level increased from 1.00 m to 3.00 m, while it decreased by 11.48% as the foundation level increased from 3.00 m to 5.00 m. Moreover, the dynamic bending moments decreased by 12.17% as the foundation level increased from 1.00 m to 3.00 m, while they increased by 7.96% as the foundation level increased from 3.00 m to 5.00 m.
Figure 13c,d show the static and dynamic shearing forces along the diaphragm wall for different foundation levels of the structure, respectively. As shown in the figures for both static and dynamic cases, the maximum shear force takes place at the location of the second strut. The results show that in the static case, the maximum shear force decreased by 40.15% as the foundation level increased from 1.00 m to 3.00 m, while it increased by 9.23% as the foundation level increased from 3.00 m to 5.00 m. Moreover, the maximum dynamic shear force increased by 1.39% and 5.40% as the foundation level increased from 1.00 m to 3.00 m and from 3.00 m to 5.00 m, respectively.
Figure 14 illustrates the ratio of the maximum dynamic bending moments to the static ones. The maximum dynamic moment is 4.29, 2.74, and 3.34 times the static one for the foundation levels of 1.00 m, 3.00 m, and 5.00 m depths, respectively. In addition, the figure also presents the ratio of the maximum dynamic shear force to the maximum static shear force. The results show that the dynamic shear force is 2.02, 3.41, and 3.29 times the static one for foundation levels of 1.00 m, 3.00 m, and 5.00 m depths, respectively.
The effect of the structure’s foundation level on the struts’ axial forces in both static and dynamic cases is illustrated in Figure 15. The results show that the static axial force of the first and second struts decreased by 29.40% and 48.00% as the foundation level increased from 1.00 m to 3.00 m, respectively. Moreover, the axial force of the first and second struts decreased by 24.00% and 11.80% as the foundation level increased from 3.00 m to 5.00 m, respectively.
Figure 15 also demonstrates that the dynamic axial forces of the first and second struts decreased by 25.00% and 11.00% as the foundation level increased from 1.0 m to 3.0 m, respectively. Furthermore, the dynamic axial force of the first strut increased by 68.00% as the foundation level increased from 3.00 m to 5.00 m, while it decreased by 7.00% for the second strut as the foundation level increased from 3.00 m to 5.00 m.
Figure 16 shows the effect of dynamic forces on axial forces of the struts for different foundation levels of the adjacent structure, compared to the static ones. The results show that for the first strut, there is nearly no change in the axial force in dynamic cases for a foundation level of 1.00 m, while it increased by 1.06 and 2.33 times the static ones as the foundation levels of the structure increased to 3.00 m and 5.00 m, respectively. On the other hand, the dynamic axial force of the second strut is 2.20, 3.77, and 3.96 times the static ones for foundation levels of 1.00 m, 3.00 m, and 5.0 m, respectively.

6.1.5. Response of the Structure Adjacent to the Shoring System

Figure 17a depicts the lateral displacement of the structure adjacent to the shoring system because of excavation for different foundation levels of the structure. The results indicated in this figure show that as the depth of the foundation increased, the lateral displacement of the structure decreased significantly due to the lateral confinement of the soil to the structure’s embedded depth and rigid diaphragm wall, which prevents excessive lateral displacements. At the base of the structure, as the foundation level increased from 1.0 m to 3.0 m depth, the lateral movement decreased by 52.20%, and as the foundation level increased from 3.00 m to 5.00 m, the lateral displacement decreased by 24.30%. Furthermore, the lateral displacement at the top roof of the structure decreased by 55.50% and 9.20% as the foundation level increased from 1.00 m to 3.00 m and from 3.00 m to 5.00 m depth, respectively.
The results of the lateral displacement of the building under dynamic load are shown in Figure 17b. A large amplification of the seismic waves and the stress relief due to excavation and, consequently, large lateral displacement of the diaphragm wall caused a large displacement of the adjacent structure. At the structure’s base, the dynamic lateral displacement increased significantly by 127.80% as the foundation level increased from 1.0 m to 3.0 m depth. However, as the foundation level increased from 3.0 m to 5.0 m depth, the lateral displacement increased by 15.90%. Moreover, the lateral displacement at the structure’s top roof increased by 19.70% as the foundation level increased from 1.0 m to 3.0 m depth, while it increased by 22.90% as the foundation level increased from 3.0 m to 5.0 m depth.
Figure 18 presents the ratio of the lateral displacement in dynamic cases to that in static cases at the base and top of the structure for different foundation levels of the structure. At the base of the structure, the dynamic lateral displacement is 0.64 times the static one at the foundation level of 1.00 m, whereas it is 3.06 and 4.69 times the static one for foundation levels of 3.00 m and 5.00 m, respectively. On the other hand, the dynamic lateral displacement at the top roof is 2.30, 6.18, and 8.37 times the static ones for the foundation levels of 1.00 m, 3.00 m, and 5.00 m, respectively. The results show that the top roof of the structure experienced large lateral displacements in dynamic cases due to large amplifications of the seismic waves while being transmitted from the base to the roof.

6.2. Effect of Earthquake Records on Shoring System–Structure Interaction

6.2.1. Displacement–Time History (DTH)

Figure 19a–c show the displacement–time history of the original ground motion at the base of the model (point A), free-field ground surface displacement (point B), top point of the diaphragm wall (point C), and top roof of the structure adjacent to the shoring system (point D), all shown and presented in Figure 5, respectively.
The maximum displacement amplitude of the Loma-Prieta earthquake, as shown in Figure 19a, is amplified by 2.10, 2.94, and 2.97 times the base ground motion for ground surface in the free field, on top of the diaphragm wall, and on the roof of the structure, respectively. Furthermore, the ground surface, top of the diaphragm wall, and roof of the structure have a permanent displacement of 18.37, −60.73, and −148.39 mm, respectively, at the end of the earthquake.
The ground surface responses, top of the diaphragm wall, and top roof of the structure under the Northridge earthquake record are illustrated in Figure 19b. The free-field ground surface, top of the wall, and top roof of the structure amplified the displacements to 1.70, 2.28, and 5.58 times the original one that was imposed at the base of the model, respectively. The results indicate that the ground surface, top of the wall, and top of the structure undergo permanent displacements of 112.31, −47.95, and −215.48 mm at the end of ground shaking, respectively.
The results of the El-Centro earthquake record are presented in Figure 19c. The amplifications of the ground surface, top of the wall, and top of the structure are 1.97, 2.23, and 2.89 times the original displacement of the earthquake, respectively. Moreover, the ground surface, top of the wall, and top of the structure experience permanent displacements of 33.71, −93.80, and −230.45 mm at the end of ground shaking, respectively.
The amplification ratio and the permanent displacement at the end of the earthquakes’ ground motions are illustrated in Table 4. The results indicate that the Loma-Prieta earthquake had a minimal effect on the ground surface displacement, while the Northridge earthquake had the largest effect on the ground surface displacement. However, the El-Centro and Northridge earthquakes had a noticeable effect on the diaphragm wall and the adjacent structure.

6.2.2. Lateral Displacement of the Shoring System

Static and dynamic lateral displacements of the diaphragm wall using the previously mentioned earthquake records are shown in Figure 20. The results indicate that the lateral displacement in static analysis has a deep inward movement pattern, and the maximum displacement occurred near the final excavation depth. Otherwise, the maximum lateral displacement in dynamic analysis takes place at the top of the wall for all earthquake records.
The ratio of the maximum lateral displacement at the top of the diaphragm wall in the dynamic analysis to that of the static analysis for different earthquake records is given in Figure 21. The results indicate that the lateral displacement at the top of the diaphragm wall in the dynamic analysis is 3.92, 3.67, and 6.18 times the static ones for the Loma-Prieta, Northridge, and El-Centro earthquake records, respectively. The results show that the El-Centro earthquake affected the wall more than the other two earthquakes, which is attributed to the fact that the El-Centro earthquake has a long duration, high frequency, and long period of strong shaking. Moreover, the predominant period of the El-Centro record is larger than the other two records, which implies that the predominant frequency of El-Centro is small and may be near the natural frequency of the diaphragm wall; this causes large displacements of the diaphragm wall, which may be attributed to some sort of resonance phenomenon.
The results show that the maximum dynamic lateral displacement at the top of the diaphragm wall falls between 0.048 and 0.094 m, indicating negligible damage to the wall itself, with noticeable damage to related structures at struts, according to the findings of Gazetas [74].

6.2.3. Settlement Trough Beside the Shoring System

Settlements behind the diaphragm wall resulting from static and dynamic analyses during the aforementioned earthquakes are depicted in Figure 22. The maximum settlement occurs along the diaphragm wall and extends to approximately twice the ultimate excavation depth, beyond which no notable changes in settlements are seen.
Settlements under dynamic analysis showed 5.10, 7.0, and 7.50 times the maximum static settlement for the Loma-Prieta, Northridge, and El-Centro records, respectively. The results of the dynamic study demonstrate that the ground surface experiences a uniform settlement of approximately 50 mm, equivalent to 0.50% of the final excavation depth. Furthermore, due to the existing buildings in proximity to the wall and the lateral displacements of the shoring system, along with stress release from the excavation operation, the structure is subjected to differential settlement and tilting towards the excavation region.
The results of maximum dynamic vertical settlement of the building adjacent to the excavation indicate that the building falls in a high-risk category, in which structural damage to the building is expected based on the Rankine guidelines [73].

6.2.4. Straining Actions of the Diaphragm Wall and Its Supporting Struts

Figure 23a,b illustrate the resulting straining actions along the diaphragm wall in both static and dynamic analyses. Dynamic bending moments along the diaphragm wall, as shown in Figure 23a, have the same pattern for different earthquakes, but the amplitude changes due to earthquake intensity and duration. The bending moment under the Loma-Prieta earthquake has the smallest value compared to the other earthquake records. The bending moment distribution pattern in the dynamic cases looks different from that in the static case due to the nature of the dynamic loads, whose values and direction change with time. The maximum static bending moment takes place at 87.5% H from the top of the diaphragm wall, whereas the maxium dynamic bending moment above the bottom of the excavation level occurs at the position of the second strut level and below the bottom of the excavation level by nearly 30% D. It was noticed that below the excavation level, El-Centro has a larger effect than other earthquakes.
Figure 23b presents the shearing forces in static and dynamic cases along the diaphragm wall. It was noticed that the shearing forces of different earthquakes have the same pattern. The shearing forces under the Loma-Prieta earthquake have the smallest values compared to the other earthquake records. The maximum shearing forces occurred at the location of the second strut.
Figure 24 shows the ratio of the maximum dynamic straining actions to the maximum static ones. The maximum wall bending moments under the Loma-Prieta, Northridge, and El-Centro records are 1.90, 3.04, and 2.93 times the static ones, respectively. Moreover, the maximum wall shear forces under the Loma-Prieta, Northridge, and El-Centro records are 2.34, 3.24, and 3.56 times the static ones, respectively. This increase emphasizes the importance of the seismic design of such structures.
Figure 25 shows the resulting axial force of the struts after static and dynamic analyses. As expected, the seismic loads cause large axial forces in the struts. The second strut exhibits a larger axial force than the first one due to the large earth pressure and shearing forces of the wall at this location.
A comparison of the axial forces of the struts under static and dynamic analyses is shown in Figure 26. The axial force in the first level of struts is 1.52, 1.94, and 1.64 times the static one under the Loma-Prieta, Northridge, and El-Centro earthquakes, respectively. Moreover, the axial force in the second level of struts is 4.39, 7.04, and 6.55 times the static one under the Loma-Prieta, Northridge, and El-Centro earthquakes, respectively. The results indicate that the Northridge and El-Centro earthquakes have a larger effect on the struts’ axial force than the Loma-Prieta earthquake.

6.2.5. Response of the Structure Adjacent to the Shoring System

Figure 27 presents the structure’s lateral displacement in static and dynamic cases. The structure undergoes a lateral displacement in the static case due to the excavation process because of stress relief resulting from removing soil adjacent to the structure and lateral deformation of the diaphragm wall due to induced lateral earth pressure.
A comparison of the resulting maximum story displacement at the base and top of the structure in static and dynamic cases is shown in Figure 28. The lateral displacement of the base of the structure is 3.87, 1.75, and 4.12 times the static one under the Loma-Prieta, Northridge, and El-Centro earthquakes, respectively. However, the lateral displacement of the roof in the dynamic analysis is 4.78, 6.74, and 6.79 times the static one for the Loma-Prieta, Northridge, and El-Centro earthquakes, respectively.
The large lateral displacement of the structure in dynamic cases is due to a large lateral displacement at the top of the diaphragm wall in the dynamic analysis. The results emphasize the seismic soil–structure interaction between the soil, shoring system, and adjacent structures.

7. Conclusions

In this study, 2D numerical non-linear time history analysis using PLAXIS 2D is performed to investigate the seismic performance of deep braced excavation in sandy soil. This study focuses on the seismic soil–structure–excavation interaction to assess the seismic susceptibility of the excavation support system and the adjacent structure. The model is validated against a real case study of deep braced excavation in silty–sandy soil. The parametric study includes the following variables: foundation level of the structure adjacent to excavation and the response of the shoring system and adjacent structure under the effect of the acceleration time histories of Loma-Prieta, Northridge, and El-Centro. Based on the analysis and discussion of this research, the following conclusions are drawn:
(a) 
Seismic performance of the entire system:
  • As the building’s foundation level adjacent to the excavation increases, the permanent ground surface displacement decreased significantly, while the displacement at the top of the diaphragm wall and top of the structure increased significantly.
  • The Loma-Prieta earthquake had a minimal effect on the soil at the ground surface, while the Northridge earthquake had a greater effect on the ground surface. However, the El-Centro and Northridge earthquakes had a great effect on the diaphragm wall and the adjacent structure.
  • Settlements under dynamic analysis were 5.10, 7.0, and 7.50 times the maximum static ones under the Loma-Prieta, Northridge, and El-Centro records, respectively.
(b) 
Seismic performance of the shoring system and its supporting elements:
  • The dynamic moment increased by an average of 3.32 times the static one, while the average increase in shear force under dynamic conditions was 2.85 times that under static conditions.
  • The average increases in the axial force of the two strut levels because of dynamic forces were 1.38 and 3.17 for the first and second struts, respectively, compared to static ones.
  • The maximum bending moments under the Loma-Prieta, Northridge, and El-Centro records were 1.90, 3.04, and 2.93 times the static ones, while the maximum wall shear forces under the Loma-Prieta, Northridge, and El-Centro records were 2.34, 3.24, and 3.56 times the static ones, respectively.
  • The axial forces in the first level of struts were 1.52, 1.94, and 1.64 times the static ones under the Loma-Prieta, Northridge, and El-Centro earthquakes, respectively. However, the axial force in the second level of struts was 4.39, 7.04, and 6.55 times the static one, respectively.
(c) 
Seismic performance of the structure adjacent to the shoring system:
  • The dynamic lateral displacement at the top roof was 2.30, 6.18, and 8.37 times the static ones as the foundation level increased from 1.00 m to 5.00 m depth, respectively.
  • The lateral displacements of the base of the structure were 3.87, 1.75, and 4.12 times the static ones under the Loma-Prieta, Northridge, and El-Centro earthquakes, respectively. On the other hand, the lateral displacements at the top roof in the dynamic analysis were 4.78, 6.74, and 6.79 times the static ones, respectively.

Author Contributions

T.N.S.: Writing—review and editing, Writing—original draft, and Conceptualization. M.S.E.: Writing—review and editing, Data curation, Conceptualization, and Writing—original draft. D.K.: Writing—review and editing, Visualization, Validation, Supervision, Data curation, and Conceptualization. E.D.: Writing—review and editing and Conceptualization. A.A.E.E.: Writing—review and editing, Writing—original draft, Methodology, Investigation, Formal analysis, Data curation, and Conceptualization. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding. We would like to thank the Scientific Grant Agency VEGA 1/0626/22—Ministry of Education, Science, Research and Youth and the Slovak Academy of Sciences of the Slovak Republic for APC and the opportunity to publish this contribution in this journal.

Data Availability Statement

The original contributions presented in the study are included in the article, further inquiries can be directed to the corresponding author.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Moormann, C. Analysis of wall and ground movements due to deep excavations in soft soil based on a new worldwide database. Soils Found. 2004, 44, 87–98. [Google Scholar] [CrossRef]
  2. Long, M. Database for retaining wall and ground movements due to deep excavations. J. Geotech. Geoenviron. Eng. 2001, 127, 203–224. [Google Scholar] [CrossRef]
  3. El Sawwaf, M.; Nazir, A.K. The effect of deep excavation-induced lateral soil movements on the behavior of strip footing supported on reinforced sand. J. Adv. Res. 2012, 3, 337–344. [Google Scholar] [CrossRef]
  4. Lam, S.Y.; Haigh, S.K.; Bolton, M.D. Understanding ground deformation mechanisms for multi-propped excavation in soft clay. Soils Found. 2014, 54, 296–312. [Google Scholar] [CrossRef]
  5. Ou, C.-Y.; Hsieh, P.-G.; Chiou, D.-C. Characteristics of ground surface settlement during excavation. Can. Geotech. J. 1993, 30, 758–767. [Google Scholar] [CrossRef]
  6. Huynh, Q.T.; Lai, V.Q.; Boonyatee, T.; Keawsawasvong, S. Behavior of a Deep Excavation and Damages on Adjacent Buildings: A Case Study in Vietnam. Transp. Infrastruct. Geotechnol. 2021, 8, 361–389. [Google Scholar] [CrossRef]
  7. Boscardin, M.D.; Cording, E.J. Building response to excavation-induced settlement. J. Geotech. Eng. 1989, 115, 1–21. [Google Scholar] [CrossRef]
  8. Hsieh, P.-G.; Ou, C.-Y. Shape of ground surface settlement profiles caused by excavation. Can. Geotech. J. 1998, 35, 1004–1017. [Google Scholar] [CrossRef]
  9. Liu, G.B.; Ng, C.W.; Wang, Z.W. Observed performance of a deep multistrutted excavation in Shanghai soft clays. J. Geotech. Geoenviron. Eng. 2005, 131, 1004–1013. [Google Scholar] [CrossRef]
  10. Kung, G.T.; Juang, C.H.; Hsiao, E.C.; Hashash, Y.M. Simplified Model for Wall Deflection and Ground-Surface Settlement Caused by Braced Excavation in Clays. J. Geotech. Geoenviron. Eng. 2007, 133, 731–747. [Google Scholar] [CrossRef]
  11. Wang, J.H.; Xu, Z.H.; Wang, W.D. Wall and Ground Movements due to Deep Excavations in Shanghai Soft Soils. J. Geotech. Geoenviron. Eng. 2010, 136, 985–994. [Google Scholar] [CrossRef]
  12. Bahrami, M.; Khodakarami, M.I.; Haddad, A. 3D numerical investigation of the effect of wall penetration depth on excavations behavior in sand. Comput. Geotech. 2018, 98, 82–92. [Google Scholar] [CrossRef]
  13. Ou, C.Y.; Hsieh, P.G.; Lin, Y.L. A parametric study of wall deflections in deep excavations with the installation of cross walls. Comput. Geotech. 2013, 50, 55–65. [Google Scholar] [CrossRef]
  14. Chowdhury, S.S.; Deb, K.; Sengupta, A. Estimation of Design Parameters for Braced Excavation in Clays. Geotech. Geol. Eng. 2017, 35, 857–870. [Google Scholar] [CrossRef]
  15. Chowdhury, S.S.; Deb, K.; Sengupta, A. Estimation of Design Parameters for Braced Excavation: Numerical Study. Int. J. Geomech. 2013, 13, 234–247. [Google Scholar] [CrossRef]
  16. Ou, C.Y.; Hsieh, P.G. A simplified method for predicting ground settlement profiles induced by excavation in soft clay. Comput. Geotech. 2011, 38, 987–997. [Google Scholar] [CrossRef]
  17. Hsiung, B.C.B. A case study on the behaviour of a deep excavation in sand. Comput. Geotech. 2009, 36, 665–675. [Google Scholar] [CrossRef]
  18. Yoo, C.; Lee, D. Deep excavation-induced ground surface movement characteristics—A numerical investigation. Comput. Geotech. 2008, 35, 231–252. [Google Scholar] [CrossRef]
  19. Arai, Y.; Kusakabe, O.; Murata, O.; Konishi, S. A numerical study on ground displacement and stress during and after the installation of deep circular diaphragm walls and soil excavation. Comput. Geotech. 2008, 35, 791–807. [Google Scholar] [CrossRef]
  20. Orazalin, Z.Y.; Whittle, A.J.; Olsen, M.B. Three-Dimensional Analyses of Excavation Support System for the Stata Center Basement on the MIT Campus. J. Geotech. Geoenviron. Eng. 2015, 141, 1–14. [Google Scholar] [CrossRef]
  21. Faheem, H.; Cai, F.; Ugai, K. Three-dimensional base stability of rectangular excavations in soft soils using FEM. Comput. Geotech. 2004, 31, 67–74. [Google Scholar] [CrossRef]
  22. Bose, S.K.; Som, N.N. Parametric study of a braced cut by finite element method. Comput. Geotech. 1998, 22, 91–107. [Google Scholar] [CrossRef]
  23. Ou, C.-Y.; Chiou, D.-C.; Wu, T.-S. Three-dimensional finite element analysis of deep excavations. J. Geotech. Eng. 1996, 122, 337–345. [Google Scholar] [CrossRef]
  24. Zdravkovic, L.; Potts, D.M.; John, H.D.S. Modelling of a 3D excavation in finite element analysis. In Stiff Sedimentary Clays: Genesis and Engineering Behaviour: Géotechnique Symposium in Print 2007; Emerald Publishing Limited: Leeds, UK, 2011; pp. 319–335. [Google Scholar] [CrossRef]
  25. Chowdhury, S.S.; Deb, K.; Sengupta, A. Effect of Fines on Behavior of Braced Excavation in Sand: Experimental and Numerical Study. Int. J. Geomech. 2016, 16, 04015018. [Google Scholar] [CrossRef]
  26. Callisto, L.; Soccodato, F.M.; Conti, R. Analysis of the Seismic Behaviour of Propped Retaining Structures. In Geotechnical Earthquake Engineering and Soil Dynamics IV; American Society of Civil Engineers: Reston, VA, USA, 2018; pp. 1–10. [Google Scholar] [CrossRef]
  27. Callisto, L.; Soccodato, F.M. Seismic design of flexible cantilevered retaining walls. J. Geotech. Geoenviron. Eng. 2010, 136, 344–354. [Google Scholar] [CrossRef]
  28. Konai, S.; Sengupta, A.; Deb, K. Behavior of braced excavation in sand under a seismic condition: Experimental and numerical studies. Earthq. Eng. Eng. Vib. 2018, 17, 311–324. [Google Scholar] [CrossRef]
  29. Ling, H.I.; Leshchinsky, D.; Wang, J.-P.; Mohri, Y.; Rosen, A. Seismic Response of Geocell Retaining Walls: Experimental Studies. J. Geotech. Geoenviron. Eng. 2009, 135, 515–524. [Google Scholar] [CrossRef]
  30. Ling, H.I.; Mohri, Y.; Leshchinsky, D.; Burke, C.; Matsushima, K.; Liu, H. Large-Scale Shaking Table Tests on Modular-Block Reinforced Soil Retaining Walls. J. Geotech. Geoenviron. Eng. 2005, 131, 465–476. [Google Scholar] [CrossRef]
  31. Ling, H.I.; Liu, H.; Mohri, Y. Parametric Studies on the Behavior of Reinforced Soil Retaining Walls under Earthquake Loading. J. Eng. Mech. 2005, 131, 1056–1065. [Google Scholar] [CrossRef]
  32. Mikola, R.G.; Sitar, N. Seismic Earth Pressures on Retaining Structures in Cohesionless Soils By Roozbeh Geraili Mikola and Nicholas Sitar Report Submitted to the California Department of Transportation (Caltrans) under Contract No. 65A0367 and NSF-NEES-CR Grant No. CMMI-093; California Department of Transportation, Division of Research and Innovation: Sacramento, CA, USA, 2013; p. 170. [Google Scholar] [CrossRef]
  33. Tufenkjian, M.R.; Vucetic, M. Dynamic Failure Mechanism of Soil-Nailed Excavation Models in Centrifuge. J. Geotech. Geoenviron. Eng. 2000, 126, 227–235. [Google Scholar] [CrossRef]
  34. Madabhushi, S.P.G.; Zeng, X. Simulating Seismic Response of Cantilever Retaining Walls. J. Geotech. Geoenviron. Eng. 2007, 133, 539–549. [Google Scholar] [CrossRef]
  35. Conti, R.; Madabhushi, G.S.P.; Viggiani, G.M.B. On the behaviour of flexible retaining walls under seismic actions. Geotechnique 2012, 62, 1081–1094. [Google Scholar] [CrossRef]
  36. Bahrami, M.; Khodakarami, M.I.; Haddad, A. Seismic behavior and design of strutted diaphragm walls in sand. Comput. Geotech. 2019, 108, 75–87. [Google Scholar] [CrossRef]
  37. 318 ACI Committee. Building Code Requirements for Structural Concrete (ACI 318-08) and Commentary; American Concrete Institute: Farmington Hills, MI, USA, 2008. [Google Scholar]
  38. AISC. Steel Construction Manual 15th Edition; AISC: Chicago, IL, USA, 2017; Volume 53, pp. 21–25. [Google Scholar]
  39. Castaldo, P.; De Iuliis, M. Effects of deep excavation on seismic vulnerability of existing reinforced concrete framed structures. Soil Dyn. Earthq. Eng. 2014, 64, 102–112. [Google Scholar] [CrossRef]
  40. Khademalrasoul, A.; Shirmohammadi, A.; Pakbaz, M.S.; Labibzadeh, M. Seismic Soft Soil-Structure Interaction Using RLBRB-Equipped Strut in Deep Excavations. J. Earthq. Eng. 2022, 26, 3572–3600. [Google Scholar] [CrossRef]
  41. Ibrahim, K.M.H.I.; Ibrahim, T.E. Effect of historical earthquakes on pre-stressed anchor tie back diaphragm wall and on near-by building. HBRC J. 2013, 9, 60–67. [Google Scholar] [CrossRef]
  42. Mo, Y.L.; Hwang, J.M. Seismic response of reinforced concrete buildings close to deep excavation. Comput. Methods Appl. Mech. Eng. 1997, 145, 117–134. [Google Scholar] [CrossRef]
  43. Chowdhury, S.S.; Deb, K.; Sengupta, A. Behavior of underground strutted retaining structure under seismic condition. Earthq. Struct. 2015, 8, 1147–1170. [Google Scholar] [CrossRef]
  44. Yeganeh, N.; Bazaz, J.B.; Akhtarpour, A. Seismic analysis of the soil-structure interaction for a high rise building adjacent to deep excavation. Soil Dyn. Earthq. Eng. 2015, 79, 149–170. [Google Scholar] [CrossRef]
  45. Hsiung, B.C.B.; Dao, S.D. Evaluation of constitutive soil models for predicting movements caused by a deep excavation in sands. Electron. J. Geotech. Eng. 2014, 19, 17325–17344. [Google Scholar]
  46. Benz, T. Small-Strain Stiffness of Soils and its Numerical Consequences; University of Stuttgart—Institute of Geotechnical Engineering: Stuttgart, Germany, 2007. [Google Scholar]
  47. Hardin, B.O.; Drnevich, V.P. Shear Modulus and Damping in Soils: Design Equations and Curves. J. Soil Mech. Found. Div. 1972, 98, 667–692. [Google Scholar] [CrossRef]
  48. Santos, J.; Correia, A. Reference threshold shear strain of soil. Its application to obtain an unique strain-dependent shear modulus curve for soil. In Proceedings of the XV International Conference on Soil Mechanics and Geotechnical Engineering, Istanbul, Turkey, 27–31 August 2001; pp. 267–270. [Google Scholar]
  49. Obrzud, R. The Hardening Soil Model: A Practical Guidebook; Zace Services: Lausanne, Switzerland, 2010. [Google Scholar]
  50. Brinkgreve, R.B.J.; Bakker, K.J.; Bonnier, P.G. The relevance of small-strain soil stiffness in numerical simulation of excavation and tunnelling projects. In Numerical Methods in Geotechnical Engineering; Taylor & Francis Group: London, UK, 2006; pp. 133–139. [Google Scholar] [CrossRef]
  51. Zienkiewicz, O.C.; Taylor, R.L. The Finite Element Method. Vol. 2: Solid and Fluid Mechanics, Dynamics and Non-Linearity, 4th ed.; McGraw-Hill: Columbus, OH, USA, 1991. [Google Scholar]
  52. Chopra, A.K. Dynamics of Structures 5th Edition SI Units Anul K Chopra; Pearson: London, UK, 2020; Volume 5. [Google Scholar]
  53. Hughes, T.J.R. The Finite Element Method: Linear Static and Dynamic Finite Element Analysis; Courier Corporation: Mineola, NY, USA, 2003. [Google Scholar]
  54. Kramer, S.L. Geotechnical Eartquake Engineering, 2nd ed.; Prentice Hall: Upper Saddle River, NJ, USA, 1996. [Google Scholar]
  55. Hudson, M.; Idriss, I.M.; Beirkae, M. Quad4m User’s Manual; University of California: Davis, CA, USA, 1994. [Google Scholar]
  56. Kuhlemeyer, R.L.; Lysmer, J. Finite element method accuracy for wave propagation problems. J. Soil Mech. Found. Div. 1973, 99, 421–427. [Google Scholar] [CrossRef]
  57. Lysmer, J.; Udaka, T.; Tsai, C.; Seed, H.B. FLUSH-A Computer Program for Approximate 3-D Analysis of Soil-Structure Interaction Problems; California Univ., Richmond (USA). Earthquake Engineering Research Center: Berkeley, CA, USA, 1975. [Google Scholar]
  58. Zienkiewicz, O.C.; Bicanic, N.; Shen, F.Q. Earthquake Input Definition and the Trasmitting Boundary Conditions. In Advances in Computational Nonlinear Mechanics; Springer: Vienna, Austria, 1989; pp. 109–138. [Google Scholar] [CrossRef]
  59. Joyner, W.B.; Chen, A.T.F. Calculation of nonlinear ground response in earthquakes. Bull. Seismol. Soc. Am. 1975, 65, 1315–1336. [Google Scholar]
  60. PLAXIS 2D Reference Manual. In PLAXIS 2D 2024.2 Reference Manual 2D; Bentley Systems: Exton, PA, USA, 2024.
  61. Ou, C.-Y. Deep Excavation: Theory and Practice; CRC Press: Boca Raton, FL, USA, 2014. [Google Scholar]
  62. Khoiri, M.; Ou, C.Y. Evaluation of deformation parameter for deep excavation in sand through case histories. Comput. Geotech. 2013, 47, 57–67. [Google Scholar] [CrossRef]
  63. Likitlersuang, S.; Surarak, C.; Wanatowski, D.; Oh, E.; Balasubramaniam, A. Finite element analysis of a deep excavation: A case study from the Bangkok MRT. Soils Found. 2013, 53, 756–773. [Google Scholar] [CrossRef]
  64. ASCE Standard ASCE/SEI 7-16; Minimum Design Loads and Associated Criteria for Buildings and Other Structures. American Society of Civil Engineers: Reston, VA, USA, 2017. [CrossRef]
  65. Hokmabadi, A.S.; Fatahi, B.; Samali, B. Recording inter-storey drifts of structures in time-history approach for seismic design of building frames. Aust. J. Struct. Eng. 2012, 13, 175–179. [Google Scholar] [CrossRef]
  66. Ebadi-Jamkhaneh, M. Pounding Risk Assessment through Soil–Structure Interaction Analysis in Adjacent High-Rise RC Structures. Buildings 2024, 14, 2779. [Google Scholar] [CrossRef]
  67. Katsimpini, P.S. Seismic Response of Vertical Hybrid Concrete/Steel Frames Considering Soil–Structure Interaction. Buildings 2024, 14, 972. [Google Scholar] [CrossRef]
  68. Ekraminia, S.; Bonab, M.H.; Ghassemi, S.; Derakhshani, R. Experimental study on seismic response of underground tunnel–soil–piled structure interaction using shaking table in loose sand. Buildings 2023, 13, 2482. [Google Scholar] [CrossRef]
  69. Bian, X.; Wang, G. Study of structural seismic damage considering seasonal frozen soil–structure interaction. Buildings 2024, 14, 1493. [Google Scholar] [CrossRef]
  70. Alnmr, A.; Sheble, A.; Ray, R.; Ahmad, H. Parametric investigation of interaction between soil-surface structure and twin tunnel excavation: A comprehensive 2D numerical study. Infrastructures 2023, 8, 124. [Google Scholar] [CrossRef]
  71. PLAXIS 2D Scientific Manual. In PLAXIS PLAXIS 2D 2024.1 Scientific Manual 2D; Bentley Systems: Exton, PA, USA, 2023.
  72. PEER. PEER Ground Motion Database—PEER Center. NGA-West2. [Online]. Available online: https://ngawest2.berkeley.edu/ (accessed on 18 August 2024).
  73. Rankin, W.J. Ground movements resulting from urban tunnelling: Predictions and effects. Geol. Soc. Lond. Eng. Geol. Spec. Publ. 1988, 5, 79–92. [Google Scholar] [CrossRef]
  74. Gazetas, G. Empirical seismic design method for waterfront anchored sheetpile walls. In Design and Performance of Earth Retaining Structures; The International Society for Soil Mechanics and Geotechnical Engineering (ISSMGE): Christchurch, New Zealand, 1990. [Google Scholar]
Figure 1. Satellite map indicating the position of the case study, i.e., Case A.
Figure 1. Satellite map indicating the position of the case study, i.e., Case A.
Buildings 15 03732 g001
Figure 2. (a) Verification model geometry, soil stratigraphy, and (b) mesh discretization with boundary conditions.
Figure 2. (a) Verification model geometry, soil stratigraphy, and (b) mesh discretization with boundary conditions.
Buildings 15 03732 g002
Figure 3. Comparison of the measured and predicted lateral displacement of the diaphragm wall at the final stage of excavation.
Figure 3. Comparison of the measured and predicted lateral displacement of the diaphragm wall at the final stage of excavation.
Buildings 15 03732 g003
Figure 4. Comparison of the measured and the predicted ground surface settlement for all stages of excavation.
Figure 4. Comparison of the measured and the predicted ground surface settlement for all stages of excavation.
Buildings 15 03732 g004
Figure 5. Model geometry, mesh discretization, and boundary conditions of the case study.
Figure 5. Model geometry, mesh discretization, and boundary conditions of the case study.
Buildings 15 03732 g005
Figure 6. Acceleration record of the Loma-Prieta (1989) earthquake of Mw (6.90).
Figure 6. Acceleration record of the Loma-Prieta (1989) earthquake of Mw (6.90).
Buildings 15 03732 g006
Figure 7. Acceleration record of the Northridge (1994) earthquake of Mw (6.70).
Figure 7. Acceleration record of the Northridge (1994) earthquake of Mw (6.70).
Buildings 15 03732 g007
Figure 8. Acceleration record of the El-Cento (1940) earthquake of Mw (6.90).
Figure 8. Acceleration record of the El-Cento (1940) earthquake of Mw (6.90).
Buildings 15 03732 g008
Figure 9. Displacement–time history of different structures’ foundation levels ((ac): free-field ground surface, top of diaphragm wall, and top roof of structure, respectively).
Figure 9. Displacement–time history of different structures’ foundation levels ((ac): free-field ground surface, top of diaphragm wall, and top roof of structure, respectively).
Buildings 15 03732 g009
Figure 10. Settlement trough for different foundation levels of the structure ((a,b): static analysis and dynamic analysis, respectively).
Figure 10. Settlement trough for different foundation levels of the structure ((a,b): static analysis and dynamic analysis, respectively).
Buildings 15 03732 g010
Figure 11. Lateral displacement of the diaphragm wall for different foundation levels of the adjacent structure ((a,b): static lateral displacement and dynamic lateral displacement, respectively).
Figure 11. Lateral displacement of the diaphragm wall for different foundation levels of the adjacent structure ((a,b): static lateral displacement and dynamic lateral displacement, respectively).
Buildings 15 03732 g011
Figure 12. Comparison of the lateral displacement at the top of the diaphragm wall in static and dynamic cases for different foundation levels of the adjacent structure.
Figure 12. Comparison of the lateral displacement at the top of the diaphragm wall in static and dynamic cases for different foundation levels of the adjacent structure.
Buildings 15 03732 g012
Figure 13. Diaphragm wall straining actions in static and dynamic cases for different foundation levels of the structure ((ad): static bending moment, dynamic bending moment, static shear force, and dynamic shear force, respectively).
Figure 13. Diaphragm wall straining actions in static and dynamic cases for different foundation levels of the structure ((ad): static bending moment, dynamic bending moment, static shear force, and dynamic shear force, respectively).
Buildings 15 03732 g013aBuildings 15 03732 g013b
Figure 14. Comparison of dynamic and static maximum straining actions along the diaphragm wall for different foundation levels of the structure.
Figure 14. Comparison of dynamic and static maximum straining actions along the diaphragm wall for different foundation levels of the structure.
Buildings 15 03732 g014
Figure 15. Axial force of the struts in both static and dynamic cases for different foundation levels of the structure.
Figure 15. Axial force of the struts in both static and dynamic cases for different foundation levels of the structure.
Buildings 15 03732 g015
Figure 16. Comparison of the maximum axial forces of struts in dynamic and static cases for different foundation levels of the structure.
Figure 16. Comparison of the maximum axial forces of struts in dynamic and static cases for different foundation levels of the structure.
Buildings 15 03732 g016
Figure 17. The structure’s story displacement in static and dynamic cases for different foundation levels ((a,b): static case and dynamic case, respectively).
Figure 17. The structure’s story displacement in static and dynamic cases for different foundation levels ((a,b): static case and dynamic case, respectively).
Buildings 15 03732 g017
Figure 18. Comparison of dynamic and static maximum story displacement at the top roof and at the base of the structure for different structures’ foundation levels.
Figure 18. Comparison of dynamic and static maximum story displacement at the top roof and at the base of the structure for different structures’ foundation levels.
Buildings 15 03732 g018
Figure 19. Displacement–time history of base ground motion, ground surface free-field motion, the top point of the diaphragm wall, and the top of the structure adjacent to the excavation ((ac): using the Loma-Prieta, Northridge, and El-Centro earthquakes, respectively).
Figure 19. Displacement–time history of base ground motion, ground surface free-field motion, the top point of the diaphragm wall, and the top of the structure adjacent to the excavation ((ac): using the Loma-Prieta, Northridge, and El-Centro earthquakes, respectively).
Buildings 15 03732 g019aBuildings 15 03732 g019b
Figure 20. Lateral displacement of the shoring system under static conditions and under Loma-Prieta, Northridge, and El-Centro earthquakes.
Figure 20. Lateral displacement of the shoring system under static conditions and under Loma-Prieta, Northridge, and El-Centro earthquakes.
Buildings 15 03732 g020
Figure 21. Dynamic/static lateral displacement ratio for different earthquake records.
Figure 21. Dynamic/static lateral displacement ratio for different earthquake records.
Buildings 15 03732 g021
Figure 22. Settlement trough behind the diaphragm wall in static and dynamic cases.
Figure 22. Settlement trough behind the diaphragm wall in static and dynamic cases.
Buildings 15 03732 g022
Figure 23. Straining actions along the diaphragm wall under static conditions and for different earthquake records ((a,b): bending moment diagram and shear force diagram, respectively).
Figure 23. Straining actions along the diaphragm wall under static conditions and for different earthquake records ((a,b): bending moment diagram and shear force diagram, respectively).
Buildings 15 03732 g023
Figure 24. Comparison of the dynamic straining actions and the static ones for different earthquake records.
Figure 24. Comparison of the dynamic straining actions and the static ones for different earthquake records.
Buildings 15 03732 g024
Figure 25. Axial force of the struts under static conditions and under different earthquake ground motions.
Figure 25. Axial force of the struts under static conditions and under different earthquake ground motions.
Buildings 15 03732 g025
Figure 26. Comparison of the axial forces of the struts under different earthquake ground motions.
Figure 26. Comparison of the axial forces of the struts under different earthquake ground motions.
Buildings 15 03732 g026
Figure 27. The story’s lateral displacement distribution in static and dynamic cases.
Figure 27. The story’s lateral displacement distribution in static and dynamic cases.
Buildings 15 03732 g027
Figure 28. Comparison of the maximum story displacement at the base and top of the structure in static and dynamic cases.
Figure 28. Comparison of the maximum story displacement at the base and top of the structure in static and dynamic cases.
Buildings 15 03732 g028
Table 1. Soil properties used in the parametric study of the dense sand layer using the HS-small model.
Table 1. Soil properties used in the parametric study of the dense sand layer using the HS-small model.
Depth of Layer (m)60
Soil typeDense sand
SPT (N-Value)37
Soil modelHardening soil model with small strain (HS-small)
Unsaturated unit weight, γunsat (kN/m3)19
Saturated unit weight, γsat (kN/m3)21
Rayleigh damping (α)0.1634
Rayleigh damping (β)1.662 × 10−3
Damping ratio 2%
Internal friction angle, ( ° ) 36
Cohesion, C ( K P a ) 1
Dilatancy angle, Ψ ( ° ) 6
Secant stiffness from drained triaxial test, E 50 r e f ( K P a ) 50,000
Reference tangent stiffness for oedometer primary loading, E o e d r e f ( K P a ) 50,000
Reference unloading/reloading stiffness, E u r r e f ( K P a ) 150,000
Unloading/reloading Poisson’s
ratio, νur
0.20
Exponent of stress-level dependency of stiffness, m0.50
Failure ratio, Rf0.90
At-rest earth pressure coefficient, K00.4122
Reference small-strain shear modulus, G0ref (kPa)180,000
Shear strain magnitude at 0.722 G0, γ0.71 × 10−4
Analysis typeDrained
Rinter0.67
Table 2. Summary of the studied parameters.
Table 2. Summary of the studied parameters.
Studied ParameterParameter VariablesChosen Parameters
Foundation level of the structure adjacent to the shoring systemDf = 1.0 mH = 10 m
B = 20 m, Tw = 80 cm
D = H = 10 m
Loma-Prieta EQ
Df = 3.0 m
Df = 5.0 m
Earthquake recordsLoma-Prieta (1989), Mw = 6.90H = D = 10 m
Tw = 80 cm
B = 20 m
Northridge (1994), Mw = 6.70
El-Centro (1940), Mw = 6.90
H = Final excavation depth. B = Width of excavation. Tw = Diaphragm wall thickness. D = Diaphragm wall embedment depth. Df = Foundation level of the adjacent structure.
Table 3. Properties of input earthquake ground motions.
Table 3. Properties of input earthquake ground motions.
Earthquake RecordMoment Magnitude
(Mw)
PGA
(g)
Duration
(s)
Predominant Period, Tp
(s)
Arias Intensity, Ia (m/s)Significant Duration, Ds 5–95 (s)
Loma-Prieta (1989)6.900.3739.900.221.3511.37
Northridge (1994)6.700.8860.000.222.728.75
El-Centro (1940)6.900.3453.760.561.7624.46
Table 4. Amplification ratio and permanent displacements at the end of earthquakes.
Table 4. Amplification ratio and permanent displacements at the end of earthquakes.
Earthquake RecordLoma-Prieta,
Mw = 6.9
Northridge,
Mw = 6.7
El-Centro,
Mw = 6.9
Ground surfaceA,max (%)210170197
Ux,per (mm)18.37112.3133.71
Top of the diaphragm wallA,max (%)294228223
Ux,per (mm)−60.73−47.95−93.8
Top of the structureA,max (%)297558289
Ux,per (mm)−148.39−215.48−230.45
Amax: Maximum amplification ratio. Ux,per: Permanent displacement at the end of the earthquake.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Salem, T.N.; Elmahdy, M.S.; Katunský, D.; Dolníková, E.; El Ela, A.A. Numerical Investigation of Seismic Soil–Structure–Excavation Interaction in Sand. Buildings 2025, 15, 3732. https://doi.org/10.3390/buildings15203732

AMA Style

Salem TN, Elmahdy MS, Katunský D, Dolníková E, El Ela AA. Numerical Investigation of Seismic Soil–Structure–Excavation Interaction in Sand. Buildings. 2025; 15(20):3732. https://doi.org/10.3390/buildings15203732

Chicago/Turabian Style

Salem, Tarek N., Mahmoud S. Elmahdy, Dušan Katunský, Erika Dolníková, and Ahmed Abu El Ela. 2025. "Numerical Investigation of Seismic Soil–Structure–Excavation Interaction in Sand" Buildings 15, no. 20: 3732. https://doi.org/10.3390/buildings15203732

APA Style

Salem, T. N., Elmahdy, M. S., Katunský, D., Dolníková, E., & El Ela, A. A. (2025). Numerical Investigation of Seismic Soil–Structure–Excavation Interaction in Sand. Buildings, 15(20), 3732. https://doi.org/10.3390/buildings15203732

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