Home Numerical modeling of site response at large strains with simplified nonlinear models: Application to Lotung seismic array
Article Open Access

Numerical modeling of site response at large strains with simplified nonlinear models: Application to Lotung seismic array

  • Francesco Di Buccio ORCID logo EMAIL logo and Alessandro Pagliaroli ORCID logo
Published/Copyright: September 18, 2024
Become an author with De Gruyter Brill

Abstract

Site response analyses at large strains are routinely carried out neglecting the shear strength of soil and the stiffness degradation due to the increase in pore pressures, leading to unrealistic predictions of the seismic response of soil deposits. The study investigates the performance of a simplified nonlinear (NL) approach, implemented in the Deepsoil code, constituted by coupling a hyperbolic model incorporating shear strength with a strain-based semi-empirical pore pressure generation model. The first part of the study, based on a large one-dimensional parametric study, shows that above a shear strain of 0.1%, it is necessary to include shear strength in the site response modelling to get more realistic results. Then, the approach has been evaluated with reference to the well-known downhole Large-Scale Seismic Test array located in Lotung (Taiwan): numerical results have been compared with recordings in terms of acceleration response spectra and pore water pressure time histories at different depths along the soil profiles. The comparison shows that the NL simplified model is characterized by an accuracy comparable with more sophisticated advanced elasto-plastic NL analyses adopting essentially the same input data of the traditional equivalent linear approaches(shear modulus and damping curves) and simple physical-mechanical properties routinely determined during geotechnical surveys (i.e., shear strength, relative density, fine content). This approach is therefore recommended for site response analyses reaching large strains (i.e., soft soil deposits and moderate-to-high input motions).

1 Introduction

The ground motion amplification phenomena due to local geological and geotechnical site conditions have a relevant impact on the response of structures and environment during earthquake events. The evaluation of site response is therefore a crucial step in the assessment of seismic risk both at urban scale, in the framework of advanced seismic microzonation studies supporting urban planning, and at building/infrastructure scale to quantify design seismic actions [1] also including soil-foundation-structure interaction phenomena [2,3,4]. The behavior of soil under cyclic loading is highly NL and depends on several geometric and geotechnical factors as well as from the characteristics of the input motion; for such a complex task, the numerical modelling of seismic response is the procedure generally adopted for the assessment of site response. Computer codes for site response analyses differ mainly in geometry dimensions, domain of analysis, numerical scheme, and soil constitutive relationships [5]. In particular, two approaches of analysis to model soil nonlinearity are available in research and practice: equivalent linear (in which a series of visco-elastic linear analyses are performed) and a true NL analysis based on different constitutive NL models spanning from the simplified hyperbolic to advanced elasto-plastic formulations [6].

Advanced elasto-plastic NL models can adequately capture the behavior at large strains including the development of plastic deformations, the shear strength of soils, and pore pressure build-up ([7,8,9,10] among others). Key elements of these models are yield surfaces, flow rules, and hardening/softening laws; the interaction between phases is modelled through the theory of consolidation generalized to the dynamic field [11,12]. However, the large number of constitutive parameters and the complex calibration processes, sometimes based on non-conventional laboratory apparatuses, limit in practice the use of these comprehensive models. More simplified NL models like those belonging to the family of hyperbolic soil models are therefore often used for NL simulations [13,14,15]. The hyperbolic models require the definition of the skeleton curve and criteria for loading-unloading-reloading behavior; regarding the pore water pressure (PWP) variation during seismic loading, semi-empirical PWP models (i.e., power water pressure generation relationships) can be employed in combination with hyperbolic models working in total stress.

The parameters of hyperbolic models are generally determined by fitting reference curves of normalized shear modulus and damping values as functions of shear strain. In this way, the cyclic soil behavior is adequately captured at small-to-medium strains while unrealistic shear stresses (i.e., non-compatible with the shear strength of soils) can develop at large strains leading to unreasonable prediction of seismic response [16,17,18].

The study deals with the importance of including shear strength and PWP build-up in the NL numerical modelling of site response at large strains. These aspects are traditionally neglected in the NL site response analyses because of the abovementioned limitations associated with the advanced elasto-plastic NL models. The study therefore focuses on the use of simplified constitutive models requiring few parameters and with clear physical significance, determined by standard surveys. In particular, the hyperbolic models (namely, Modified Kondner Zelasko; MKZ and general quadratic/hyperbolic; GQ/H models) coupled with strain-based pore pressure generation model implemented in the 1D NL code Deepsoil are employed [13]. In the first part, based on the results of an extensive parametric study, a shear strain threshold above which the shear strength should be included as an input parameter in numerical site response analyses is proposed. Then, the performance of the hyperbolic models in predicting ground motion at large strains, including PWP development, is tested in one of the most studied sites in the literature: the downhole Large-Scale Seismic Test (LSST) array located in Lotung (Taiwan). Seismic events characterized by increasing peak ground acceleration (PGA) have been considered, to investigate soil behavior from small to high strains. The results of the NL analyses have been compared to recordings in terms of response spectra and PWP build-up at different locations along soil profiles. Differences between the results here obtained with simplified models and predictions from advanced constitutive models in literature are discussed.

2 NL hyperbolic models

2.1 Deepsoil computer code

Deepsoil [13] is a one-dimensional site-response analysis program mainly used for nonlinear (NL) time-domain analysis with or without pore pressure generation. The equivalent linear frequency domain analysis is also available as implemented in other popular codes like SHAKE [19] or STRATA [20]. For NL analysis, a multi-degree-of-freedom lumped model is used in which each individual layer is represented by a corresponding mass, NL spring, and a dashpot for viscous damping. The soil response is obtained from a constitutive model that describes the cyclic behavior. To represent the initial stress–strain backbone curve, Deepsoil employs an updated version of the MKZ model proposed by Matasovic [21]; moreover, a GQ/H model [22], explicitly considering soil strength, is available for backbone curve. Deepsoil uses both standard and a modified version of the extended Masing rules [23] to simulate the unloading-reloading branches.

In order to model the soil damping at small strains where damping from hyperbolic models approaches zero values, Deepsoil provides in addition to the two-control frequencies full Rayleigh formulation [24], an extended Rayleigh formulation [25], in which the user must specify four control frequencies, thus extending the frequency range in which the damping variation is small, and a frequency independent damping formulation [13].

The main features of MKZ and GQ/H models employed in the present study are described in Sections 2.2 and 2.3, respectively.

In Deepsoil, semi-empirical PWP models are available to compute the PWP variation during seismic loading. As stated in the introduction, these models are coupled with the NL hyperbolic models working in total stress (“loosely coupled approach” according to Chiaradonna [26] as an alternative to more complex approach based on advanced constitutive effective-stress models (“fully couped approach”). A summary of available models and codes for both loosely and fully coupled methods is reported in the study by Chiaradonna [26] while main features of strain-based loosely coupled models used in the present study are summarized here.

2.2 MKZ model

The Modified Kondner Zelasko model, known as MKZ, is a constitutive model based on a modification of the hyperbolic model first suggested by Kondner [27] that uses a hyperbolic shear stress–strain relationship according to the following equation:

(1) τ = G 0 γ 1 + β γ γ r s ,

where G 0 = initial shear modulus, τ = shear stress, γ = shear strain, while β, s, and γ r are the model parameters to be determined by fitting the normalized shear modulus and damping curves. Moreover, Deepsoil extends the MKZ through the “pressure-dependent hyperbolic model” proposed by Hashash and Park [28] allowing the reference strain confining to be pressure-dependent (i.e., depending on the effective vertical stress).

The extended unload-reload Masing rules [23] are then used to model soil behavior under irregular cyclic loading:

  1. For initial loading, the stress–strain curve follows the backbone curve:

    τ = F bb ( γ )

    where F bb(γ) is the backbone curve function.

  2. If a stress reversal occurs at a point (γ rev, τ rev), the stress–strain curve follows a path given by:

    (2) τ τ rev 2 = F bb γ γ rev 2 .

  3. If the unloading or reloading curve intersects the backbone curve, it follows the backbone curve until the next stress reversal.

  4. If an unloading or reloading curve crosses an unloading or reloading curve from the previous cycle, the stress–strain curve follows that of the previous cycle.

It is well known that Masing rules results in overestimation of damping at large strains [29]. To overcome this problem, the modulus reduction and damping curve fitting (MRDF) Pressure-Dependent Hyperbolic model introducing a reduction factor which effectively alters the Masing rules has been implemented in the code. The reduction factor F(γ m) is calculated as a function of γ m, the maximum shear strain experienced by the soil at any given time, according to two alternative formulations modulus reduction and damping curve fitting-University of Illinois at Urbana-Champaign and MRDF-Darendeli described in the study by Park and Hashash [13]. Non Masing MRDF-UIUC approach has been employed in this study. The hyperbolic unload-reload equation is modified with the reduction factor, F(γ m), as follows:

(3) τ = F ( γ m ) 2 G 0 γ γ rev 2 1 + β γ γ rev 2 γ r s G 0 ( γ γ rev ) 1 + β γ m γ r s + G 0 ( γ γ rev ) 1 + β γ m γ r s + τ rev .

The MKZ model is generally calibrated on modulus reduction and damping curve without control of maximum shear stress, i.e., the soil shear strength is not explicitly considered in the model. A four-step procedure to adjust the modulus reduction curve to fit with shear stress at large strains (i.e., shear strength) has been proposed by Hashash et al. [6] and followed in the present study:

  1. Calibrate the backbone to fit the reference stiffness and damping curves (G/G 0-γ and D-γ) with MKZ model;

  2. Compute the implied shear strength as the maximum shear stress value calculated using the following equation where G(γ)/G 0 is the modulus reduction curve obtained from backbone curve:

    (4) τ = ρ · V s 2 · G ( γ ) G 0 · γ .

  3. Compare the implied shear strength obtained from the previous expression and the real shear strength of the soil, estimated in dynamic conditions as 1.1–1.4 of the static shear strength [30].

  4. In the case of the implied shear strength greater/smaller than the real shear strength, manually change backbone curve parameters to decrease/increase, respectively, the G(γ)/G 0 curve at large strain (indicatively for strains higher than 0.1%).

The procedure is iteratively repeated until the implied shear strength match the real soil strength.

The limiting values of shear stresses in soils can be estimated assuming simple shear conditions in the soil element subjected to vertical propagating shear waves. Before earthquake, the soil element is subjected to geostatic vertical and horizontal stresses σ v and σ h = k 0 · σ v being k 0 coefficient of earth pressure at rest; in 1D condition (horizontal ground surface and soil layering), vertical and horizontal directions are principal directions of stress state and therefore, there are no shear stress on the vertical and horizontal planes. During S-waves propagation, the soil element is subjected to shear stress τ hv on the vertical and horizonal planes under constant normal stresses. Assuming a Mohr-Coulomb failure criterion, when the soil comes to failure, the maximum shear stress acting on vertical and horizontal planes (τ hv)max that cannot be exceeded during 1D site response analyses, is [31,32]:

(5) ( τ hv ) max = = 1 + k 0 2 σ v sin φ + c cos φ 2 1 k 0 2 σ v 2 ,

with c′ and φ′ cohesion and friction angle of soil in terms of effective stresses.

2.3 GQ/H model

Groholski et al. [22] developed a new general/hyperbolic model characterized by the following backbone curve:

(6) τ τ max = 2 γ γ r 1 + γ γ r + 1 + γ γ r 2 4 θ t γ γ r ,

where τ is the shear stress, τ max is the shear stress at failure, γ is the shear strain, γ r is the reference shear strain, and θ t is a curve-fitting parameter. This model allows to fit both an initial (or maximum) shear modulus G 0 at zero strain and a limiting shear stress at large shear strains (i.e., soil shear strength), assuming a quadratic model controlling the NL behavior between these two “boundary conditions.” Following Hardin and Drnevich [31], the reference stain γ r is defined as

(7) γ r = τ max G 0 .

The coefficient θ t is a model-fitting function adjusting the curvature of the stress–strain curve and does not affect the boundary conditions of the model. It is a function of the shear strain according to the following expression:

(8) θ t = θ 1 + θ 2 · θ 4 · γ γ r θ 5 θ 3 θ 5 + θ 4 · γ γ r θ 5 ,

where parameters θ 1θ 5 are chosen to provide the best fit to normalized shear modulus vs shear strain curve over the whole strain range.

As for traditional hyperbolic models like MKZ, the extended unload-reload Masing rules as well as MRDF approach, which modify the Masing rules by applying a reduction factor, can be used in conjunction with GQ/H model. An automatic fitting curve scheme is available in Deepsoil and only the shear velocity, unit weight, and the shear strength must be assigned by the user. The shear strength can be estimated as (τ hv)max according to the formula reported in the previous section as a function of soil resistance parameters.

2.4 PWP models

As stated in the introduction, the PWP variation induced by seismic loading can be predicted by semi-empirical PWP models coupled with hyperbolic models working in total stress. Pore pressure generation models can generally be categorized into stress-based, strain-based, and energy-based models [6]; the first models are based on the results of cyclic stress-controlled tests while further studies showed a better correlation between PWP build-up and the shear strain level or the energy dissipated within the soil deposit [6,26]. In Deepsoil code, the stress-based model developed by Park and Ahn [33] for sands has been implemented in the code while for strain-based relationships, the Dobry/Matasovic [21] and Matasovic/Vucetic [34] models, for sands and clays, respectively, are available. Moreover, the user can apply the energy-based model GMP [35] for sandy soils and energy-based Generalized model [35,36] for any soil.

In the present study, strain-based relationships have been employed for PWP build-up; these models are briefly described here while more details, including the formulation of stress- and energy-based models, can be found in the study by Park and Hashash [13].

The Dobry/Matasovic model has been originally developed by Dobry et al. [37] and later modified by other authors [21,38,39]. The normalized excess pore pressure (r u = Δu/ σ vo ) is generated using the following equation:

(9) r u = p f N eq F ( γ γ tvp ) s 1 + f N eq F ( γ γ tvp ) s ,

where N eq is the equivalent number of cycles; for uniform strain cycles, the equivalent number of cycles is the same as the number of loading cycles while for irregular strain cycles, N eq is calculated incremented by 0.5 at each strain reversals; γ c is the current reversal shear strain; γ tvp is the volumetric threshold: the excess PWP is different from zero only if the shear strain level is greater than the threshold shear strain value; this threshold value is generally determined by means of cyclic or dynamic laboratory tests;

f is a parameter introduced by Vucetic [40,41] in order to simulate the cyclic loading in multiple dimensions: f is equal to 1 in 1D motion and 2 for 2D motion;

p, s, F, and curve are fitting parameters and can be obtained by calibration procedures based on undrained cyclic strain-controlled laboratory tests [21]. In the absence of laboratory data, for practical purposes, p can be assumed equal to 1 while Carlton [42] proposed empirical correlations for the curve fitting parameters F and s for sands:

(10) F = 3 , 810 V s ( 1.55 ) ,

(11) s = ( FC + 1 ) 0.1252 ,

where V s is the shear wave velocity in m/s, and FC is the percentage of fines content. Mei and Olson [43] developed further correlation for parameter F using 123 cyclic shear test results compiled from literature referring to sub-angular to sub-rounded clean sands. Two soil index properties, relative density (Dr), and uniformity coefficient (Cu) are used in the correlation.

For clayey soils, Matasović [39] proposed the following equation for the normalized excess PWP:

(12) r u = A N eq 3 s ( γ c γ tvp ) r + B N eq 2 s ( γ c γ tvp ) r + C N eq s ( γ c γ tvp ) r + D ,

where N eq and γ tvp are the same parameters described in the Dobry/Matasovic model for sands, while the s, r, A, B, C, and D are fitting parameters that can be obtained from the empirical correlations proposed by Carlton [42] as a function of plasticity index (PI) and overconsolidation ratio (OCR) of the clay [13].

The degradation of shear strength and shear stiffness due to the generation of excess PWP during dynamic loading can be considered in the hyperbolic model (MKZ or GQ/H) by introducing the following two degradation indices [21]:

(13) δ G = 1 r u ,

(14) δ τ = 1 ( r u ) v ,

where δ G is the shear modulus degradation factor, δ τ is the shear stress degradation factor, r u is the normalized excess PWP, and ν is a curve-fitting parameter to better model the degradation of shear strength with excess pore pressure generation. These degradation parameter formulations are implemented for all soil models except the Matasovic and Vucetic models for clays where the degradation parameters are defined [21] as follows:

(15) δ G = δ τ = N eq 1 .

Based on Terzaghi 1D consolidation theory, the code allows to also include PWP dissipation simultaneously to generate during ground shaking.

3 Shear strain threshold for “large strains” effects

It is well known that pore pressure build-up becomes significant for seismic-induced shear strains exceeding the volumetric threshold γ tvp which is typically in the range of 0.01–0.1% [44,45]; lower values pertain to sandy soils while higher values characterize clayey soils with γ tvp increasing with plasticity index PI.

For dynamic problems, in which shear strains exceed γ tvp site response, analyses should in principle including models for pore pressure build-up. Analogously, one might ask what is the threshold for which the effect of soil resistance becomes relevant and therefore, this parameter should be included in the site response modeling. A large 1D parametric numerical study has been carried out by Di Buccio [32] to answer this question and briefly summarized here.

Sandy and clayey ideal soil deposits characterized by shear wave velocity increasing with the depth have been considered. The seismic bedrock (characterized by unit weight γ = 22 kN/m3 and V s = 800 m/s) has been placed at depths H of 20, 40, 60, and 80 m from ground surface. Two main lithologies were considered: clay with γ = 18 kN/m3 and NL behavior defined by Darendeli [46] curves for PI = 30 and OCR = 1 and sand characterized by γ = 20 kN/m3 and Seed and Idriss [47] curves. Darendeli curves are pressure-dependent being function of the confining stress here computed assuming 0.6 for the at-rest earth pressure coefficient k 0. Analogously for sands, the dependence of NL curves from confining stress has been roughly considered by assuming the mean or upper bound of the Seed and Idriss curves depending on the thickness of the soil deposit.

Different V s profiles were adopted by setting reference values of equivalent shear wave velocity V S,eq/V S,30 at 150 and 250 m/s for clayey soil deposits while equivalent values of 300 and 450 m/s were assigned to sandy soil deposits. Considering the different V s profiles, thicknesses, and lithologies, a total of 16 ideal soil deposits were defined and subjected to numerical analyses (Figure 1).

Figure 1 
               Ideal sandy and clayey soil profiles considered in the parametric analyses.
Figure 1

Ideal sandy and clayey soil profiles considered in the parametric analyses.

Natural accelerograms with increasing PGA have been considered as input motions. The signals were divided into six set of five recordings, with PGA falling in the following ranges: 1–10, 10–50, 50–100, 100–200, 200–400 cm/s2, and >400 cm/s2. The accelerograms were extracted from the Italian ITACA database (http://itaca.mi.ingv.it/) and the international PEER database (https://ngawest2.berkeley edu/), assuming magnitude and distance in the ranges of M = 4.5–6.9 and D = 5–50 km, respectively.

The analyses were performed with the 1D code Deepsoil version 7.0.20 [13] considering two approaches for NL modelling: analyses with the hyperbolic MKZ model and Masing rules modified with the MRDF-UIUC damping reduction factor at large strains (NL approach) and NL MRDF-UIUC analyses with the inclusion of shear strength τ max in the analyses (NL + τ max approach). For this last approach, no significant differences have been obtained using MKZ with Hashash adjustment to include the shear strength and GQ/H model directly incorporating τ max in the analyses [32]. In the following, the results of the NL analyses with shear strength only refer to the GQ/H model. The shear strength for each layer has been defined according to the Hardin and Drnevich [31] formulation in effective stress for sands as expressed by equation (5).

The comparison between the results obtained with the two approaches allows us to study the influence of shear strength as an input parameter in numerical site response analyses.

In order to assume a shear strength profile somewhat compatible with the stiffness gradient, for the sandy deposits, the friction angle of each layer has been obtained according to the [48] correlation:

(16) tan φ = 0.1 + 0.38 ln · q c σ v ,

where σ v = effective vertical stress in the midpoint of the layer considering water table at the ground surface and q c is related to small-strain shear modulus by [49]

(17) G 0 ( MPa ) = 49.5 · q c 0.51 .

Cohesion c’ has been considered equal to zero in equation (5)

The results were processed in terms of PGA and acceleration response spectrum (structural damping 5%) at ground surface. The comparisons were made using the following synthetic parameter Δ:

(18) Δ = Output A 1 Output A 2 ,

in which A1 and A2 are the outputs of the two analyses to be compared. In particular, in the previous equation, the “reference” analysis A2 is the approach typically used in local response analyses, i.e., the NL analysis without shear strength inclusion. In the following, the synthetic parameters Δ is represented as a function of the highest strain level reached in the maximum shear strains profile γ max (as computed in the “reference” A2 analysis) to investigate the differences in the two approaches at increasing NL levels.

The results of 8 (sandy deposits) × 30 (input motions) × 2 (modelling approaches) = 480 analyses carried out on sandy deposits (Table 1) are reported in Figure 2 in terms of Δ computed for PGA and spectral accelerations at 0.2, 0.5, 1, and 2 s period values.

Table 1

Summary of main parameters (thickness H, V s profiles; G/G 0 and D curves) adopted for parametric analyses on sandy deposits

H (m) V s,eq/V s, 30 (m/s) Curves G/G 0γ; D−γ
20 300; 450 z < 30 m Seed-Idriss 1970 – mean
40
60 z > 30 m Seed-Idriss 1970 – upper
80
Figure 2 
               Results of the parametric analyses on sandy deposits in terms of PGA and spectral accelerations at selected values of period: parameters Δ as a function of the highest strain reached in the maximum shear strain profile γ
                  max.
Figure 2

Results of the parametric analyses on sandy deposits in terms of PGA and spectral accelerations at selected values of period: parameters Δ as a function of the highest strain reached in the maximum shear strain profile γ max.

Below a peak shear strain of 0.1%, no significant differences exist in terms of PGA and spectral accelerations provided at the surface by the two approaches. At higher strains, the discrepancy becomes more significative: the differences are generally higher than 20% approaching maximum values of 30 and 60% for PGA and spectral accelerations, respectively. Note that Δ is generally lower than 1, indicating that the more accurate analyses including shear strength predict lower PGA and spectral accelerations. In other words, neglecting the shear strength leads to an overestimation of the PGA and response spectra at the surface. This behavior (i.e., the underestimation or overestimation) is obviously related to NL G/G 0 curves and the corresponding implied limit stress values in comparison to real shear strength of soils.

Similar considerations were obtained with reference to the 480 analyses carried out on clayey deposits, not reported here for sake of brevity [32].

This parametric study allowed therefore to define a preliminary threshold shear strain value, beyond which it is necessary to execute numerical analyses with more advanced models, able to capture the real behavior of the soil at large strains. Indicatively above shear strains of 0.1%, traditional NL models neglecting soil strength can provide unrealistic results, with significant mismatch of the seismic motion, up to 30 and 60% at the surface in terms of PGA and Response Spectra, respectively.

4 Lotung

4.1 Earthquake data

In the framework of a LSST program at a site near Lotung, a seismically active region in northeast Taiwan, the construction of ¼ and ½ scale models of a nuclear plant containment structure was carried out for soil-structure interaction research. The installed instrumentation (Figure 3) involves three linear surface arrays (arms 1, 2, and 3) and two downhole arrays (DHA and DHB). Down-hole array recordings are extremely useful to investigate soil amplification phenomena and to validate constitutive models and numerical modelling techniques for site response analyses [50]. Since its establishment in 1985, the Lotung site has been extensively studied by a large number of authors ([51,52,53,54,55,56,57] among others).

Figure 3 
                  Lotung array: (a) Plain view and (b) schematic cross-section of the site with the deployment of instruments at the ground surface and at the depths of 6, 11, 17, and 47 m.
Figure 3

Lotung array: (a) Plain view and (b) schematic cross-section of the site with the deployment of instruments at the ground surface and at the depths of 6, 11, 17, and 47 m.

The location of the downhole arrays, with respect to the edge of the ¼ scale model, is illustrated in Figure 3b. Extensive instrumentations was deployed to record seismic structural and ground response and to monitor soil PWP build-up. More specifically, the arrays were equipped with three-component accelerometers located at the ground surface (FA1-5) and approximately at depths of 6, 11, 17, and 47 m; in addition, a total of 27 pore pressure transducers have been installed. In this study, the free-field site response was studied using the downhole array DHB considering the recordings at borehole sensor located at 47 m depth as input motion.

During a 6-year operation from 1985 to 1990, 30 earthquakes of magnitude comprised between 4.0 and 6.5 triggered this array. In this study, in order to explore different levels of shaking, two events were considered: a weak event (LSST8) and one strong event (LSST7) to investigate small and high strain soil behavior, respectively. Events 12 and 16 (LSST12-16) were also used to validate the semi-empirical PWP models comparing numerical results with the pore water generation recorded at 6 m depth. As matter of fact, to the Authors knowledge, PWP recordings are available in the literature only for these two events.

The records of the transducers installed at 6 m are illustrated in Figure 4, showing a corresponding maximum normalized excess PWP r u of about 0.17 for event #12 and 0.25 for event #16.

Figure 4 
                  PWP build-up recorded during event LSST12 and LSST16 (after Zeghal et al., [65]).
Figure 4

PWP build-up recorded during event LSST12 and LSST16 (after Zeghal et al., [65]).

However, for these events (LSST12-16), the registration of the acceleration at 47 m depth is not available in the downhole array DHB; therefore, the 47 m depth recordings at downhole array DHA were employed as input in these analyses. Main seismological and ground motion parameters of the events selected for the analyses are listed in Table 2.

Table 2

Characteristics of LSST events considered in this study

Event# Date Local magnitude (M L) Epicentral distance (km) Focal depth (km) PGA (FA1-5)
EW (g) NS (g)
LSST7 5/20/86 6.5 66.2 15.8 0.16 0.21
LSST8 5/20/86 6.2 69.2 21.8 0.03 0.03
LSST12 7/30/86 6.2 5.2 1.6 0.16 0.19
LSST16 11/14/86 7.0 77.9 6.9 0.13 0.17

4.2 Geophysical and geotechnical data

Lotung geology consists of recent alluvium and Pleistocene materials approximately 40–50 and 350 m thick, respectively, overlying Miocene basement. Upper layers comprise about 30 m of predominantly silty sands and sandy silts with some gravels followed by 20 m of interlayered sandy silts, clayey silts, and silty clays [58]. The water table is at about 1 m depth from the ground surface.

As said before, several authors have studied this site, the shear wave velocity profiles measured and adopted by [52,55,59,60,61] are summarized in Figure 5a.

Figure 5 
                  (a) V
                     s profiles from various studies and (b) numerical viscoelastic linear amplification function.
Figure 5

(a) V s profiles from various studies and (b) numerical viscoelastic linear amplification function.

In this research, the soil profile and parameters assumed in the geotechnical model are referred from the study by Lee et al. [55]. The numerical viscoelastic linear amplification function between surface and the depth of 47 m computed assuming this V s profile and a uniform damping of 3% is reported in Figure 5b; the obtained fundamental frequency f 0 = 1.4 Hz is in good agreement with the experimental one in the linear range [53,62], thus confirming the reliability of the V s profile.

In Figure 6, normalized shear stiffness modulus and damping curves obtained from laboratory testing reported in the study by Tsai [58] and EPRI [63] are illustrated.

Figure 6 
                  Normalized modulus and damping curves for the Lotung site from various literature studies.
Figure 6

Normalized modulus and damping curves for the Lotung site from various literature studies.

In particular, EPRI [63] reports resonant column and torsional shear tests conducted on eight “undisturbed” samples of silty sands and silts at the University of Texas at Austin. The range of experimental curves represented as a grey region in Figure 6 suggests dynamic soil properties close to sands and low plasticity clays as expected being the plastic index of silt of about 7–8% (see the Seed curves [47] for reference). The researchers working on Lotung site have executed numerical analyses assuming different sets of curves. Chang et al. [59] performed equivalent linear analyses by two sets of curves assumed, respectively, for 0–6 and 6–17 m depth ranges, more recently used also by Lee et al. [55] in NL simulations of the Lotung site (Figure 6). Elgamal et al. [64] also computed three material curves from the in situ recorded seismic response for 0–6, 6–11 and 11–17 m layers. Since data were only available to a depth of 17 m, from 17 to 47 m were assumed the same curves as that at 11–17 m depth. This set of curves was also adopted by Borja et al. [52] and Stewart [61] in their seismic response verification studies. As may be observed, considering the stiffness modulus curves, the set proposed by Elgamal et al. [64] fall within the range proposed for sandy soils by Seed and Idriss [47], whereas the Chang et al. [59] curves, also employed by Lee et al. [55], fall along with the lower bound. Analogously, damping curves by Zeghal et al. [65] fall along the upper bound of the Seed and Idriss range, while curves by Chang et al. [59] exhibit much lower damping. As shown in Figure 6, a significant variation in the stiffness and damping curves is evident, confirming a large uncertainty in setting NL cyclic properties of soils.

4.3 Soil parameters and geotechnical model

The soil parameters assumed in the analyses are summarized in Table 3. As shown in Section 4.2, given the significant variation in terms of stiffness and damping curves, numerical analyses have been performed by using different sets of curves with reference to a weak (LSST8) and a strong event (LSST7) to calibrate the NL models in Deepsoil.

Table 3

Soil properties adopted for numerical simulations in Deepsoil

Soil z (m) γ (kN/m3) V s (m/s) ϕ′ (°) c′ (kPa) OCR τ max (kPa) G/G 0 and D curves
Silty sand 0–17 *16.45–22.50 *115–234 30 2 *16–70 see Table 4
Sand with gravel 17–23 19.33 234 35 97
Silty clay 23–29 244 24 10 122
Sand with gravel 29–36 286 37 160
Silty clay 36–47 262 24 10 190

Note: z: layer depth; V s: shear wave velocity; γ: unit weight; φ′: friction angle in terms of effective stress; c′: cohesion in terms of effective stress; OCR overconsolidation ratio; τ max shear strength according to Hardin & Drnevich formulation; “*” indicates variable values with the depth, minimum and maximum values are reported.

Table 4

Set of curves for the Lotung site used in the calibration analyses

Set curves Depth
0–17 m 17–47 m
1 Seed & Idriss 1970 – lower
2 Seed & Idriss 1970 – mean
3 Seed & Idriss (1970) – lower Seed & Idriss (1970) – mean
4 Seed & Idriss (1970) – mean Seed & Idriss (1970) – upper
5 Suwal et al. (2015) – 4b Darendeli (2001)
6 Suwal et al. (2015) – 4b Suwal et al. (2015) – 4c
7 Zeghal 6-11 Suwal et al. (2015) – 4b

Note that Seed & Idriss – Lower refers to the lower limit of G/G 0 curve and the D upper limit proposed by Authors while Seed & Idriss – Upper represents the upper limit of G/G 0 curve and the D lower limit. Seed & Idriss – Mean obviously refers to the average G/G 0 and D curves (Figure 6). Darendeli (2001) curves assigned to set 5 are referred to a plasticity index of PI = 0 because the layer 17–47 m consists of low plasticity clay (PI = 7–8) and sands; two PI = 0 curves were adopted varying the confining pressure (17–29 and 29–47 m). The curves by Suwal 4b and c are those used in numerical analyses performed by [67] considering the events LSST7 and LSST8.

The curves employed for the NL calibration involved both a unique set for the whole deposit consisting of Seed & Idriss Lower or Seed & Idriss Mean (set curves 1 and 2 in Table 4) and two sets of curves, respectively, for the 0–17 and 17–47 m layers (set curves 3–7). A summary of all curves is reported in Table 4 and Figure 7 where the employed curves are compared with the range defined by the experimental results reported in Figure 6.

Figure 7 
                  Normalized modulus and damping ratio curves for the Lotung site used in the calibration analyses.
Figure 7

Normalized modulus and damping ratio curves for the Lotung site used in the calibration analyses.

The Figures 8 and 9 show the results in terms of computed vs recorded acceleration response spectra at 17, 11, 6 m and at the surface (0 m).

Figure 8 
                  Recorded and numerical acceleration response spectra (ξ = 5%) for the Event LSST8; numerical analyses carried out using motion at 47 m as input (DHB47); East-West component on the left side and North-South component on the right side. Label 1–7 refer to the different sets of curves reported in Table 4.
Figure 8

Recorded and numerical acceleration response spectra (ξ = 5%) for the Event LSST8; numerical analyses carried out using motion at 47 m as input (DHB47); East-West component on the left side and North-South component on the right side. Label 1–7 refer to the different sets of curves reported in Table 4.

Figure 9 
                  Recorded and numerical acceleration response spectra (ξ = 5%) for the Event LSST7; numerical analyses carried out using motion at 47 m as input (DHB47); EW comp. on the left side and NS comp. on the right side. Label 1–7 refer to the different sets of curves reported in Table 4.
Figure 9

Recorded and numerical acceleration response spectra (ξ = 5%) for the Event LSST7; numerical analyses carried out using motion at 47 m as input (DHB47); EW comp. on the left side and NS comp. on the right side. Label 1–7 refer to the different sets of curves reported in Table 4.

For the weak event, LSST8, there is no substantial differences in the numerical results of the analyses carried out using the different sets of curves; each set provides a good agreement with the recordings at all depths. On the contrary, the analyses of the strong event, LSST7, highlight two groups of curves providing a substantial different performances: sets 1–2–3 show a poor prediction of ground motion with a significant underestimation of recorded spectral accelerations in the whole period range and at all depths while a satisfactory agreement with recordings is obtained with curves 4–5–6–7. The selection of the best set is not an easy task (Huang et al., [66]) and is somewhat based on subjective expert judgment. A simple objective statistical analysis has been performed to reach a decision. In the following analyses, set 6 that correspond to Suwal et al. – 4b curves for 0–17 m and Suwal et al. – 4c curves for 17–47 m, has been employed as it provides the lower mean residual (Figure 10):

[19] R = 1 N i = 1 N ( S a , num S a , rec ) 2 ,

where S a,num and S a,rec are the numerical and recorded spectral acceleration, respectively, and N is the number of points at which the response spectrum is defined between 0 and 2 s.

Figure 10 
                  Performance of the different NL curves: residuals between recorded and numerical spectra for events LSST8 and LSST7.
Figure 10

Performance of the different NL curves: residuals between recorded and numerical spectra for events LSST8 and LSST7.

4.4 Influence of shear strength as input parameter in numerical site response analyses

The parameters employed in the analysis are reported in Table 3. The shear strength was computed by the Hardin & Drnevich formulation equation (6) in which the friction angle and cohesion in terms of effective stress (c′ and φ′) have been assigned according to the modelling assumed by Amorosi et al. [57] based on available experimental data of the Lotung site. In Figure 11 is reported, for Event 7, a comparison of the MKZ (no shear strength included in the analyses) and the GQ/H model (accounting for τ max) to assess the influence of shear strength as an input parameter in numerical analyses. The results indicate a slight improvement of the numerical prediction at all sensors by introducing the shear strength in the analyses only for the EW component; no appreciable differences can be noted for NS component. Results can be interpreted by observing the stress–strain (τγ) curves associated with MKZ and GQ/H models for representative soil elements at 7.5, 20.5, 26.5, and 31.5 m (Figure 12). The GQ/H stress–strain curve obviously perfectly matches the shear strength while MKZ overestimates the maximum shear stress especially for strains higher than 1%. In this analysis, the maximum shear strain experienced by soils is lower than 0.2% for EW component and even lower (<0.1%) for NS component. In this strain range, the discrepancies between MKZ and GQ/H stress curves are quite limited; therefore, the explicit inclusion of shear strength in the modelling lead only to a slight improvement of the prediction. It should be noted that the calibration of NL reference curves reported above allowed us to select curves (set #6) providing a satisfactory numerical prediction of event 7 and therefore modelling a response already compatible with the soil behavior at medium-to-high strains. The explicit inclusion of shear strength in the modelling with the GQ/H analyses allowed a further refinement of the prediction with respect to the standard MKZ NL modelling.

Figure 11 
                  Recorded and numerical acceleration response spectra (ξ = 5%) for the Event LSST7; numerical analyses carried out with MKZ (no shear strength included in the analyses) and the GQ/H (accounting for τ
                     max) models using motion at 47 m as input (DHB47); EW comp. on the left side and NS comp. on the right side.
Figure 11

Recorded and numerical acceleration response spectra (ξ = 5%) for the Event LSST7; numerical analyses carried out with MKZ (no shear strength included in the analyses) and the GQ/H (accounting for τ max) models using motion at 47 m as input (DHB47); EW comp. on the left side and NS comp. on the right side.

Figure 12 
                  Normalized shear modulus curves from MKZ and GQ/H models (on the left) and stress–strain relationship (on the right) for the soil elements at 7.5, 20.5, 26.5, and 31.5 m depth. The reference curves are Suwal et al. (2015) 4b and c for layers 0–17 and 17–47 m, respectively.
Figure 12

Normalized shear modulus curves from MKZ and GQ/H models (on the left) and stress–strain relationship (on the right) for the soil elements at 7.5, 20.5, 26.5, and 31.5 m depth. The reference curves are Suwal et al. (2015) 4b and c for layers 0–17 and 17–47 m, respectively.

It is important to note that the accuracy of the numerical prediction obtained in this study is consistent with previous research utilizing more advanced elasto-plastic NL models [57,68,69]. These studies have also shown that the numerical prediction for event 7 is less accurate for the NS component where an underestimation of recorded response is generally observed like in the present study.

4.5 NL analyses with PWP models

NL analyses have finally been performed by coupling to the GQ/H model with two different semi-empirical PWP models (described in Section 2.4): the Dobry/Matasovic model for the silty sand and sand with gravel layers (0–17, 17–23, 29–36 m), and the Matasovic and Vucetic model for the silty clays (23–29 and 36–47 m layers). A summary of the parameters of PWP models is reported in Table 5. The analyses were carried out for LSST12 and LSST16 events, for which pore pressure recordings are available.

Table 5

Summary of PWP models parameters

Soil z (m) Dobry/Matasovic (sands) Matasovic and Vucetic (clays)
f p F s γ s r A B C D
(−) (−) (−) (−) (%) (−) (−) (−) (−) (−) (−)
Silty sand 0–17 2 1 0.8–2.4 1.35–1.63 0.01
Sand with gravel 17–23
Silty clay 23–29 0.33 0.59 13.23 26.80 15.51 1.92
Sand with gravel 29–36 2 1 0.68 1.35 0.01
Silty clay 36–47 0.33 0.59 13.23 26.80 15.51 1.92

In particular, for the Dobry/Matasovic formulation, the f parameter is used to account for loading in multiple dimensions; in this analysis, it has been set equal to 2 to take into correctly the dimensionality of this problem. The parameter p is set to be 1, as suggested for practical purposes. F and s parameters have been computed according to equations (11) and (12) based on the shear wave velocity and estimated fine content of the materials. Regarding the Matasovic and Vucetic models, the fitting parameters have been obtained by the suggested empirical correlations proposed by Carlton [42] as a function of OCR and PI of soils. No information is available for the volumetric threshold γ tvp , it was set equal to 0.01% for both models, which is a typical value for sandy and low plasticity soils [44,45].

Figure 13 shows the comparison between the recorded and numerical response spectra at the different sensors for LSST12 and LSST16 events using GQ/H total stress analysis and GQ/H coupled with PWP models (GQ/H PWP models). Reference is made to NS component; similar considerations can be made for EW component. The addition of the PWP generation models does not significantly affect the seismic response prediction in terms of response spectra with respect to total stress analyses.

Figure 13 
                  Recorded and numerical acceleration response spectra (ξ = 5%) for Events LSST12 (left side) and LSST16 (right side): comparison between GQ/H and GQ/H PWP analyses (NS component).
Figure 13

Recorded and numerical acceleration response spectra (ξ = 5%) for Events LSST12 (left side) and LSST16 (right side): comparison between GQ/H and GQ/H PWP analyses (NS component).

The pore pressure comparison in terms of normalized excess PWP time history at 6 m depth is finally shown in Figure 14 for both LSST12 and LSST16 events, with reference to the NS component. For both events, the excess PWP is underestimated even if the order of magnitude is captured by the simplified model. A satisfactory prediction can be observed for LSST16: the recorded pore-pressures reach a peak value of 0.25 with a 0.2 residual value after a dissipation; the numerical model shows a residual r u of about 0.15 with a 25% error with respect to the experimental value. For comparison, for the same event, a residual value of 0.45 at 6 m (almost double with respect to the recorded one) is predicted by Elgamal et al. [70] with a multi-surface plasticity model and u-p solid-fluid formulation. Even if more research for other events is necessary, the simplified NL strategy proposed here show an accuracy comparable with more advanced models also in terms of PWP build-up.

Figure 14 
                  Build-up pore pressure normalized by the effective vertical stress for events LSST12 and LSST16: computed vs recorded time history at 6 m depth.
Figure 14

Build-up pore pressure normalized by the effective vertical stress for events LSST12 and LSST16: computed vs recorded time history at 6 m depth.

The low values of excess PWP reached in the analyses for LSST12 and LSSTR16 justify the negligible differences shown by GQ/H and GQ/H PWP analyses in terms of response spectra observed in Figure 13.

5 Conclusion

The shear strength of soil is generally neglected in site response analyses at large strains leading to unrealistic predicted seismic response of soil deposits. Moreover, above the volumetric threshold, the pore pressure build-up causes degradation of soil properties that can strongly affect the NL behavior. The simplified NL models of the hyperbolic family largely used in site response analyses are generally calibrated on stiffness and damping curves and not necessarily represent the large-strain shear stresses of soil; moreover, PWP build-up models are also usually neglected in site response being used only for liquefaction analyses.

The aim of this study was to validate a simplified NL procedure based on hyperbolic model including both shear strength and PWP build-up, implemented in 1D Deepsoil computer code. The first part of the study identified limiting shear strain values, beyond which it is highly recommended to execute numerical analyses including shear strength. A large parametric study on ideal sandy and clayey deposits has been presented by comparing two modelling approaches: NL analyses with the hyperbolic model and non Masing rules and the NL analyses with the hyperbolic model including shear strength. The study shows that above a shear strain of 0.1%, that is easily reached in the case of highly deformable soils and medium-to-strong input motions, it is recommended to consider, in addition to the stiffness and damping curves, the shear strength of the soil to achieve a more realistic prediction of site response. In this strain range, it is also suggested to add models for excess PWPs being the volumetric threshold in the order of 0.01–0.1% moving from sandy to high plasticity clayey soils.

The proposed NL approach, implemented in the Deepsoil code, has been then applied to the Lotung array site widely studied in the literature, also with advanced NL elasto-plastic models. Given the large uncertainties in NL cyclic properties of soils highlighted by experimental data and previous studies, the first part of the work consisted of a sensitivity analysis on the influence of different NL shear modulus and damping curves. The set of curves providing the lower mean residuals with respect to experimental recordings in terms of acceleration response spectra has been selected. This calibration allowed us to get good prediction for both weak and strong motion recordings. A further refinement of modeling reliability, fully consistent with results produced by more advanced elasto-plastic NL models, has then been obtained by including shear strength in site response analyses. NL analyses have been finally carried out for two moderate events in which PWP build-up recordings are available. For both events, the computed excess PWP is underestimated even if the order of magnitude is captured by the PWP simplified empirical models showing an accuracy comparable or even better than advanced models used in the literature.

As a conclusion, the NL models here adopted require the same input data adopted in the traditional equivalent linear approaches (V s, G, and D curves) and physical-mechanical properties routinely determined during geotechnical surveys (i.e., shear strength, relative density, fine content). The hyperbolic NL models coupled with semi-empirical PWP generation and the inclusion of soil shear strength can be therefore applied in routine site response analyses with no costs aggravation both at local (i.e., determination of seismic action on buildings) and urban scale (detailed microzonation studies in terms of ground shaking and liquefaction).

  1. Author contributions: Francesco di Buccio: Methodology, Data curation, Numerical analyses, Writing – original draft, Writing – review & editing. Alessandro Pagliaroli: Conceptualization, Methodology, Supervision, Data curation, Writing – review & editing.

  2. Conflict of interest: The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

References

[1] Pagliaroli A. Key issues in seismic microzonation studies: lessons from recent experiences in Italy. Riv Ital di Geotech – Italian Geotech J. 2018;52(1):5–48. 10.19199/2018.1.0557-1405.05.Search in Google Scholar

[2] Karkush MO. Numerical analysis of piles group surrounded by grouting under seismic load. In Geotechnical engineering and sustainable construction. Singapore: Springer; 2022. p. 379–89. 10.1007/978-981-16-6277-5.Search in Google Scholar

[3] Al-Mirza HA, Karkush MO. Numerical modeling of circular tunnel alignment under seismic loading. In Geotechnical engineering and sustainable construction. Singapore: Springer; 2022. p. 15–27. 10.1007/978-981-16-6277-5_2.Search in Google Scholar

[4] Ali AM, Karkush MO, Al-Jorany AN. Numerical modeling of connected piled raft foundation under seismic loading in layered soils. J Mech Behav Mater. 2023;32(1):20220250. 10.1515/jmbm-2022-0250.Search in Google Scholar

[5] Régnier J, Bonilla L, Bard P, Bertrand E, Hollender F. International benchmark on numerical simulations for 1D, nonlinear site response (PRENOLIN): verification phase based on canonical cases. Bull Seismol Soc Am. 2016;106(5):2112–35. 10.1785/0120150284.Search in Google Scholar

[6] Hashash YMA, Phillips C, Groholski DR. Recent advances in non-linear site response analysis. Proceedings of the 5th International Conference on Recent Advances in Geotechnical Earthquake Engineering and Soil Dynamics. Rolla, MO: Missouri University of Science and Technology; 2010. p. 0–22.Search in Google Scholar

[7] Prevost JH. DYNA1D: A computer program for nonlinear site response analysis. Technical documentation. Soil Dyn Earthquake Eng. 1989;63:98–109.Search in Google Scholar

[8] Schanz T, Vermeer PA. BPG. The hardening soil model: Formulation and verification. In Beyond 2000 in computational geotechnics. Rotterdam, The Netherlands: Balkema; 1999. p. 281–96.10.1201/9781315138206-27Search in Google Scholar

[9] Pisanò F, Jeremi B. Simulating stiffness degradation and damping in soils via a simple visco-elastic – plastic model. Soil Dyn Earthquake Eng. 2014;63:98–109. 10.1016/j.soildyn.2014.02.014.Search in Google Scholar

[10] Elgamal A, Yang ZLJ. Cyclic1D: a computer program for seismic ground response. San Diego, La Jolla, CA; 2006.Search in Google Scholar

[11] Biot MA. The mechanism of deformation and acoustic propagation in porous media. J Appl Phys. 1962;33(4):1482–98.10.1063/1.1728759Search in Google Scholar

[12] Sica S. Analisi del comportamento dinamico di dighe in terra. PhD Dissertation in Geotechnical Engineering. University of Rome ‘La Sapienza’ and University of Naples ‘Federico II’; 2001.Search in Google Scholar

[13] Park D, Hashash YMA. Soil damping formulation in nonlinear time domain site response analysis. J Earthq Eng. 2004;8(2):249–74.Search in Google Scholar

[14] Matasović NOGA. A computer program for seismic response analysis of horizontally layered soil deposits, earthfill dams, and solid waste landfills. Lacey, Washington, D.C; 2010.Search in Google Scholar

[15] Conti R, Angelini M, Licata V. Nonlinearity and strength in 1D site response analyses: a simple constitutive approach. Bull Earthquake Eng. 2020;18(10):4629–57. 10.1007/s10518-020-00873-5.Search in Google Scholar

[16] Hashash YMA, Dashti S, Romero MI, Ghayoomi M, Musgrove M. Evaluation of 1-D seismic site response modeling of sand using centrifuge experiments. Soil Dyn Earthquake Eng. 2015;78:19–31. 10.1016/j.soildyn.2015.07.003.Search in Google Scholar

[17] Shi J, Asimaki D. From stiffness to strength: formulation and validation of a hybrid hyperbolic nonlinear soil model for site-response analyses. Bull Seismol Soc Am. 2017;107(3):1336–55. 10.1785/0120150287.Search in Google Scholar

[18] Aaqib M, Sadiq S, Park D, Hashash YMA, Pehelivan M. Importance of implied strength correction for 1D site response at shallow sites at a moderate to low seismicity region. In Proceedings of Geotechnical Earthquake Engineering and Soil Dynamics V. Austin, Texas, USA: 2018. 10.1061/9780784481462.043.10.1061/9780784481462.043Search in Google Scholar

[19] Schnabel John Lysmer H, Bolton Seed A. Computer program for earthquake response analysis of horizontally layered sites report No. EERC 72 - December 1972 College of Engineering Univeristy of California Berkeley, California; 1972.Search in Google Scholar

[20] Kottke ARREM. Technical manual for Strata. Berkeley, California: Pacific Earthquake Engineering Research Center; 2009.Search in Google Scholar

[21] Matasovic N. Seismic response of composite horizontally-layered soil deposits. Vol. 119, No. 11, Los Angeles, USA: University of California; 1993. p. 1805–22.Search in Google Scholar

[22] Groholski DR, Hashash YMA, Kim B, Musgrove M, Harmon J, Stewart JP. Simplified model for small-strain nonlinearity and strength in 1D seismic site response analysis. J Geotech Geoenviron Eng. 2016;142(9):04016042. 10.1061/(asce)gt.1943-5606.0001496.Search in Google Scholar

[23] Masing G. Eigenspannungen und verfestigung beim messing. Proceedings of the 2nd International Congress on Applied Mechanics. Zurich, Switzerland: Orell Füssli; 1926.Search in Google Scholar

[24] Verrucci L, Pagliaroli A, Lanzo G, Di Buccio F, Pio A, Cucci C. Damping formulations for finite difference linear dynamic analyses: Performance and practical recommendations. Comput Geotech. 2022;142(February):104568. 10.1016/j.compgeo.2021.104568.Search in Google Scholar

[25] Park D, Hashash YMA. Soil damping formulation in nonlinear time domain site response analysis. J Earthquake Eng. 2004;8(2):249–74.10.1080/13632460409350489Search in Google Scholar

[26] Chiaradonna A. Development and assessment of a numerical model for non-linear coupled analysis on seismic response of liquefiable soils. PhD thesis, University of Napoli Federico II; 2016.Search in Google Scholar

[27] Kondner RL. A hyperbolic stress-strain formulation for sands. Proceedings of the Second Panamerican Conference on Soil Mechanics and Foundation Engineering. Vol. 1, Brazil: 1963. p. 289–324.Search in Google Scholar

[28] Hashash YMA, Park D. Non-linear one-dimensional seismic ground motion propagation in the Mississippi embayment. 2001;62(1–3):185–206.10.1016/S0013-7952(01)00061-8Search in Google Scholar

[29] Phillips C, Hashash YMA. Damping formulation for nonlinear 1D site response analyses. 2009;29:1143–58. 10.1016/j.soildyn.2009.01.004.Search in Google Scholar

[30] Ishihara K, Kasuda K. Dynamic strength of cohesive soil. Proceedings of the Sixth Budapest Conference on Soil Mechanics and Foundation Engineering. Budapest, Hungary: 1984.Search in Google Scholar

[31] Hardin BO, Drnevich VP. Shear modulus and damping in soils. J Soil Mech Found Div. 1972;98(7):667–92.10.1061/JSFEAQ.0001760Search in Google Scholar

[32] Di Buccio F. Numerical modelling of seismic response at large strains with simplified nonlinear models. PhD dissertation. Italy: University “G. d’Annunzio” of Chieti Pescara; 2021.Search in Google Scholar

[33] Park D, Ahn J. Accumulated stress based model for prediction of residual pore pressure. IOS Press; Vol. 1, p. 1567–70.Search in Google Scholar

[34] Matasović N, Vucetic M. Generalized cyclic-degradation-pore-pressure generation model for clays. J Geotech Eng. 1995;121(1):33–42.Search in Google Scholar

[35] Green RA, Mitchell JK, Polito CP. An energy-based pore pressure generation model for cohesionless soils. In Proceedings of the John Booker Memorial Symposium, Melbourne, Australia, 2000.Search in Google Scholar

[36] Berrill JB, Davis RO. Energy dissipation and seismic liquefaction of sands: revised model. Soils Found. 1985;25(2):106–18. 10.3208/sandf1972.25.2_106 Search in Google Scholar

[37] Dobry R, Vasquez-Herrera A, Mohamad R, Vucetic M. Liquefaction flow failure of silty sand by torsional cyclic tests. Advances in the art of testing soils under cyclic condition. ASCE; 1985. p. 29–50.Search in Google Scholar

[38] Matasović N, Vucetic M. Cyclic characterization of liquefiable sands. J Geotech Eng. 1993;119(11):1805–22.10.1061/(ASCE)0733-9410(1993)119:11(1805)Search in Google Scholar

[39] Matasović N, Vucetic M. Generalized cyclic-degradation-pore pressure generation model for clays. J Geotech Eng. 1995;121(1):33–42.10.1061/(ASCE)0733-9410(1995)121:1(33)Search in Google Scholar

[40] Vucetic MDR. Pore pressure buildup and liquefaction at level sandy sites during earthquakes. California, Japan: Rensselaer Polytechnic Institute; 1986.Search in Google Scholar

[41] Vucetic MDR. Cyclic triaxial strain-controlled testing of liquefiable sands. ASTM International; 1988.10.1520/STP29093SSearch in Google Scholar

[42] Carlton B. An improved description of the seismic response of sites with high plasticity soils, organic clays, and deep soft soil deposit. PhD dissertation. Berkeley: University of California; 2014.Search in Google Scholar

[43] Mei X, Olson SHY. Empirical curve-fitting parameters for a porewater pressure generation model for use in 1-D effective stress-based site response analysis. In Proc, 6th Int Conf on Earthquake Geotechnical Engineering Technical Committee on Earthquake Geotechnical Engineering and Associated Problems (TC203) of ISSMGE. Wellington, New Zealand: 2015.Search in Google Scholar

[44] Vucetic M. Cyclic threshold shear strains in soils. J Geotech Eng. 1994;120(12):2208–28.10.1061/(ASCE)0733-9410(1994)120:12(2208)Search in Google Scholar

[45] Ciancimino A, Maria R, Sebastiano C, Lanzo G, Pagliaroli A, Pallara O. The poliTO– UniRoma1 database of cyclic and dynamic laboratory tests: assessment of empirical predictive models [Internet]. Bulletin of Earthquake Engineering. Vol. 21, Netherlands: Springer; 2023. p. 2569–601. 10.1007/s10518-022-01573-y.Search in Google Scholar

[46] Darendeli MB. Development of a new family of normalized modulus reduction and material damping. IEEE Trans Commun. 2001;25(3):374–9. 10.1109/TCOM.1977.1093818.Search in Google Scholar

[47] Seed HB, Idriss IM. Soil moduli and damping factors for dynamic response analyses. Report No. EERC 7010, Earthquake Engineering Research Center, University of California, Berkeley, California. 1970; p. 40.Search in Google Scholar

[48] Robertson PKCRG. Interpretation of cone penetration tests. Part I: Sand. Can Geotech J. 1983;20(4):718–33.10.1139/t83-078Search in Google Scholar

[49] Simonini P, Cola S. Use of piezocone to predict maximum stiffness of Venetian soils. J Geotech Geoenviron Eng. 2000;126(4):378–82.10.1061/(ASCE)1090-0241(2000)126:4(378)Search in Google Scholar

[50] Régnier J, Bonilla L, Bard P, Bertrand E, Hollender F. International benchmark on numerical simulations for 1D, nonlinear site response (PRENOLIN): verification phase based on canonical cases. Bull Seismol Soc Am. 2018;108(2):876–900. 10.1785/0120170210.Search in Google Scholar

[51] Li XS, Shen CKWZ. Fully coupled inelastic site response analysis for the 1986 Lotung earthquake. J Geotech Geoenviron Eng. 1998;124(7):560–73.10.1061/(ASCE)1090-0241(1998)124:7(560)Search in Google Scholar

[52] Borja RI, Chao HY, Montáns FJLCH. Nonlinear ground response at Lotung LSST site. J Geotech Geoenviron Eng. 1999;125(3):187–97.10.1061/(ASCE)1090-0241(1999)125:3(187)Search in Google Scholar

[53] Huang H-C, Shieh C-SCH-C. Linear and nonlinear behaviors of soft soil layers using Lotung downhole array in Taiwan. Terr Atmos Ocean Sci. 2001;12(3):503–24.Search in Google Scholar

[54] Borja RI, Duvernay BG, Lin C. Ground response in Lotung: total stress analyses and parametric studies. 2002;128(1):54–63.10.1061/(ASCE)1090-0241(2002)128:1(54)Search in Google Scholar

[55] Lee C, Tsai Y, Wen K. Analysis of nonlinear site response using the LSST downhole accelerometer array data. 2006;26:435–60. 10.1016/j.soildyn.2005.10.005.Search in Google Scholar

[56] Borja RI, Ching ÆW. Estimating inelastic sediment deformation from local site response simulations. 2007;2:183–95. 10.1007/s11440-007-0044-x.Search in Google Scholar

[57] Amorosi A, Boldini D, di Lernia A. Seismic ground response at Lotung: Hysteretic elasto-plastic-based 3D analyses. Soil Dyn Earthquake Eng. 2016;85:44–61. 10.1016/j.soildyn.2016.03.001.Search in Google Scholar

[58] Tsai J. Validation of soil-structure interaction analysis methods. Berkeley: University of California; 1990.Search in Google Scholar

[59] Chang C, Mok C, Power M, Tang Y, Tang HSJ. Equivalent linear versus nonlinear ground response analyses at Lotung seismic experiment site. Fourth US National Conference on Earthquake Engineering; 1990. p. 327–36.Search in Google Scholar

[60] EPRI. Guidelines for determining design ground motions. Appendices for ground motion estimation, TR 102293. Vol. 2, 1993.Search in Google Scholar

[61] Stewart JPKAO. Nonlinear seismic ground response analysis: Code usage protocols and verification against vertical array data. Geotech Earthquake Eng soil Dyn. 2008;4:1–24.10.1061/40975(318)1Search in Google Scholar

[62] Huang H-C, Shieh C-SCH-C. Linear and nonlinear behaviors of soft soil layers using Lotung downhole array in Taiwan. Proceedings of the 12th World Conference on Earthquake Engineering. Auckland, New Zealand: 2000.10.3319/TAO.2001.12.3.503(T)Search in Google Scholar

[63] EPRI. Guidelines for determining design ground motions. Appendices for laboratory investigations, TR 102293. Vol. 4, 1993.Search in Google Scholar

[64] Elgamal A, Zeghal M, Forecasting I, Box PO, Tang HT. Identification and modeling of earthquake ground response I. Site Amplif. 1996;15:499–522.Search in Google Scholar

[65] Zeghal M, Elgamal A-W, Tang HSJ. Lotung downhole array. II: Evaluation of soil nonlinear properties. J Geotech Eng. 1995;121(4):363–78.10.1061/(ASCE)0733-9410(1995)121:4(363)Search in Google Scholar

[66] Huang G, Xiao L, Zhang G. An integrated design concept evaluation method based on best-worst best–worst entropy and generalized TODIM considering multiple factors of uncertainty. Appl Soft Comput. 2023;40:110165.10.1016/j.asoc.2023.110165Search in Google Scholar

[67] Suwal S, Pagliaroli A, Lanzo G. Numerical modelling of site response at the LSST downhole array in Lotung. Geotechnical earthquake engineering and soil dynamics; 2015.Search in Google Scholar

[68] Amorosi A, Boldini D, di Lernia A. Dynamic soil-structure interaction: A three-dimensional numerical approach and its application to the Lotung case study. Comput Geotech. 2017;90:34–54. 10.1016/j.compgeo.2017.05.016.Search in Google Scholar

[69] Elia G, Rouainia M, Karofyllakis D, Guzel Y. Modelling the non-linear site response at the LSST down-hole accelerometer array in Lotung. 2018;102:1–14.10.1016/j.soildyn.2017.08.007Search in Google Scholar

[70] Elgamal A, Zeghai M, Parra E, Gunturi R, Tang HT, Stepp JC. Identification and modeling of earthquake ground response – I. Site amplification. Soil Dyn Earthquake Eng. 1996;15:499–552.10.1016/S0267-7261(96)00021-8Search in Google Scholar

Received: 2023-10-03
Revised: 2024-02-28
Accepted: 2024-03-18
Published Online: 2024-09-18

© 2024 the author(s), published by De Gruyter

This work is licensed under the Creative Commons Attribution 4.0 International License.

Articles in the same Issue

  1. Regular Articles
  2. Theoretical magnetotelluric response of stratiform earth consisting of alternative homogeneous and transitional layers
  3. The research of common drought indexes for the application to the drought monitoring in the region of Jin Sha river
  4. Evolutionary game analysis of government, businesses, and consumers in high-standard farmland low-carbon construction
  5. On the use of low-frequency passive seismic as a direct hydrocarbon indicator: A case study at Banyubang oil field, Indonesia
  6. Water transportation planning in connection with extreme weather conditions; case study – Port of Novi Sad, Serbia
  7. Zircon U–Pb ages of the Paleozoic volcaniclastic strata in the Junggar Basin, NW China
  8. Monitoring of mangrove forests vegetation based on optical versus microwave data: A case study western coast of Saudi Arabia
  9. Microfacies analysis of marine shale: A case study of the shales of the Wufeng–Longmaxi formation in the western Chongqing, Sichuan Basin, China
  10. Multisource remote sensing image fusion processing in plateau seismic region feature information extraction and application analysis – An example of the Menyuan Ms6.9 earthquake on January 8, 2022
  11. Identification of magnetic mineralogy and paleo-flow direction of the Miocene-quaternary volcanic products in the north of Lake Van, Eastern Turkey
  12. Impact of fully rotating steel casing bored pile on adjacent tunnels
  13. Adolescents’ consumption intentions toward leisure tourism in high-risk leisure environments in riverine areas
  14. Petrogenesis of Jurassic granitic rocks in South China Block: Implications for events related to subduction of Paleo-Pacific plate
  15. Differences in urban daytime and night block vitality based on mobile phone signaling data: A case study of Kunming’s urban district
  16. Random forest and artificial neural network-based tsunami forests classification using data fusion of Sentinel-2 and Airbus Vision-1 satellites: A case study of Garhi Chandan, Pakistan
  17. Integrated geophysical approach for detection and size-geometry characterization of a multiscale karst system in carbonate units, semiarid Brazil
  18. Spatial and temporal changes in ecosystem services value and analysis of driving factors in the Yangtze River Delta Region
  19. Deep fault sliding rates for Ka-Ping block of Xinjiang based on repeating earthquakes
  20. Improved deep learning segmentation of outdoor point clouds with different sampling strategies and using intensities
  21. Platform margin belt structure and sedimentation characteristics of Changxing Formation reefs on both sides of the Kaijiang-Liangping trough, eastern Sichuan Basin, China
  22. Enhancing attapulgite and cement-modified loess for effective landfill lining: A study on seepage prevention and Cu/Pb ion adsorption
  23. Flood risk assessment, a case study in an arid environment of Southeast Morocco
  24. Lower limits of physical properties and classification evaluation criteria of the tight reservoir in the Ahe Formation in the Dibei Area of the Kuqa depression
  25. Evaluation of Viaducts’ contribution to road network accessibility in the Yunnan–Guizhou area based on the node deletion method
  26. Permian tectonic switch of the southern Central Asian Orogenic Belt: Constraints from magmatism in the southern Alxa region, NW China
  27. Element geochemical differences in lower Cambrian black shales with hydrothermal sedimentation in the Yangtze block, South China
  28. Three-dimensional finite-memory quasi-Newton inversion of the magnetotelluric based on unstructured grids
  29. Obliquity-paced summer monsoon from the Shilou red clay section on the eastern Chinese Loess Plateau
  30. Classification and logging identification of reservoir space near the upper Ordovician pinch-out line in Tahe Oilfield
  31. Ultra-deep channel sand body target recognition method based on improved deep learning under UAV cluster
  32. New formula to determine flyrock distance on sedimentary rocks with low strength
  33. Assessing the ecological security of tourism in Northeast China
  34. Effective reservoir identification and sweet spot prediction in Chang 8 Member tight oil reservoirs in Huanjiang area, Ordos Basin
  35. Detecting heterogeneity of spatial accessibility to sports facilities for adolescents at fine scale: A case study in Changsha, China
  36. Effects of freeze–thaw cycles on soil nutrients by soft rock and sand remodeling
  37. Vibration prediction with a method based on the absorption property of blast-induced seismic waves: A case study
  38. A new look at the geodynamic development of the Ediacaran–early Cambrian forearc basalts of the Tannuola-Khamsara Island Arc (Central Asia, Russia): Conclusions from geological, geochemical, and Nd-isotope data
  39. Spatio-temporal analysis of the driving factors of urban land use expansion in China: A study of the Yangtze River Delta region
  40. Selection of Euler deconvolution solutions using the enhanced horizontal gradient and stable vertical differentiation
  41. Phase change of the Ordovician hydrocarbon in the Tarim Basin: A case study from the Halahatang–Shunbei area
  42. Using interpretative structure model and analytical network process for optimum site selection of airport locations in Delta Egypt
  43. Geochemistry of magnetite from Fe-skarn deposits along the central Loei Fold Belt, Thailand
  44. Functional typology of settlements in the Srem region, Serbia
  45. Hunger Games Search for the elucidation of gravity anomalies with application to geothermal energy investigations and volcanic activity studies
  46. Addressing incomplete tile phenomena in image tiling: Introducing the grid six-intersection model
  47. Evaluation and control model for resilience of water resource building system based on fuzzy comprehensive evaluation method and its application
  48. MIF and AHP methods for delineation of groundwater potential zones using remote sensing and GIS techniques in Tirunelveli, Tenkasi District, India
  49. New database for the estimation of dynamic coefficient of friction of snow
  50. Measuring urban growth dynamics: A study in Hue city, Vietnam
  51. Comparative models of support-vector machine, multilayer perceptron, and decision tree ‎predication approaches for landslide ‎susceptibility analysis
  52. Experimental study on the influence of clay content on the shear strength of silty soil and mechanism analysis
  53. Geosite assessment as a contribution to the sustainable development of Babušnica, Serbia
  54. Using fuzzy analytical hierarchy process for road transportation services management based on remote sensing and GIS technology
  55. Accumulation mechanism of multi-type unconventional oil and gas reservoirs in Northern China: Taking Hari Sag of the Yin’e Basin as an example
  56. TOC prediction of source rocks based on the convolutional neural network and logging curves – A case study of Pinghu Formation in Xihu Sag
  57. A method for fast detection of wind farms from remote sensing images using deep learning and geospatial analysis
  58. Spatial distribution and driving factors of karst rocky desertification in Southwest China based on GIS and geodetector
  59. Physicochemical and mineralogical composition studies of clays from Share and Tshonga areas, Northern Bida Basin, Nigeria: Implications for Geophagia
  60. Geochemical sedimentary records of eutrophication and environmental change in Chaohu Lake, East China
  61. Research progress of freeze–thaw rock using bibliometric analysis
  62. Mixed irrigation affects the composition and diversity of the soil bacterial community
  63. Examining the swelling potential of cohesive soils with high plasticity according to their index properties using GIS
  64. Geological genesis and identification of high-porosity and low-permeability sandstones in the Cretaceous Bashkirchik Formation, northern Tarim Basin
  65. Usability of PPGIS tools exemplified by geodiscussion – a tool for public participation in shaping public space
  66. Efficient development technology of Upper Paleozoic Lower Shihezi tight sandstone gas reservoir in northeastern Ordos Basin
  67. Assessment of soil resources of agricultural landscapes in Turkestan region of the Republic of Kazakhstan based on agrochemical indexes
  68. Evaluating the impact of DEM interpolation algorithms on relief index for soil resource management
  69. Petrogenetic relationship between plutonic and subvolcanic rocks in the Jurassic Shuikoushan complex, South China
  70. A novel workflow for shale lithology identification – A case study in the Gulong Depression, Songliao Basin, China
  71. Characteristics and main controlling factors of dolomite reservoirs in Fei-3 Member of Feixianguan Formation of Lower Triassic, Puguang area
  72. Impact of high-speed railway network on county-level accessibility and economic linkage in Jiangxi Province, China: A spatio-temporal data analysis
  73. Estimation model of wild fractional vegetation cover based on RGB vegetation index and its application
  74. Lithofacies, petrography, and geochemistry of the Lamphun oceanic plate stratigraphy: As a record of the subduction history of Paleo-Tethys in Chiang Mai-Chiang Rai Suture Zone of Thailand
  75. Structural features and tectonic activity of the Weihe Fault, central China
  76. Application of the wavelet transform and Hilbert–Huang transform in stratigraphic sequence division of Jurassic Shaximiao Formation in Southwest Sichuan Basin
  77. Structural detachment influences the shale gas preservation in the Wufeng-Longmaxi Formation, Northern Guizhou Province
  78. Distribution law of Chang 7 Member tight oil in the western Ordos Basin based on geological, logging and numerical simulation techniques
  79. Evaluation of alteration in the geothermal province west of Cappadocia, Türkiye: Mineralogical, petrographical, geochemical, and remote sensing data
  80. Numerical modeling of site response at large strains with simplified nonlinear models: Application to Lotung seismic array
  81. Quantitative characterization of granite failure intensity under dynamic disturbance from energy standpoint
  82. Characteristics of debris flow dynamics and prediction of the hazardous area in Bangou Village, Yanqing District, Beijing, China
  83. Rockfall mapping and susceptibility evaluation based on UAV high-resolution imagery and support vector machine method
  84. Statistical comparison analysis of different real-time kinematic methods for the development of photogrammetric products: CORS-RTK, CORS-RTK + PPK, RTK-DRTK2, and RTK + DRTK2 + GCP
  85. Hydrogeological mapping of fracture networks using earth observation data to improve rainfall–runoff modeling in arid mountains, Saudi Arabia
  86. Petrography and geochemistry of pegmatite and leucogranite of Ntega-Marangara area, Burundi, in relation to rare metal mineralisation
  87. Prediction of formation fracture pressure based on reinforcement learning and XGBoost
  88. Hazard zonation for potential earthquake-induced landslide in the eastern East Kunlun fault zone
  89. Monitoring water infiltration in multiple layers of sandstone coal mining model with cracks using ERT
  90. Study of the patterns of ice lake variation and the factors influencing these changes in the western Nyingchi area
  91. Productive conservation at the landslide prone area under the threat of rapid land cover changes
  92. Sedimentary processes and patterns in deposits corresponding to freshwater lake-facies of hyperpycnal flow – An experimental study based on flume depositional simulations
  93. Study on time-dependent injectability evaluation of mudstone considering the self-healing effect
  94. Detection of objects with diverse geometric shapes in GPR images using deep-learning methods
  95. Behavior of trace metals in sedimentary cores from marine and lacustrine environments in Algeria
  96. Spatiotemporal variation pattern and spatial coupling relationship between NDVI and LST in Mu Us Sandy Land
  97. Formation mechanism and oil-bearing properties of gravity flow sand body of Chang 63 sub-member of Yanchang Formation in Huaqing area, Ordos Basin
  98. Diagenesis of marine-continental transitional shale from the Upper Permian Longtan Formation in southern Sichuan Basin, China
  99. Vertical high-velocity structures and seismic activity in western Shandong Rise, China: Case study inspired by double-difference seismic tomography
  100. Spatial coupling relationship between metamorphic core complex and gold deposits: Constraints from geophysical electromagnetics
  101. Disparities in the geospatial allocation of public facilities from the perspective of living circles
  102. Research on spatial correlation structure of war heritage based on field theory. A case study of Jinzhai County, China
  103. Formation mechanisms of Qiaoba-Zhongdu Danxia landforms in southwestern Sichuan Province, China
  104. Magnetic data interpretation: Implication for structure and hydrocarbon potentiality at Delta Wadi Diit, Southeastern Egypt
  105. Deeply buried clastic rock diagenesis evolution mechanism of Dongdaohaizi sag in the center of Junggar fault basin, Northwest China
  106. Application of LS-RAPID to simulate the motion of two contrasting landslides triggered by earthquakes
  107. The new insight of tectonic setting in Sunda–Banda transition zone using tomography seismic. Case study: 7.1 M deep earthquake 29 August 2023
  108. The critical role of c and φ in ensuring stability: A study on rockfill dams
  109. Evidence of late quaternary activity of the Weining-Shuicheng Fault in Guizhou, China
  110. Extreme hydroclimatic events and response of vegetation in the eastern QTP since 10 ka
  111. Spatial–temporal effect of sea–land gradient on landscape pattern and ecological risk in the coastal zone: A case study of Dalian City
  112. Study on the influence mechanism of land use on carbon storage under multiple scenarios: A case study of Wenzhou
  113. A new method for identifying reservoir fluid properties based on well logging data: A case study from PL block of Bohai Bay Basin, North China
  114. Comparison between thermal models across the Middle Magdalena Valley, Eastern Cordillera, and Eastern Llanos basins in Colombia
  115. Mineralogical and elemental analysis of Kazakh coals from three mines: Preliminary insights from mode of occurrence to environmental impacts
  116. Chlorite-induced porosity evolution in multi-source tight sandstone reservoirs: A case study of the Shaximiao Formation in western Sichuan Basin
  117. Predicting stability factors for rotational failures in earth slopes and embankments using artificial intelligence techniques
  118. Origin of Late Cretaceous A-type granitoids in South China: Response to the rollback and retreat of the Paleo-Pacific plate
  119. Modification of dolomitization on reservoir spaces in reef–shoal complex: A case study of Permian Changxing Formation, Sichuan Basin, SW China
  120. Geological characteristics of the Daduhe gold belt, western Sichuan, China: Implications for exploration
  121. Rock physics model for deep coal-bed methane reservoir based on equivalent medium theory: A case study of Carboniferous-Permian in Eastern Ordos Basin
  122. Enhancing the total-field magnetic anomaly using the normalized source strength
  123. Shear wave velocity profiling of Riyadh City, Saudi Arabia, utilizing the multi-channel analysis of surface waves method
  124. Effect of coal facies on pore structure heterogeneity of coal measures: Quantitative characterization and comparative study
  125. Inversion method of organic matter content of different types of soils in black soil area based on hyperspectral indices
  126. Detection of seepage zones in artificial levees: A case study at the Körös River, Hungary
  127. Tight sandstone fluid detection technology based on multi-wave seismic data
  128. Characteristics and control techniques of soft rock tunnel lining cracks in high geo-stress environments: Case study of Wushaoling tunnel group
  129. Influence of pore structure characteristics on the Permian Shan-1 reservoir in Longdong, Southwest Ordos Basin, China
  130. Study on sedimentary model of Shanxi Formation – Lower Shihezi Formation in Da 17 well area of Daniudi gas field, Ordos Basin
  131. Multi-scenario territorial spatial simulation and dynamic changes: A case study of Jilin Province in China from 1985 to 2030
  132. Review Articles
  133. Major ascidian species with negative impacts on bivalve aquaculture: Current knowledge and future research aims
  134. Prediction and assessment of meteorological drought in southwest China using long short-term memory model
  135. Communication
  136. Essential questions in earth and geosciences according to large language models
  137. Erratum
  138. Erratum to “Random forest and artificial neural network-based tsunami forests classification using data fusion of Sentinel-2 and Airbus Vision-1 satellites: A case study of Garhi Chandan, Pakistan”
  139. Special Issue: Natural Resources and Environmental Risks: Towards a Sustainable Future - Part I
  140. Spatial-temporal and trend analysis of traffic accidents in AP Vojvodina (North Serbia)
  141. Exploring environmental awareness, knowledge, and safety: A comparative study among students in Montenegro and North Macedonia
  142. Determinants influencing tourists’ willingness to visit Türkiye – Impact of earthquake hazards on Serbian visitors’ preferences
  143. Application of remote sensing in monitoring land degradation: A case study of Stanari municipality (Bosnia and Herzegovina)
  144. Optimizing agricultural land use: A GIS-based assessment of suitability in the Sana River Basin, Bosnia and Herzegovina
  145. Assessing risk-prone areas in the Kratovska Reka catchment (North Macedonia) by integrating advanced geospatial analytics and flash flood potential index
  146. Analysis of the intensity of erosive processes and state of vegetation cover in the zone of influence of the Kolubara Mining Basin
  147. GIS-based spatial modeling of landslide susceptibility using BWM-LSI: A case study – city of Smederevo (Serbia)
  148. Geospatial modeling of wildfire susceptibility on a national scale in Montenegro: A comparative evaluation of F-AHP and FR methodologies
  149. Geosite assessment as the first step for the development of canyoning activities in North Montenegro
  150. Urban geoheritage and degradation risk assessment of the Sokograd fortress (Sokobanja, Eastern Serbia)
  151. Multi-hazard modeling of erosion and landslide susceptibility at the national scale in the example of North Macedonia
  152. Understanding seismic hazard resilience in Montenegro: A qualitative analysis of community preparedness and response capabilities
  153. Forest soil CO2 emission in Quercus robur level II monitoring site
  154. Characterization of glomalin proteins in soil: A potential indicator of erosion intensity
  155. Power of Terroir: Case study of Grašac at the Fruška Gora wine region (North Serbia)
  156. Special Issue: Geospatial and Environmental Dynamics - Part I
  157. Qualitative insights into cultural heritage protection in Serbia: Addressing legal and institutional gaps for disaster risk resilience
Downloaded on 8.9.2025 from https://www.degruyterbrill.com/document/doi/10.1515/geo-2022-0627/html
Scroll to top button