You are on page 1of 13

3D Numerical Analysis of a Single Footing on Soft Soil

Reinforced by Rigid Inclusions


Jiamin Zhang1; Orianne Jenck2; and Daniel Dias3

Abstract: In the absence of high-quality soils, the rigid inclusion reinforcement technique has been extensively used in geotechnical
Downloaded from ascelibrary.org by University of Central Florida on 03/10/24. Copyright ASCE. For personal use only; all rights reserved.

engineering practice to meet the serviceability requirements of construction projects. This article aims to investigate the behavior of a single
footing placed on soft soil reinforced by four rigid inclusions using three-dimensional (3D) finite difference modeling (FDM). A set of full-
scale experimental tests are used as references for the development of the numerical approach. The concurrence of the numerical and
experimental data allows us to assess the load transfer platform thickness influence on the system under centered and eccentric loading con-
ditions. Some preliminary insights can thus be obtained based on this analysis for the soft soil improvement using rigid inclusions at the
design stage. DOI: 10.1061/(ASCE)GM.1943-5622.0002412. © 2022 American Society of Civil Engineers.
Author keywords: Numerical modeling; Soil/structure interaction; Soft soil; Soil improvement; Rigid inclusion; Footing.

Introduction In the case of a piled embankment, the load is transmitted to the in-
clusions mainly through the granular layer by shearing mecha-
Soft soils (clay, silt, or peat), as a well-known problematic soil nisms. The thickness of LTP and/or embankment is large enough
category, have some typical characteristics: low bearing capacity, to meet an equal settlements plan development. While in the pres-
high compressibility, and long consolidation times. This type of ence of a slab or a raft because the structure also contributes to the
soil is widespread in certain regions (notably coastal regions) and load transfer, the LTP thickness in this type of work is generally
needs to be enhanced to meet the design requirements. The rigid small compared with the inclusions spacing (Simon 2012;
inclusion (RI) technique, referring to both pile-supported embank- Messioud et al. 2016, 2017; Houda et al. 2021).
ment (with or without geosynthetic) and pile-supported earth plat- Different methodologies were developed for the design of RI
form (without geosynthetic reinforcement), was adopted to address systems considering the load transfer mechanism through soil arch-
soft-soil-related problems as a cost and time-consuming effective ing under static loadings (Terzaghi 1943; Guido et al. 1987; Com-
solution since the late 1970s. It combines a set of regularly arranged barieu 1988; Kempfert et al. 1997; Russell and Pierpoint 1997;
rigid inclusions floating, placed, or anchored in the support layer Svanø et al. 2000; Collin 2004; Filz 2006; van Eekelen and Bezui-
according to the soft soil layer thickness, with a load transfer plat- jen 2008; Zhang et al. 2016). Some comparisons demonstrate a
form (LTP) made of granular material connecting the foundation good agreement between the analytical models and the real projects
(raft or footing) and the inclusions. (Xing et al. 2014; Ariyarathne and Liyanapathirana 2015), different
At the base of the load transfer platform, the stiffness difference results are noted for the same situation using various design meth-
between the inclusions and soft soil causes differential settlements, ods (Bhasi and Rajagopal 2015), and neither of them considers the
which leads to shearing mechanisms and arching effects within the influence of complex loading (e.g., eccentric loading). The need of
granular layer. The arching transmits part of the loads to the inclu- using experimental and numerical approaches is thus emphasized
sions and reduces the global settlements as well as the loads sus- to better investigate the load transfer mechanism within the RI sys-
tained by the soft ground (Simon and Schlosser 2006; Briançon tem. Numerous studies concerning the behavior of piled embank-
et al. 2015). The skin friction along inclusions is also an instructive ments under static loading were carried out in the past (Han and
component for load transfer in the system (Okyay and Dias 2010). Gabr 2002; Jenck et al. 2009; Van Eekelen et al. 2012; Zhuang
In general, depending on whether there is a concrete slab (or et al. 2012; Nunez et al. 2013; Wang et al. 2018; Pham et al.
raft) above the LTP or not, two types of RI works can be derived. 2019; Tran et al. 2021).
However, the RI reinforcement which combines structural ele-
1 ments such as a rigid slab or a raft has not been fully addressed.
Ph.D. Candidate, 3SR Laboratory, Grenoble INP, CNRS, Univ.
Grenoble Alpes, F-38000 Grenoble, France. ORCID: https://orcid.org Raithel et al. (2008) took consideration of using a concrete slab
/0000-0003-2151-6551. Email: jiamin.zhang@3sr-grenoble.fr placed directly on the top of piles, as one of the solutions to guar-
2
Professor, 3SR Laboratory, Grenoble INP, CNRS, Univ. Grenoble antee a sufficient load transfer and distribution of pile-supported
Alpes, F-38000 Grenoble, France. ORCID: https://orcid.org/0000-0002 embankments (pile-slab–supported embankment) based on experi-
-2623-1830. Email: orianne.jenck@3sr-grenoble.fr mental results of a railway line. They showed that the embankment
3
Professor, 3SR Laboratory, Grenoble INP, CNRS, Univ. Grenoble settlement using a slab is two-thirds compared with the solution of
Alpes, F-38000 Grenoble, France; School of Automotive and Transporta- using a geosynthetic reinforcement. Jiang et al. (2014) investigated
tion Engineering, Hefei Univ. of Technology, Hefei 230009, China the performance of a pile-slab–supported railway embankment
(corresponding author). Email: daniel.dias@3sr-grenoble.fr
Note. This manuscript was submitted on July 26, 2021; approved on
using a two-dimensional (2D) hydro-mechanical finite-element nu-
February 7, 2022; published online on May 26, 2022. Discussion period merical model. The effects of combining the rigid piles and the re-
open until October 26, 2022; separate discussions must be submitted for in- inforced concrete slab are shown by the significantly reduced
dividual papers. This paper is part of the International Journal of Geome- vertical stresses applied on to the soil between piles. Besides, the
chanics, © ASCE, ISSN 1532-3641. shear forces and bending moments in the concrete slab increase

© ASCE 04022113-1 Int. J. Geomech.

Int. J. Geomech., 2022, 22(8): 04022113


with the slab thickness increase. Chevalier et al. (2011) modeled a Experimental Project Description
case of soft soil reinforcement with rigid inclusions under a slab
placed on top of a granular platform, using a discrete element
Site Characteristics
method. In the presence of the slab, a high concentration of vertical
stresses on the slab over the piles is observed. The influence of the The full-scale experimental tests site is located in France and is re-
thickness and frictional characteristics of the granular layer on the ported and detailed by Baroni et al. (2016). The ground surface is
system performance is underlined as well. Briançon et al. (2015) covered with a 1-m-thick in-place fill layer, followed by a
realized a three-dimensional (3D) finite-difference back-analysis 4-m-thick silty clay layer lying on the substratum made of hard
for a rigid inclusion-reinforced industrial building using a concrete gravel and sand. The groundwater table is around 3.7 m deep. In
raft as a foundation. The simulations identified the efficiency of order to evaluate the main geotechnical parameters of the soil lay-
using a unit cell model and were in good agreement with experi- ers, site investigations were carried out including cone penetration
mental data for the load transfer as well as the raft differential set- test and pressuremeter tests, as well as laboratory tests. Represen-
tlement. A 2D axisymmetric finite-element analysis was conducted tative values were selected for different soils assuming that each
Downloaded from ascelibrary.org by University of Central Florida on 03/10/24. Copyright ASCE. For personal use only; all rights reserved.

by Dias and Grippon (2017) for a pile-supported embankment layer is homogeneous (Table 1).
using variable inertia piles in the case of industrial slabs. Compared
with classical circular piles, the negative friction is not fully devel-
oped at the pile head and the effort along the pile shaft is redistrib- Loading Program
uted; the bending moment resulting in the pavement is more Two concrete square footings of 2-m long (B) and 0.5-m thick were
important. Rui et al. (2020) performed a set of 2D multitrapdoor set up over the ground surface, one without reinforcement and the
model tests to investigate the developed load transfer mechanisms other is reinforced by four rigid inclusions arranged in a square
within the platform layer between the rigid raft and column founda- grid. The inclusions are 5.5-m long with 0.5-m embedded in the
tion. According to the deformation characteristics observed in these substratum layer. The diameter d of the inclusions is 0.32 m, the in-
tests, a method for evaluating the load efficiency based on the clusion spacing s is 1.2 m, and their Young’s modulus is 5 GPa
pressure diffusion angle on the columns was presented. The load with a compressive strength of 10 MPa (for 28 days). A thin
efficiency decreases with an increase of the relative LTP thickness layer of concrete lean is placed between the footing and the inclu-
and the relative column net spacing. sions with Young’s modulus of 2 GPa in the construction to pro-
Studies of the RI system behaviors for the cases under a foot- vide a uniform surface for the foundation.
ing are more limited. Dias and Simon (2015) carried out a 3D As shown in Table 3, three types of loading tests were
finite-difference model to simulate a single spread footing over performed:
soil reinforced by rigid inclusions under inclined loading condi- 1. A classical vertical loading test on a single rigid inclusion
tions. The numerical results were compared with the simplified [Fig. 1(a)], set up with a loading–unloading cycle of 300 kN
method proposed by Simon (2010) and showed a fair agreement (Test 1A) followed by a loading–unloading cycle of 480 kN
for the investigated cases. Al-Naddaf et al. (2019) carried out (Test 2A). This test was performed on two rigid inclusions
trapdoor model tests under plane-strain conditions to investigate under the same condition, to assess the test repeatability.
the effects of static surface footing loading on soil arching mech- 2. A centered vertical loading test on a footing without reinforce-
anism in geosynthetic-reinforced and unreinforced embank- ment [Fig. 1(b)], carried out with two successive loading–un-
ments. As the load applied to the rigid footing increases, a loading cycles of 400 kN (Test 1B) and 1,800 kN (Test 2B),
narrow contact zone is created between the soil wedge below respectively.
the footing due to nonuniform contact pressure underneath the 3. A centered vertical loading test on a footing reinforced by four
footing and the soil wedge above the trapdoor. It is caused by rigid inclusions [Fig. 1(c)], with two successive loading–un-
soil arching and generates a higher pressure on the trapdoor by loading cycles of 1,000 kN (Tests 1C and 2C).
transferring the load downward. Zhang et al. (2021a, 2021b) For each loading path, the applied load is divided into five
conducted 2D trapdoor tests with transparent soil to evaluate phases, which account for 20%, 40%, 60%, 80%, and 100% of
the soil arching phenomenon under static and cyclic surface foot- the total load. For the unloading stage, the applied load is fir re-
ing loading. It was shown that as the loading increased, the soil duced to 60% and then reduced to zero. The load duration of
arching degraded locally at its crown and then expanded to both each phase is 4 h to reach a stable system state and check its robust-
sides of the arch till full degradation was reached. ness under different loads.
The literature indicates that experimental and numerical studies The instrumentation used consists of pressure cells and dis-
are scarce and therefore knowledge about the behavior of rigid placement transducers that allow measuring the stress applied
inclusion-reinforced structures for the cases with the presence of on the inclusion heads and the vertical displacements of the foot-
a footing, especially under eccentric loading. Consequently, a full- ing center.
scale experimental test considering rigid inclusions under a single
footing has been carried out to perform studies under vertical cen-
tered/eccentric loading (Baroni et al. 2016). The case without any
Table 1. Geotechnical parameters obtained from in situ and laboratory
load transfer platform has been investigated. Based on this exper- tests
imental database, a numerical approach is developed in this cur-
rent work, using a finite-difference code, in order to study the γ
system’s behavior under various conditions. After validation of Depth (kN/ Em pL α qc c′ φ′
Soil type (m) m3) (MPa) (MPa) (—) (MPa) (kPa) (°)
the experimental data, a parametric study is thus conducted to an-
alyze the impact of the load transfer platform thickness and the In-place fill 0–1 20 15 0.96 1/2 5 11 25
surface loading eccentricity. The result analysis mainly focuses Silt clay 1–5 20 5.5 0.43 1/2 1 6 31
on the footing settlements, stress distribution, and solicitations Substratum 5 16.5 35 2.50 1/3 20 0 43
generated in the RI. Source: Data from Baroni et al. (2016).

© ASCE 04022113-2 Int. J. Geomech.

Int. J. Geomech., 2022, 22(8): 04022113


Downloaded from ascelibrary.org by University of Central Florida on 03/10/24. Copyright ASCE. For personal use only; all rights reserved.

(a) (b) (c)

Fig. 1. Experimental loading tests arrangement: (a) Tests 1A and 2A; (b) Tests 1B and 2B; and (c) Tests 1C and 2C.

Numerical Simulation considering the Mohr–Coulomb failure criteria (termed MC). The
footing, the inclusions, and the concrete lean are modeled with vol-
ume elements as linear elastic materials (termed EL). The soft soil
Model Properties
layers are simulated using a hardening/softening elastoplastic con-
Several numerical models were developed. Due to the model sym- stitutive model, that is, the modified Cam-clay model (named
metry, a quarter of the inclusion is simulated for the loading tests of MCC). It allows considering nonlinear elasticity and a hardening/
single rigid inclusions, while half of the footing is modeled for the softening behavior governed by the volumetric plastic strains.
loading tests of unreinforced and reinforced footings because it will The shear flow rule is associated (Itasca 2017). An oedometer
be also used for the eccentric loading analyses afterward. The test was available for the model parameter characterization (sample
finite-difference program FLAC 3D (Version 6.0) was used for taken at the soft soil layer). As shown in Fig. 3, a fair agreement of
the numerical calculations. The model geometries and the mesh the numerical and experimental results is observed by adjusting the
distributions adopted are shown in Fig. 2 are based on a mesh sen- proper compressibility and swelling line index (λ and κ parame-
sitivity analysis. The model comprises 46,686 elements and 51,438 ters). Because the preconsolidation pressure varies along the soil
gridpoints for the single inclusion; 140,480 elements and 151,487 depth, the calibration of this value is completed using the loading
gridpoints for the footing case. Mesh refinement is assigned to the tests on the unreinforced footing (experimental Test 2B). A value
zone close to the rigid inclusions and the fill layer to achieve greater of pc0 = 315 kPa is adopted because it permits to obtain the same
accuracy. In a preliminary study, it was also found that to eliminate footing settlement as in the experiment [Fig. 4(b)].
the boundary effects, using a model length in the horizontal direc- Interface elements are placed on the outer perimeter of the foot-
tion 9.6 times the inclusion spacing and extending the model to a ing and all the outer inclusion faces of the inclusions (heads/tips/
depth of 12 m are sufficient conditions. The displacements at the along the shafts). They are implemented to model the interaction
bottom boundary are blocked in all directions (x, y, z) and the ver- between the soil and the inclusions, with an interface friction
tical boundaries for the horizontal displacements. The phreatic angle set equal to two-thirds of the surrounding soil value. The
level is set at 3.7-m depth as observed in situ; all the calculations shear and normal interface stiffnesses were calculated depending
are carried out in drained conditions as the saturated soft soil on the elastic properties of the neighboring material and adjacent
layer is relatively thin and lies on a permeable stratum, so the mesh zone dimensions by using the FLAC3D user’s manual to
pore pressure increase phase in response to the loading conditions limit time calculation (Itasca 2013).
is neglected in the calculation. Weightless beam elements were added at the inclusion axis with
For the simulation of each loading test, four calculation phases a reduced Young’s modulus compared with the concrete original
are carried out and the model equilibrium is reached at each phase: value (10,000 times lower). This very small added stiffness will
• Initial phase: setup of the soil layers without the inclusions. The not cause influence on the system stiffness and allows the inclusion
initial soil stress state is calculated using a nonlinear initial K0 bending moments and internal forces to be determined easily.
profile based on Mayne’s formula considering K0 = K0_NC
OCR0.5 (Mayne and Kulhawy 1982). A constant preconsolida-
Comparison with the Experimental Results
tion stress is considered (explained in the following paragraph).
• Setup of the rigid inclusions, without consideration of the instal- Numerical calculations were successively performed for the three
lation effects. loading tests (A, B, and C) described previously. The two cycles
• Setup of the concrete footing and underlying lean. of the vertical loading tests on single rigid inclusions are named
• Application of the vertical additional surcharge, by applying an Tests 1A and 2A, and the two different rigid inclusions are called
evenly distributed vertical stress in a circular area with a 1-m di- RI-A and RI-B, respectively. As shown in Fig. 5, the load–dis-
ameter on the top of the footing, with an increasing magnitude. placement curves indicate a quasi-linear-elastic behavior during
The system equilibrium is reached at each loading increment. the first loading circle for both RIs [Fig. 5(a)]. This is not the
The geotechnical parameters used for the soils and structural el- case for the two experimental results in which initial stiffness is
ements are presented in Table 2. The fill and substratum are sup- much higher than in the numerical model. As the preconsolidation
posed to behave like a linear elastic-perfectly plastic material stress value is based on experimental data, the K0 distribution on

© ASCE 04022113-3 Int. J. Geomech.

Int. J. Geomech., 2022, 22(8): 04022113


Downloaded from ascelibrary.org by University of Central Florida on 03/10/24. Copyright ASCE. For personal use only; all rights reserved.

(a) (b)

Fig. 2. Schematic view of the 3D numerical models and mesh distributions: (a) Test A; and (b) Tests B and C.

Table 2. Geomechanical parameters for fill and substratum layers and structural elements
Material Constitutive model E (MPa) ν φ′ (°) c′ (kPa) ψ (°) γ (kN/m3) λ κ M pc0 (kPa) vλ
Fill MC 30 0.3 25 11 0 19 — — — — —
Silty clay MCC — 0.3 — — — 18 0.06 0.007 1.244 315 1.89
Substratum MC 350 0.3 38 250 10 19 — — — — —
Inclusion EL 5 × 103 0.2 — — — 24 — — — — —
Footing EL 24 × 103 0.2 — — — 25 — — — — —
Lean EL 2 × 103 0.2 — — — 22 — — — — —

Table 3. Summary of the loading program


Maximum surface
Test no. Model condition loading level (kN)
1A Single rigid inclusion (RI-A and RI-B) 300
2A Single rigid inclusion (RI-A and RI-B) 480
1B Footing without reinforcement 400
2B Footing without reinforcement 1,800
1C Footing reinforced by four rigid inclusions 1,000
2C Footing reinforced by four rigid inclusions 1,000

the Mayne’s formula, the main probable reason of this mismatch is


the fact that the considered constitutive model is not able to con-
sider the shear modulus degradation curve with the shear strains.
When the vertical load exceeds 350 kN during the second loading
cycle, the experimental tests for the two RIs lead to different results
in terms of displacements [Fig. 5(b)]. From this value, a nonlinear
behavior is observed for RI-A, also noticeable considering the un- Fig. 3. Calibration of the oedometer test results with the modified
loading path with a residual settlement of 9.8 mm. The numerical Cam-clay model.
results are similar to those obtained for the two RIs when their de-
formation remains in the elastic range: The curve correlates well
with the measurements carried out on the two RIs during Test elastically, because the residual settlement of the footing is only
1A, then the curve agrees only with RI-B test result between 350 1 mm [Fig. 4(a)]. For Test 2B, the hyperbolic shape of the experi-
and 480 kN with a maximum difference of 1.2 mm during Test 2A. mental results and the significant 40-mm residual settlement after a
The results of the vertical loading test on the footing over the load of 1,800 kN indicate a nonlinear and plastic behavior of the
unreinforced soil are given in Fig. 4 (Test 1B for the first cycle soil under large loading conditions [Fig. 4(b)]. The numerical dis-
and Test 2B for the second cycle), in terms of footing center dis- placements are overpredicted at the beginning of the loading pro-
placements according to the applied vertical loading. During the cess for both tests, whereas the numerical and experimental
1B test (loading up to 400 kN), in both the experimental test and results start to be in better agreement for larger loadings (after
numerical model, the system behaves linearly and almost 1,400 kN).

© ASCE 04022113-4 Int. J. Geomech.

Int. J. Geomech., 2022, 22(8): 04022113


Downloaded from ascelibrary.org by University of Central Florida on 03/10/24. Copyright ASCE. For personal use only; all rights reserved.

(a)
(a)

(b)
(b)
Fig. 4. Experimental and numerical footing center displacements:
(a) Test 1B; and (b) Test 2B. Fig. 5. Experimental and numerical inclusion head displacement in
single rigid inclusion loading tests: (a) Test 1A; and (b) Test 2A.

In fact, during numerical Test 1B, the soft soil layer remains in
the overconsolidated zone owing to the high preconsolidation pres- footing is placed directly on top of them (cf. value of EL parameter
sure, and the shallow soil layers (fill and silty clay layer) behave in defined afterward). Moreover, the high soft soil preconsolidation
their linear-elastic domain. When the applied load becomes larger pressure of 315 kPa makes the soft soil behavior remain in the
(between 1,200 and 1,400 kN), the nonlinear behavior of the overconsolidated zone, where no plasticity occurs under compres-
MCC constitutive model is mobilized in the silty clay layer and sion. The numerical system thus stays in the elastic range, hence the
shear failure zones also appear in the fill. This actually results in residual settlements observed in the experimental test are not
plastic settlements. represented in the numerical simulation.
The same numerical modeling procedure and model properties as Meanwhile, a new parameter was defined: The settlement
efficiency E s , which denotes the reduction ratio of the footing
for the two previous cases were implemented to simulate the case of
settlement with reinforcement u and the footing settlement
the footing on soil improved with the four inclusions (Test 1C for the
without reinforcement u un under the same loading condition.
first cycle and Test 2C for the second cycle). As depicted in Fig. 6(a),
For a load of 1,000 kN, E s equals to 0.75, demonstrating a
during experimental Test 1C (first loading cycle), the vertical footing
great settlement reduction by combining the footing with the
displacement increases linearly with the applied vertical loading,
rigid inclusions.
reaches 5 mm for a load of 1,000 kN, and a residual settlement of
the footing of 2 mm is recorded after unloading. Experimental Test u
2C (second loading cycle) then shows a linear-elastic behavior. Es = 1 − (1)
uun
The numerical model succeeds in producing the footing center dis-
placement of the first loading path but does not vary from the first The average vertical stress on the inclusion head is obtained by
loading cycle to the second one, contrary to the experimental obser- dividing the sum of the vertical forces on the inclusion-head-level
vations. Additionally, the numerical simulation shows a roughly uni- elements by the surface of the inclusion cross section. Fig. 6(b)
form distribution of settlements along the footing base. shows the average stress extracted from the central area of the in-
In the system, a large part of the load applied on the footing is clusion, on a 0.16-m-diameter disk (whereas the inclusions have
transferred to the substratum through the inclusions, because the a 0.32-m diameter), taking into account the fact that the pressure

© ASCE 04022113-5 Int. J. Geomech.

Int. J. Geomech., 2022, 22(8): 04022113


Downloaded from ascelibrary.org by University of Central Florida on 03/10/24. Copyright ASCE. For personal use only; all rights reserved.

(a)
Fig. 7. Distribution of the vertical stress on the left inclusion head pres-
sure at different loading levels.

(b)

Fig. 6. Experimental and numerical footing displacements: (a) and av-


(a)
erage vertical stress of the left-side inclusion central area; and (b) for
Tests 1C and 2C.

sensor is not large enough to cover the whole inclusion head and
thus records the stress concentration effect on the inclusion edges,
as highlighted in Fig. 7. Only the results of one inclusion are pre-
sented here because the four inclusions bear the same force under
the centered loading condition. As the system stays in a linear-elastic
range, the measured stress increases linearly with the applied load.
The slopes of the unloading–reloading lines of the numerical analy-
sis are almost parallel to those of the experiment.
The global load efficiency EL is determined as the ratio of the
load transmitted to the four inclusions Qp to the sum of external
load applied on the footing Q and the weight of footing combined
with the weight of the LTP layer below the footing, W (in this case,
the weight of the LTP layer is null). The numerical load transfer ef-
ficiency is equal to 80% under an external load Q = 1,000 kN. This (b)
important value is due to the fact that the rigid footing is directly
placed above the inclusions, that is, without possible distribution Fig. 8. (a) Influence of the LTP thickness on the settlement efficiency;
in the load transfer platform. and (b) the average stress efficiency.

Qp
EL = (2)
Q+W
observed at the inclusion edge is about 3.5 times the stress at the
Fig. 7 depicts the stress distribution at the inclusion head level, inclusion center: considering all the volume elements of the inclu-
along a cross line A-A′ through the left side inclusion head center, sion head, the average pressure is 2,600 kPa for Q = 1,000 kN,
at two different loading levels. This figure shows that the stress whereas it is only 1,600 kPa when considering the inclusion central

© ASCE 04022113-6 Int. J. Geomech.

Int. J. Geomech., 2022, 22(8): 04022113


Table 4. Corresponding H/s and H/B ratios of different LTP thicknesses
H (m) H/s (inclusion spacing s = 1.2 m) H/B (footing length B = 2 m)
0 0 0
0.25 0.2 0.125
0.5 0.42 0.25
1 0.83 0.5
2.5 2.1 1.25
Downloaded from ascelibrary.org by University of Central Florida on 03/10/24. Copyright ASCE. For personal use only; all rights reserved.

Fig. 10. Influence of the LTP thickness on bending moments (Mfy) of


the left inclusion under a centered load of 1,000 kN.

Parametric Analysis

This section studies the thickness effect of the load transfer plat-
(a) form on the behavior of a footing placed over soil reinforced
with rigid inclusions under vertical centered and eccentric loading
conditions. The discussion mainly focuses on the analysis of sur-
face displacements of the footing, stress-related results, and solici-
tation generated in the rigid inclusions.
The presence of an LTP may significantly affect the reinforce-
ment performance because the arching effect is developed within
this LTP and leads to a load distribution between the inclusions
and the soft soil (Jenck et al. 2009). Based on the reference case
model configuration, five different thicknesses of the LTP layer
were considered between the footing and the inclusion head to
study the LTP thickness impact H: 0 m (the reference case, without
LTP), 0.25, 0.5, 1, and 2.5 m. The case with the larger value H =
2.5 m is not representative of a real case but is investigated as an ex-
treme case. The corresponding H/s and H/B ratios are presented in
Table 4, with s being the inclusion spacing, B being the side length
of the square footing. The LTP is made of the same material as the
(b) fill, thus has the same numerical constitutive model and model
parameters.
Fig. 9. Distribution of the vertical displacements: (a) the stress; and (b)
along AA′ at footing and inclusion head levels with an LTP thickness
of 0.5 m under a centered vertical load of 1,000 kN. Under Centered Vertical Loading
The values of settlement efficiency ES [Eq. (1)] and global load ef-
ficiency EL [Eq. (2)] according to the LTP thickness are shown in
part of diameter 0.16 m. It is also noticeable that the edge stress Fig. 8, at three centered vertical loading stages (Q up to 1,000 Kn).
concentration is larger on the inner side. As the LTP thickness increases, ES gradually decreases. As ex-
Results in terms of both footing vertical displacements and pres- pected, with a large LTP thickness of 2.5 m, there is no more ef-
sure carried by the inclusions show that the proposed numerical fect of adding reinforcement on the settlement reduction. This is
model permits rather satisfactorily simulating the main features because the LTP is thick compared with the footing size (H/B =
of the rigid inclusion reinforcement system behavior because it 1.25). The load applied to the footing is dispersed in this layer
can be observed under various experimental loading tests (from and only a partial loading is transmitted to the inclusion vicinity.
1A to 2C). However, the main limitation is that the numerical Thus, the RIs cannot contribute to the footing support. Addition-
model cannot correctly reproduce the effects of the cyclic loading, ally, Fig. 8(a) also shows that the settlement efficiency does al-
the system features in the elastic domain or due to the nonsimula- most not vary with the loading amplitude, whatever the LTP
tion of the soil’s small strain behavior. Thereby, the centered load- thickness H.
ing test of a footing over reinforced soil (Test 1C) is considered the The load efficiency also decreases with the increase in LTP
reference case in the following simulations. thickness, from 80% when H equals 0 m to less than 10% for

© ASCE 04022113-7 Int. J. Geomech.

Int. J. Geomech., 2022, 22(8): 04022113


H = 2.5 m. Indeed, when an LTP is added, the load is not directly inclination angle for an H of 2.5 m highlights the effective thickness
transferred from the rigid footing to the inclusions, but also through of the load transfer platform in RI reinforcement applications beyond
this platform, where shear mechanisms are developed. which the supporting effect of rigid inclusions almost disappears.
Fig. 9 depicts the distributions of the settlements and vertical Thereby, despite the centered vertical loading condition, bend-
stresses along a horizontal crossline A-A′ , at the inclusion head ing moments are generated in the inclusions under a footing. Due
and the footing base levels (i.e., base and top of the LTP). The to symmetry conditions, the values of the bending moments related
case of an LTP thickness H = 0.5 m under a load Q = 1,000 kN is to the y-axis and z-axis (defined in Fig. 10) are the same for each
selected as a representative case to highlight the system behavior inclusion, therefore only the bending moment for the left side inclu-
below the footing in the LTP presence. sion related to the y-axis (Mfy) is analyzed. The inclusion bending
A uniform distribution of settlements underneath the footing is moments are extracted from the beam elements placed in the
obtained due to its high rigidity. This footing settlement is equal to inclusion vertical axis. Fig. 10 shows the repartition of Mfy along
18 mm, that is, 3.6 times the settlement obtained for the reference the inclusion, for different LTP thicknesses, under a load
case (with H = 0 m). The settlement of the inclusions head is half Q = 1,000 kN. Whatever might be the LTP thickness, the bending
Downloaded from ascelibrary.org by University of Central Florida on 03/10/24. Copyright ASCE. For personal use only; all rights reserved.

of this value when H = 0 m. Simultaneously, a nonuniform distribu- moment is negligible in the lower part of the inclusions (from a
tion of the stress on the footing undersurface is observed [Fig. 9(b)]. depth equal to 3 m). Without LTP (H = 0 m), the maximum bend-
Stress concentrations occur at the inclusions’ edges and at the foot- ing moment is equal to 1.2 kN · m at a buried depth of 0.3 m. With
ing’s edges. This is due to the strong contrast between the soil and an LTP of 0.25 m, the maximum bending moment value becomes
structural components’ stiffnesses. Moreover, the inner side of the smaller and is moved down to a buried depth close to 1 m. When
inclusions is more loaded than the outer side: The inclusions are the thickness of the LTP is greater than 0.5 m, the bending moment
thus submitted to a nonaxial loading. The average vertical stress act- sign changes compared with a thin LTP, illustrating a change in the
ing on the inclusions is here equal to 950 kPa, which is 37% of the inclusion deformation. For H larger than 0.5 m, the maximum
value of 2,600 kPa obtained in the reference case H = 0 m, indicating bending moments are different in depth for different thicknesses,
the average load efficiency decreases with the LTP thickness in- but the trends of the three bending moment curves along the inclu-
crease. A larger amount of the load is transmitted to the soft soil be- sion are similar. The maximum value increases up to 3.3 kN · m for
tween the RIs when the LTP thickness increases. H = 1 m. The axial–flexural response of inclusions can then be an-
With the LTP presence and increase of its thickness, the inclina- alyzed by combining the effects of both the bending moments and
tion angle of the inclusion head plane increases. Table 5 gives this axial loads in the inclusions (Mirsepahi et al. 2021a, b). Assuming
inclination angle value for a surface load Q = 1,000 kN. The angle that the inclusions are perfectly elastic isotropic and have homog-
is defined as positive when the left-inclusion upper face is inclined enous sections, the axial normal stress σxx distribution in an inclu-
counterclockwise relatively to the horizontal plane (as observed in sion cross section is given by
Fig. 9(a), for H = 0.5 m). The inclination angle is negative when
the thickness of the LTP layer is below 0.25 m, because the inclusion N M fz M fy
σ xx = − ×y+ ×z (3)
head is relatively close to the footing center, with a centered loading, A Iz Iy
and is therefore inclined toward it. From an LTP thickness greater
than 0.5 m, the angle becomes positive and increases with the LTP where N = axial load of the inclusion cross section; A = inclusion
thickness increase until a certain thickness between 1 and 2.5 m. cross sectional area; Mfy and Mfz = bending moment related to the
Due to the LTP thickness increase, the load exerted on the footing y-direction and z-direction, respectively; Iy and Iz are the second
is no longer directly transmitted to the inclusions but through the moment of area of the inclusion’s cross section along the y-axis
LTP soil and generates an eccentric thrust, causing the inclusion and z-axis (Iy = Iz = 5.15e−4 m4), respectively; y and z are the mate-
head plane to be inclined outward the footing. The decreased rial extensions along the y-axis and z-axis, respectively (Fig. 10),
both varying from –r to r, with r being the inclusion radius
Table 5. Variation of the left inclusion head inclination angle with the LTP (r = 0.16 m). Table 6 completes the axial forces on the cross sec-
thickness under a centered load of 1,000 kN tions where the maximum bending moments occur (N ) and gives
H (m) Inclination angle (°) the corresponding extreme normal stresses in the inclusion section,
obtained
√on the inclusion√ external
 fibers on the diagonal, that is, for
0 —
y = ± 2/2r and z = ± 2/2r (as Mfy = Mfz). Positive values of
0.25 −0.025
0.5 0.06 σxx indicate tension. Table 6 highlights that tension is obtained
1 0.2 within the inclusions for H ≥ 1 m. In fact, as the LTP thickness in-
2.5 0.08 creases, the absolute value of the normal compressive force N de-
creases (which is consistent with the EL decrease), whereas the
bending moments display larger values. It should then be verified
that this tension in the inclusions would not lead to the use of rein-
forced concrete piles.

Table 6. Maximum bending moments, corresponding axial forces, and


extreme normal stresses in the cross sections, under a centered load
Q = 1,000 kN
H Depth of critical Mfy = Mfz N σxx min σxx max
(m) section (m) (kN · m) (kN) (kPa) (kPa)
0.25 −0.9 −0.8 −143 −2,142 −1,404
0.5 −1.7 1.3 −100 −1,807 −692
1 −1.4 3.3 −61 −2,211 691
Fig. 11. Load transfer platform layer in the numerical model. 2.5 −1.2 2.6 −31 −1,510 744

© ASCE 04022113-8 Int. J. Geomech.

Int. J. Geomech., 2022, 22(8): 04022113


Under Eccentric Loading
Eccentric loading conditions on a footing often occur in construc-
tion practice. This numerical work aims at investigating its impact
on the reinforcement system behavior. A vertical loading Q up to
1,000 kN is applied incrementally at two different eccentric posi-
tions on the footing top, located at e = 0.3 m or 0.45 m from the
footing center [Fig. 11, corresponding to eccentricity ratios (e/B)
of 0.15 and 0.225]. Calculations are performed to visualize the ef-
fect of the five different LTP thicknesses described previously on
the eccentric loading cases.
Fig. 12 compares the settlement of the footing center, for the
centered loading and different eccentricities, at two loading levels
Downloaded from ascelibrary.org by University of Central Florida on 03/10/24. Copyright ASCE. For personal use only; all rights reserved.

(Q = 600 then 1,000 kN). The eccentricity of the loading induces


an increase in the footing settlement, except for the case without
LTP. The footing settlement increases with eccentricity and is
higher for large LTP heights. This figure also shows that, for a
fixed e/B value, the footing settlement increases as the LTP thick-
ness increases (as previously observed for the centered loading). It Fig. 13. Load efficiency evolution for left and right rigid inclusions for
should be noted that with an eccentricity ratio of 0.225, the footing different external load eccentricities under a load magnitude of 600 kN.
is strongly inclined, and under a load that exceeds 800 kN, shear
failure occurs on the fill layer and the model cannot converge
when the LTP thickness is greater than 0.5 m.
The eccentric loading influence on the load efficiency is demon-
strated by the efficiency of each inclusion Ei, which is defined in
Eq. (4) as the ratio of the vertical load on the top of one inclusion
Qp_L or Qp_R (with the labels _L and _R referring to the left/right
sides) to a quarter of the sum of the total load applied on the footing
Q and the weight of footing combined with the weight of LTP layer
below the footing, W.
Q p L/R
Ei = (4)
(1/4)(Q + W )
In the centered loading case, the load efficiency for each inclu-
sion is obviously the same (equals to the global load efficiency EL).
Fig. 13 depicts the left and right load efficiency for Q = 600 kN
(representative for other loading magnitudes). As for the centered
loading case, the load efficiency of every single inclusion decreases
with the increase of the LTP thickness. In eccentric loading cases,
the difference between left and right load efficiency decreases as
the LTP thickness increases (and is null for H = 2.5 m).
(a)

(b)

Fig. 14. Distribution of the vertical displacement: (a) the stress; and
Fig. 12. Vertical displacements of the footing center under different (b) along AA′ at footing and inclusion head levels with an LTP thick-
eccentric loading conditions. ness of 0.5 m under an eccentric load of 1,000 kN (e/B = 0.15).

© ASCE 04022113-9 Int. J. Geomech.

Int. J. Geomech., 2022, 22(8): 04022113


Furthermore, this load efficiency difference between the left and
right inclusions is obviously higher for larger eccentricities.
The global load efficiency EL for the eccentric cases can also
be determined as the average between the left and right inclusion
efficiency. The results are presented in Fig. 13 and labeled with
_EL. The corresponding curves are all superimposed with the
e/B = 0 curve showing that the global load efficiency is not af-
fected by the eccentricity. However, the variation of the single in-
clusion load efficiency puts forward another requirement at the
design stage which is the strength of rigid inclusions, which
needs to be determined according to the possible eccentric loading
conditions.
Downloaded from ascelibrary.org by University of Central Florida on 03/10/24. Copyright ASCE. For personal use only; all rights reserved.

Fig. 14 depicts the distribution of the settlement and vertical


stress along the crossline A-A′ at different LTP elevations (base
and top), with an LTP thickness of 0.5 m and under an eccentric
load of 1,000 kN, with e/B = 0.15 (failure occurs with e/B =
0.025). Compared with the centered loading case, where the footing
settlement is equal to 18 mm, cf. Fig. 9(a), the footing center settle-
Fig. 15. Variation of the footing inclination angle with the LTP thick-
ment is increased by 44%. Under an eccentric loading, footing tilts
ness under an eccentric load of 1,000 kN (e/B = 0.15).
to the left with a linear settlements’ distribution. The maximum

(a) (b)

(c) (d)

Fig. 16. Influence of the LTP thickness on bending moments: Mfy in the left inclusion: (a) Mfy in the right inclusion; (b) Mfz in the left inclusion;
(c) Mfz in the right inclusion; and (d) under an eccentric load of 1,000 kN (e/B = 0.15).

© ASCE 04022113-10 Int. J. Geomech.

Int. J. Geomech., 2022, 22(8): 04022113


Table 7. Influence of the LTP thickness on axial forces and axial normal stresses of the cross sections with maximum bending moments under an eccentric
load of 1,000 kN (e/B = 0.15)
Left inclusion Right inclusion

Depth of Depth of
the critical Mfy Mfz σxx min σxx max critical Mfy Mfz σxx min σxx max
H (m) section (m) (kN · m) (kN · m) N (kN) (kPa) (kPa) section (m) (kN · m) (kN · m) N (kN) (kPa) (kPa)
0.25 −0.3 −0.3 −0.7 −189 −2,569 −2,131 −0.6 2.5 −0.8 −84 −1,858 −237
0.5 −1.4 3.2 4.0 −119 −3,063 94 −1.7 −4.8 0.8 −73 −2,410 584
1 −1.4 4.4 5.8 −71 −3,150 1,389 −1.4 −7.5 2.5 −53 −3,119 1,792
2.5 −1.2 2.1 3.1 −35 −1,610 748 −1.2 −3.1 −3.3 −29 −1,784 1,051
Downloaded from ascelibrary.org by University of Central Florida on 03/10/24. Copyright ASCE. For personal use only; all rights reserved.

settlement of the footing left end reaches 50 mm and is almost null inclusion sections, thus a special attention should be paid to the in-
on the right end. The settlements for all inclusions remain in a small clusion design in practical applications.
range (less than 4 mm), with slightly greater displacement on the
left compared with the right hand. In terms of the stress distribution,
the eccentric load increases the stress concentration of the left-side Conclusion
inclusion as well as on the left-side footing edge. The average ver-
tical stress carried by the left inclusion is equal to 1,210 kPa, which The mechanical behavior of a single footing placed on a soft soil
is about 2.4 times the average stress obtained on the right inclusion. layer reinforced by four rigid inclusions and subjected to centered
A peak of stress equal to 1,750 kPa is obtained on the inner side of and eccentric loading conditions has been investigated through a
the left inclusion edge. Compared with the centered loading case, finite-difference numerical approach. The numerical simulations
this peak value at the inclusion level increases by 40%, while the were developed according to three experimental loading tests.
peak value along the footing increases by 48%. The reliability of the developed numerical model was first assessed
Fig. 15 depicts the LTP thickness influence on the footing incli- through the comparison of the settlements and stresses database ob-
nation with e/B = 0.15 and Q = 1,000 kN. Results obtained without tained during the in situ experimental tests. A relatively good agree-
rigid inclusions improvement are also presented. As the LTP thick- ment between the numerical model and experimental data were
ness increases, the footing tilt increases. The figure also highlights obtained. Some mismatches were observed probably because of
that the footing inclination angle is smaller with rigid inclusions the considered constitutive model which is not able to reproduce
than without, especially for thin LTPs. This effect of reducing the the small strain behavior of soils (shear modulus degradation
footing tilt due to the RI gradually decreases when the LTP thickness curve with shear strains).
increases, and nearly disappears when H equals or exceeds 1 m. This Then, for valuing the influence of a load transfer platform and its
threshold value differs from the effective LTP thickness limit in- thickness on the reinforcement performance, together with the in-
ferred using footing settlement and load efficiency analysis. fluence of the surface load eccentricity, a numerical parametric
The bending moments are related to y- and z-axes (termed study was conducted. Relevant indicators such as the footing settle-
Mfy and Mfz) with an eccentric loading of e/B = 0.15 for the two- ment and inclination, the load distribution on the piles and under-
side inclusions with different LTP thickness under an external neath the footing, and the solicitations generated in the rigid
load Q = 1,000 kN are shown in Fig. 16. With the eccentric load- inclusions were discussed and interpreted.
ing, the bending moments Mfy and Mfz are not equal in each inclu- Regarding the use of a load transfer platform in a rigid inclusion
sion section, but the maximum absolute values are reached in the system under a single footing, the settlement efficiency and load ef-
same sections. Table 7 completes the axial force values obtained ficiency decrease with the platform thickness increase, under both
in the cross sections where these maximum bending moments centered and eccentric loading conditions. As the platform becomes
occur. Compared with the centered loading conditions (Table 6), sufficiently thick compared with the footing size (i.e., from H/B =
these axial forces N are greater (in absolute value) in the left inclu- 0.5–1.25 for both centered and eccentric loading conditions), the
sions and smaller in the right inclusions, as already shown by the inclusion reinforcement stops contributing to the load transfer
load efficiency analysis, while the maximum bending moment mechanisms.
magnitudes are generally higher than for the centered loading The loading eccentricity has a great impact on the system behav-
cases. The axial normal stresses in the inclusion cross section σxx ior because it results in the footing tilt and an increase of the inclu-
can be computed using Eq. (3), and the extreme values (minimum sion efforts. The single inclusion load efficiency variation becomes
and maximum) are indicated in Table 7. These values are obtained the main factor at the design stage instead of the global load effi-
on the inclusion external fiber, but not necessary on the inclusion ciency when considering an eccentric loading condition.
diagonal, as Mfy and Mfz values differ. Zones in tension (σxx The analysis of axial loads and bending moments within the
positive) are obtained in numerous situations: from H = 0.25 m rigid inclusions also indicates that tension stresses could be gener-
and for both left and right inclusions. Moreover, the values of the ated, for centered and eccentric loading cases, and for a sufficient
extreme cross-section normal stresses are higher (in absolute LTP thickness H (from H = 0.5–1 m for centered load; from H =
value, in both tension and compression) for the eccentric loading 0.25–0.5 m for eccentric load with e/B = 0.15). In this way, steel
than the centered loading. For instance, at Q = 1,000 kN and bars reinforcement could be required for the inclusions.
e/B = 0.15, the maximum tension stress is reached for H = 1 m in
the right inclusion, where σxx_max = 1,792 kPa, whereas the maxi-
mum tension in the inclusions for H = 1 m was 691 kPa for the cen- Data Availability Statement
tered load case.
The eccentric loading condition has a significant impact on the All data, models, and code generated or used during the study ap-
bending moments, axial loads, and thus normal stresses in the pear in the published article.

© ASCE 04022113-11 Int. J. Geomech.

Int. J. Geomech., 2022, 22(8): 04022113


Acknowledgments 145 (12): 04019114. https://doi.org/10.1061/(ASCE)GT.1943-5606
.0002190.
The authors gratefully acknowledge the financial support provided Ariyarathne, P., and D. S. Liyanapathirana. 2015. “Review of existing de-
by the China Scholarship Council (ID: 201908070075). The labo- sign methods for geosynthetic-reinforced pile-supported embank-
ments.” Soils Found. 55 (1): 17–34. https://doi.org/10.1016/j.sandf
ratory 3SR is part of the LabEx Tec 21 (Investissement d’avenir—
.2014.12.002.
Grant Agreement No. ANR-11-LABX-0030). Baroni, M., L. Briançon, J. Racinais, F. Maucotel, and H. Scache. 2016.
“Semelles Sur Inclusions Rigides: Validation Du Nouveau Cahier Des
Charges De Menard—Footings Over Rigid Inclusions: Experimental
Notation Validation of the Menard New Specifications.” Journées Nationales
de Géotechnique et de Géologie de l’Ingénieur—Nancy 2016
SEMELLES, 1–8.
The following symbols are used in this paper:
Bhasi, A., and K. Rajagopal. 2015. “Numerical study of basal reinforced
A = inclusion cross-sectional area (m2); embankments supported on floating/end bearing piles considering
Downloaded from ascelibrary.org by University of Central Florida on 03/10/24. Copyright ASCE. For personal use only; all rights reserved.

B = side length of the concrete square footing (m); pile–soil interaction.” Geotext. Geomembr. 43 (6): 524–536. https://
c’ = effective cohesion (kN/m2); doi.org/10.1016/j.geotexmem.2015.05.003.
d = diameter of circular inclusion (m); Briançon, L., D. Dias, and C. Simon. 2015. “Monitoring and numerical in-
E = Young’s modulus (MPa); vestigation of a rigid inclusions-reinforced industrial building.” Can.
Ei = average load efficiency for a single inclusion (%); Geotech. J. 52 (10): 1592–1604. https://doi.org/10.1139/cgj-2014-0262.
EL = average load efficiency for all inclusions (%); Chevalier, B., P. Villard, and G. Combe. 2011. “Investigation of load-
Em = pressuremeter (Menard) modulus in pressuremeter tests transfer mechanisms in geotechnical earth structures with thin fill plat-
(MPa); forms reinforced by rigid inclusions.” Int. J. Geomech. 11 (3): 239–250.
Es = settlement efficiency; https://doi.org/10.1061/(ASCE)GM.1943-5622.0000083.
Collin, J. G. 2004. “Column supported embankment design consider-
e = horizontal distance between the eccentric loading
ations.” In Proc., 52nd Annual Geotechnical Engineering Conf.,
center and the footing center (m); 51–78. Minneapolis, MN: University of Minnesota.
e/B = eccentricity ratio; Combarieu, O. 1988. “amélioration des sols par inclusions rigides verti-
G = shear modulus; cales application à l’édification de remblais sur sols médiocres amelio-
H = thickness of the load transfer platform (m); ration of soils by vertical rigid piles application to construction of
K0 = earth pressure coefficient; embankments on soft soils Laboratoire régional des Ponts-et-Cha.”
K0_NC = earth pressure coefficient at rest; 57–79.
r = the radius of circular inclusion (m); Dias, D., and B. Simon. 2015. “Spread foundations on rigid inclusions sub-
Iy and Iz = second moment of area of the inclusion’s cross section jected to complex loading: Comparison of 3D numerical and simplified
along y- and z-directions; analytical modelling.” Am. J. Appl. Sci. 12 (8): 533–541. https://doi.org
M = critical state ratio; /10.3844/ajassp.2015.533.541.
Dias, D., and J. Grippon. 2017. “Numerical modelling of a pile-supported
Mfy/Mfz = inclusion bending moment in the y/z-direction (kN · m);
embankment using variable inertia piles.” Struct. Eng. Mech. 61 (2):
N = axial load of the inclusion cross section (kN); 245–253. https://doi.org/10.12989/sem.2017.61.2.245.
pc0 = preconsolidation pressure (kPa); Filz, G. M. 2006. Design of bridging layers in geosynthetic-reinforced,
pL = limit pressure in pressuremeter tests (MPa); column-supported embankments. Rep. No. VTRC 06-CR12.
p′ = mean effective stress (kPa); Charlottesville, VA: Virginia Center for Transportation Innovation
Q = total load applied on the footing (kN); and Research.
Qp = load transmitted to one inclusion (kN); Guido, V. A., J. D. Knueppel, and M. A. Sweeney. 1987. “Plate loading
Qp_L/_R = load transmitted to the left/right side inclusion (kN); tests on geogrid-reinforced earth slab.” In Geosynthetic’ 87 Conf.,
qc = static cone penetration resistance (MPa); 216–225. St. Paul, MN: Industrial Fabrics Association International.
s = inclusion center-to-center spacing (for inclusions on a Han, J., and M. A. Gabr. 2002. “Numerical analysis of
square grid) (m); geosynthetic-reinforced and pile-supported earth platforms over soft
soil.” J. Geotech. Geoenviron. Eng. 128 (1): 44–53. https://doi.org/10
u = footing center settlement with reinforcement (mm);
.1061/(ASCE)1090-0241(2002)128:1(44).
uun = footing center settlement without reinforcement (mm);
Houda, M., O. Jenck, and F. Emeriault. 2021. “Soft soil improvement by
v = specific volume; rigid inclusions under vertical cyclic loading: Numerical back analysis.”
W = sum of the footing weight and the LTP layer weight; Eur. J. Environ. Civ. Eng. 25 (3): 409–428. https://doi.org/10.1080
x/y/z = coordinate system directions for inclusions; /19648189.2018.1531268.
α = rheological coefficient of soil; Itasca Consulting Group, Inc. 2017. FLAC3D — Fast Lagrangian Analysis
σxx = axial normal stress of inclusion (kPa); of Continua in Three Dimensions, Ver. 6.0. Minneapolis, MN: Itasca.
γ = soil unit weight (kN/m3); Jenck, O., D. Dias, and R. Kastner. 2009. “Three-dimensional numerical
ν = Poisson ratio; modeling of a piled embankment.” Int. J. Geomech. 9 (3): 102–112.
vλ = value of the specific volume for p′ —1 kPa (the https://doi.org/10.1061/(ASCE)1532-3641(2009)9:3(102).
reference pressure) on the critical state line; Jiang, Y., J. Han, and G. Zheng. 2014. “Numerical analysis of a
φ′ = internal friction angle (°); pile-slab-supported railway embankment.” Acta Geotech. 9 (3): 499–
511. https://doi.org/10.1007/s11440-013-0285-9.
ψ = dilation angle (°);
Kempfert, H., M. Stadel, and D. Zaeske. 1997. “Berechnung von geokunst-
λ = isotropic virgin consolidation line index; and stoffbewehrten Teagschichten über Pfahlelementen.” Bautechnik
κ = isotropic swelling line index. (Berlin, 1984), (12).
Mayne, P. W., and F. H. Kulhawy. 1982. “Ko-OCR relationships in soil.”
References J. Geotech. Eng. Div. 108 (6): 851–872. https://doi.org/10.1061
/AJGEB6.0001306.
Al-Naddaf, F., J. Han, C. Xu, S. Jawad, and G. Abdulrasool. 2019. Messioud, S., U. S. Okyay, B. Sbartai, and D. Dias. 2016. “Dynamic re-
“Experimental investigation of soil arching mobilization and degrada- sponse of pile reinforced soils and piled foundations.” Geotech. Geol.
tion under localized surface loading.” J. Geotech. Geoenviron. Eng. Eng. 34 (3): 789–805. https://doi.org/10.1007/s10706-016-0003-0.

© ASCE 04022113-12 Int. J. Geomech.

Int. J. Geomech., 2022, 22(8): 04022113


Messioud, S., B. Sbartai, and D. Dias. 2017. “Estimation of dynamic im- Simon, B., and F. Schlosser. 2006. “Soil reinforcement by vertical stiff in-
pedance of the soil–pile–slab and soil–pile–mattress–slab systems.” clusions in France.” In Vol. 1 of Symposium Rigid Inclusion in Difficult
Int. J. Struct. Stab. Dyn. 17 (06): 1750057. https://doi.org/10.1142 Subsoil Conditions, ISSMGE TC36, 3–23. Mexico: Universidad
/S0219455417500572. Nacional Autónoma de México.
Mirsepahi, M., A. Nayeri, S. M. Mirhosseini, and S. H. Lajevardi. 2021a. Svanø, G., T. Ilstad, G. Eiksund, and A. Want. 2000. “Alternative calcula-
“Investigation of the effects of twin tunneling on ground surface tion principle for design of piled embankments with base reinforce-
settlement and internal forces of a single pile in 3D analysis.” Soil ment.” In Proc., 4th Conf., of the GIGS, 542–548. Helsinki, Finland:
Mech. Found. Eng. 58 (1): 63–70. https://doi.org/10.1007/s11204-021 Finnish Geotechnical Society.
-09707-1. Terzaghi, K. 1943. Theoretical soil mechanics. New York: Wiley.
Mirsepahi, M., A. Nayeri, S. H. Lajevardi, and S. M. Mirhosseini. 2021b. Tran, Q. A., P. Villard, and D. Dias. 2021. “Geosynthetic reinforced piled
“Effect of multi-faced twin tunneling in different depths on a single embankment modeling using discrete and continuum approaches.”
pile.” Innovative Infrastruct. Solutions 6 (1): 1–12. https://doi.org/10 Geotext. Geomembr. 49 (1): 243–256. https://doi.org/10.1016/j
.1007/s41062-020-00425-5. .geotexmem.2020.10.026.
Nunez, M. A., L. Briançon, and D. Dias. 2013. “Analyses of a pile- van Eekelen, S. J., and A. Bezuijen. 2008. “Design of piled embankments,
Downloaded from ascelibrary.org by University of Central Florida on 03/10/24. Copyright ASCE. For personal use only; all rights reserved.

supported embankment over soft clay: Full-scale experiment, analytical considering the basic starting points of the British standard BS8006.” In
and numerical approaches.” Eng. Geol. 153: 53–67. https://doi.org/10 Proc., EuroGeo4. International Geosynthetics Society.
.1016/j.enggeo.2012.11.006. van Eekelen, S. J., A. Bezuijen, H. J. Lodder, and A. F. Van Tol. 2012. “Model
Okyay, U. S., and D. Dias. 2010. “Use of lime and cement treated soils as experiments on piled embankments. Part I.” Geotext. Geomembr. 32:
pile supported load transfer platform.” Eng. Geol. 114 (1–2): 34–44. 69–81. https://doi.org/10.1016/j.geotexmem.2011.11.002.
https://doi.org/10.1016/j.enggeo.2010.03.008. Wang, K., H. Liu, Y. Zhuang, and H. Xiao. 2018. “Multilayered low-
Pham, H. V., L. Briançon, D. Dias, and J. Racinais. 2019. “Investigation of strength geogrid-reinforced piled embankment.” Geotech. Res. 5 (4):
behavior of footings over rigid inclusion-reinforced soft soil: 231–246. https://doi.org/10.1680/jgere.18.00001.
Experimental and numerical approaches.” Can. Geotech. J. 56 (12): Xing, H., Z. Zhang, H. Liu, and H. Wei. 2014. “Large-scale tests of pile-
1940–1952. https://doi.org/10.1139/cgj-2018-0495. supported earth platform with and without geogrid.” Geotext.
Raithel, M., A. Kirchner, and H. G. Kempfert. 2008. “Pile-supported em- Geomembr. 42 (6): 586–598. https://doi.org/10.1016/j.geotexmem
bankments on soft ground for a high speed railway: Load transfer, dis- .2014.10.005.
tribution and concentration by different construction methods.” In Zhang, C., G. Jiang, X. Liu, and O. Buzzi. 2016. “Arching in
Proc., of the Int. Conf. Held in Nottingham, UK, on Advances in geogrid-reinforced pile-supported embankments over silty clay of
Transportation Geotechnics, 25–406. Boca Raton, FL: CRC Press. medium compressibility: Field data and analytical solution.”
Rui, R., J. Han, Y. Ye, C. Chen, and Y. Zhai. 2020. “Load transfer mech- Comput. Geotech. 77: 11–25. https://doi.org/10.1016/j.compgeo
anisms of granular cushion between column foundation and rigid raft.” .2016.03.007.
Int. J. Geomech. 20 (1): 04019139. https://doi.org/10.1061/(ASCE)GM Zhang, Z., F. Tao, J. Han, G. Ye, B. Cheng, and L. Liu. 2021a. “Influence
.1943-5622.0001539. of surface footing loading on soil arching above multiple buried struc-
Russell, D., and N. Pierpoint. 1997. “An assessment of design methods for tures in transparent sand.” Can. J. Civ. Eng. 48 (2): 1–10..
piled embankments.” Ground Eng. 30 (10): 39–44. Zhang, Z., F.-J. Tao, J. Han, G.-B. Ye, B.-N. Cheng, and C. Xu. 2021b.
Simon, B. 2010. “Une méthode simplifiée pour le calcul des semelles sur “Arching development in transparent soil during multiple trapdoor
sol renforcé par inclusions rigides.” JNGG 1: 529–536. movement and surface footing loading.” Int. J. Geomech. 21 (3):
Simon, B. 2012. “General report S5 rigid inclusions and stone columns.” 04020262. https://doi.org/10.1061/(ASCE)GM.1943-5622.0001908.
In Vol.1 of Proc. ISSMGE-TC 211 International Symposium on Zhuang, Y., E. Ellis, and H. S. Yu. 2012. “Three-dimensional finite-
Ground Improvement (IS-GI), 127–168. Belgium: TC 211 Ground element analysis of arching in a piled embankment.” Géotechnique
Improvement. 62: 1127–1131. https://doi.org/10.1680/geot.9.P.113.

© ASCE 04022113-13 Int. J. Geomech.

Int. J. Geomech., 2022, 22(8): 04022113

You might also like