Asiri National Project
Asiri National Project
Asiri National Project
Recommendations
for the design,
construction and control
of rigid inclusion
ground improvements
Summary
Close
ForeWord
These Recommendations on the improvement of foundation soils by inserting vertical rigid inclusions embody the achievement of the ASIRI National Project, whose research activities and
studies were conducted between 2005 and 2011 with a total budget allocation of 2.7 million.
This collective effort has relied on input from 39 partner firms and organizations; it has received
the financial support of the French Government and steady backing from IREX for all project
management aspects.
This project arose from a personal initiative coordinated through IREXs Soil Specialist Cluster, in conjunction with public works contractors and engineering companies within France, for
the purpose of accompanying the development of a new ground improvement technique. This
technique first appeared in France during the 1990s and consists of associating a network of
vertical rigid inclusions with a granular layer in order to compose the foundation platform for
civil engineering works (rafts, reservoirs, ground slabs, embankments) and construction works
(ground slabs). Though previous uses were already familiar and this technique beneath embankments had already been practiced in other European countries, the increasing application
of such procedures for civil engineering structures and buildings had become something of a
French national specificity, owing to the dynamic role of actors within the sector (construction
companies and consulting engineers). Yet in the absence of regulatory references, a set of dedicated and adapted guidelines needed to be developed and implemented.
The ASIRI National Project has focused on designing, conducting and interpreting a series of
physical experiments and numerical models as a fundamental step to understanding the mechanisms acting in this innovative foundation system. The applied research program presented
herein has been facilitated by project participants broad range of competences, stemming from
both academia and the professional world; the program has also greatly benefited from the
physical and human resources each partner organization made available.
A project database with an exceptional collection of experimental recordings and numerical
results has provided a valuable support tool for nine doctoral theses performed within the scope
of this very Project. A full slate of observations and model output could thus be compared for
a number of varied structures. New orientations were introduced, revealing the extent to which
associating such distinct elements (soil, inclusions, load transfer platform) produces an efficient
and highly economical composite foundation system. These Recommendations are derived
from the projects layout, which has yielded a set of practical rules for the design, construction
and control of these structures, intended for all industry actors interested in further developing
this technique.
Another specificity of these Project Recommendations lies in the proposed design strategy,
which conforms to the general Limit State design framework set forth in the Eurocodes, especially Eurocode 7-1 relative to geotechnical design. This application of general principles to
devise a method for ground improvement design constitutes an original and forward-looking
Foreword
approach, given that it was necessary to adopt specific rules for structures not easily identified
from the foundation categories established in Eurocode 7-1. The novelty of this technique has
been well substantiated, and the field of applications foreseen thanks to this breakthrough is
now wide open.
This valuable contribution was made possible thanks to the support of all partner organizations,
as well as to the tremendous motivation on the part of all who drove this work program so enthusiastically. Gratitude is extended to each and every one of them.
B. Simon
Director of the Scientific and Technical Committee
Summary
Summary
Foreword......................................................................................................................................3
Partners........................................................................................................................................5
Notations......................................................................................................................................7
Chapter 1. Description, history, initial developments and launching.
of the national project.............................................................................................19
Chapter 2. Operating mechanisms............................................................................................41
Chapter 3. Design methods......................................................................................................91
Chapter 4. Design considerations...........................................................................................145
Chapter 5. Justifications.........................................................................................................155
Chapter 6. Geotechnical investigations..................................................................................245
Chapter 7. Execution conditions............................................................................................263
Chapter 8. Controls and instrumentation................................................................................293
Cover
Close
Review Panel
President: F. SCHLOSSER (Professor Emeritus)
R. KASTNER (Professor Emeritus), A. GUILLOUX (Terrasol), P. LIAUSU (Mnard),
P. AGUADO (Apave) and F. BAGUELIN (Fondasol).
Figures: Mr. ANIC ANTIC (Terrasol)
English translation: R. SACHS
Summary
Notations
Notation
Descriptif
Description
Dimension
Units/
Units
L2
(m)
Ac
L2
(m)
AS
L2
(m)
AP
Surface area
of the inclusion head
L2
(m)
Longueur caractristique de la
dallette (tte dinclusion)
(m)
a, b
(m)
Reference dimension
of a shallow foundation
(m)
Cc
Indice de compression
Compression index
Cd
Cmax
Cs
Indice de recompression
Recompression index
()
CD
Taux de consolidation
secondaire
()
Creep index
()
CDe
Indice de fluage
PROJET
NATIONAL ASIRI
c
ck
()
ML1T2
ML1T2
(Pa)
(Pa)
()
cu
ML1T2
(Pa)
cuu
Cohsion apparente
dun sol non satur
Apparent cohesion
of an unsaturated soil
ML1T2
(Pa) 9
cv
Coefficient de consolidation
verticale
Coefficient of consolidation
in the vertical direction
L2T1
(m2/s)
Inclusion diameter
(m)
Notations
ck
()
cu
ML1T2
(Pa)
cuu
Cohsion apparente
dun sol non satur
Apparent cohesion
of an unsaturated soil
ML1T2
(Pa)
cv
Coefficient de consolidation
verticale
Coefficient of consolidation
in the vertical direction
L2T1
(m2/s)
Inclusion diameter
(m)
Dp
(m)
Ds
Diamtre quivalent du ft
pour une inclusion de section
non circulaire
(m)
dmax
(m)
di
(m)
dn
Dimension n % (n % des parti- n percent dimension (n% of particles by weight with a dimencules en poids ont des dimension of less than dn)
sions infrieures dn)
(m)
Load efficiency
()
Etass
Efficacit en terme
de tassement
Settlement efficiency
()
EV
()
ET
()
EQP
()
Notations
Ed
Module dYoung
ML1T2
(Pa)
E*
Module de dformation
apparent du sol renforc sous
chargement vertical
ML1T2
(Pa)
EM
ML1T2
(Pa)
Ei
ML1T2
(Pa)
Eoed
Module domtrique
Oedometric modulus
ML1T2
(Pa)
10
Eoedref
(Pa)
ML1T2
ASIRI
National
Project
1 2
EM
ML1T2
(Pa)
Ei
ML1T2
(Pa)
Eoed
Module domtrique
Oedometric modulus
ML1T2
(Pa)
Eoedref
ML1T2
(Pa)
Eoedo*
ML1T2
(Pa)
Ep
Youngs modulus
of the inclusion
ML1T2
(Pa)
Es
Module de dformation
apparent du sol
ML1T2
(Pa)
Esol
Module de dformation
du sol non renforc
Deformation modulus
of the non-reinforced soil
ML1T2
(Pa)
ref
ML1T2
(Pa)
Eurref
ML1T2
(Pa)
Void ratio
()
e0
()
emax
()
emin
()
FN
E50
MLT2
(N)
ML1T2
(Pa)
fck
Rsistance caractristique en Characteristic value of the compressive strength of the concompression du bton, coulis ou
mortier mesure sur cylindres crete, grout or mortar measured
on cylinders at 28 days
28 jours
ML1T2
(Pa)
fck (t)
Rsistance caractristique en Characteristic value of the compressive strength of the concompression du bton, coulis ou
mortier mesure sur cylindres crete, grout or mortar measured
on cylinders at t days
t jours
ML1T2
(Pa)
fck*
Rsistance caractristique la
compression du bton, coulis ou
mortier dune inclusion
Characteristic value of the compressive strength of the concrete, grout or mortar used on
an inclusion
ML1T2
(Pa)
fcvd
ML1T2
(Pa)
fctd
ML1T2
(Pa)
Valeur caractristique de la
Notations
11
an inclusion
fcvd
ML1T2
(Pa)
fctd
ML1T2
(Pa)
fc28
Valeur caractristique de la
rsistance la compression
28 j du bton dun dallage
ML1T2
(Pa)
ftfendage28
Valeur caractristique de la
rsistance en traction par fendage 28 j du bton dun dallage
ML1T2
(Pa)
fn
ML1T2
(Pa)
Module de cisaillement
Shear modulus
G11,G12,
G2,G3
G4
G*
Module de cisaillement
quivalent du sol renforc
Acclration de la pesanteur
terrestre
(g =
9,81m/s2)
(g = 9.81 m/s2)
()
ML1T2
(Pa)
ML1T2
(Pa)
LT2
(m/s2)
Notations
ML1T2
(Pa)
(m)
(m)
(m)
He
(m)
HR ou hR
Hauteur du remblai
Embankment height
(m)
HM ou hm
Hauteur du matelas
Platform height
(m)
HM
(m)
H, HS
(m)
(m)
hc
(m)
hi
(m)
g
g (z)
gmax
12 H
hr, ha
Hauteur dapplication du frotte- Height of negative friction application on the fictitious column
ment ngatif sur la colonne
positioned on top
fictive surmontant la tte
of the inclusion head
dinclusion
(m)
H, HS
(m)
(m)
hc
(m)
hi
(m)
Hauteur dapplication du frotte- Height of negative friction application on the fictitious column
ment ngatif sur la colonne
positioned on top
fictive surmontant la tte
of the inclusion head
dinclusion
(m)
L4
(m4)
hr, ha
ID
Indice de densit
Density index
()
IC
Indice de consistance
Consistency index
()
Ip
Indice de plasticit
Plasticity index
()
MT2
Coefficient de la loi de variation Coefficient of the modulus variadu module dans le modle rho- tion law in the "Hardening soil"
rheological model
logique Hardening soil
(N/m)
()
ML2T2
(kN/m3)
()
Ko
Facteur de pression
des terres au repos
()
Ka
Coefficient de pousse
des terres
()
Kp
()
Kq
Coefficient de transmission
dune surcharge dans un massif
non pesant
Coefficient of surcharge
transmission within a weightless medium
()
Kx
Impdance horizontale
dune fondation superficielle
MT2
(kN/m)
kh, kv
Coefficients de permabilit
horizontale et verticale
Coefficients of horizontal
and vertical permeability
LT1
(m/s)
ki
ML2T2
(kN/m3)
ks
ML2T2
(kN/m3)
ML1T2
(Pa)
kW, kq
Notations
13
ks
kW, kq
ML2T2
(kN/m3)
ML1T2
(Pa)
k1
()
k2
()
Notations
14k3
(m)
Inclusion length
(m)
(m)
(m)
Lal
(m)
Lat
(m)
Lfr
(m)
Lht
L1
L1 , L2
Reference dimensions of a
ground slab
()
ASIRI
Project
L National
(m)
Lat
(m)
Lfr
(m)
Lht
(m)
LM
(m)
ML1T2
(Pa)
ML2T2
(Nm)
Bending moment
ML2T2
(Nm)
Lmax
M
M NATIONALMoment
PROJET
ASIRI flchissant
Md
ML2T2
(Nm)
MR
ML2T-2
(Nm)
Mr
ML2T-2
(Nm)
Mx
ML2T-2
(Nm)
My
ML2T-2
(Nm)
15
Msup
ML2T-2
(Nm)
Minf
ML2T-2
(Nm)
MO
()
Exposant de la loi de variation Exponent of the modulus variadu module dans le modle rho- tion law used in the Hardening
soil rheological model
logique Hardening soil
ma
ML2T2
(Nm)
mb
ML2T2
(Nm)
mc
ML2T-2
(Nm)
mW , mq
()
()
()
Nc, Nq, N
Notations
Nk
()
ML2T2
(Nm)
ML2T-2
(Nm)
mc
mW , mq
()
Nc, Nq, N
()
Nk
()
Nkt
()
mb
Notations
()
Porosit
Porosity
Primtre de linclusion
Inclusion perimeter
(m)
Contrainte horizontale
contre une inclusion
Horizontal stress
against an inclusion
ML1T2
(Pa)
ML1T2
(Pa)
16 p
pl
Pression limite conventionnelle Standard limit pressure determidtermine lors de lessai pres- ned when conducting the stansiomtrique Mnard normalis dard Menard pressuremeter test
ML1T2
(Pa)
pl *
Pression limite conventionnelle Standard net limit pressure corcompte partir de la contrainte rected from the total horizontal
totale horizontale rgnant dans stress acting in the soil at test
elevation
le sol lors de lessai
ML1T2
(Pa)
plim
ML1T2
(Pa)
MLT2
(N)
Qd
MLT2
(N)
Q*
MLT2
(N)
Q p*
MLT2
(N)
MLT2
(N)
MLT2
(N)
MLT2
(N)
QR
Qs(z)
Qs
Qs(z)
de portance ELU
capacity
MLT2
(N)
MLT2
(N)
MLT-2
(N)
Qp(0) ou
MLT2
(N)
Qs(0) ou
MLT2
(N)
(Pa)
Qs
Q P+
QS
17
ML1T2
Variable component
of a surcharge
ML1T2
(Pa)
q0
ML1T2
(Pa)
q*
ML1T2
(Pa)
qa
Valeur asymptotique du dviaAsymptotic value of the hyperteur dans lajustement hyperbobolic fitting of stress-strain curve
lique de la courbe effortin primary triaxial loading
dformation dun essai triaxial
qc
Cone resistance
ML1T2
(Pa)
qp
ML1T2
(Pa)
qs
ML1T2
(Pa)
qt
ML1T2
(Pa)
q p*
ML1T2
(Pa)
qs*
ML1T2
(Pa)
qp+
ML1T2
(Pa)
qp+(P)
ML1T2
(Pa)
Notations
Notations
qp+(L)
ML1T2
(Pa)
qp,d+
ML1T2
(Pa)
q s+
ML1T2
(Pa)
(m)
(m)
Rc
MLT2
(N)
Rb
MLT2
(N)
(m)
RC
Rd
Rf
Rapport qf/qa
qf/qa ratio
Rk
Valeur caractristique
de rsistance
Rs
Roc
Rapport de surconsolidation
Overconsolidation ratio
Rc,cr ; d
()
MLT2
(N)
()
MLT2
(N)
MT2
(N/m)
MT2
(N/m)
Rt-L ; d
Valeur de calcul de la rsistance Design value of the ultimate tenultime de traction du gosynth- sile strength of the geosynthetic
in the longitudinal direction of
tique dans le sens longitudinal
the embankment
du remblai
MT2
(N/m)
Rt-T ; d
Valeur de calcul de la rsistance Design value of the ultimate tenultime de traction du gosynth- sile strength of the geosynthetic
tique dans le sens transversal in the transverse direction of the
embankment
du remblai
MT 2
(N/m)
19
Rv ; d
Valeur de calcul de la rsistance Design value of the ground bearing resistance beneath a shaldu terrain sous une fondation
low foundation
superficielle
ASIRI
Project
L National(m)
du remblai
the embankment
Rt-T ; d
Valeur de calcul de la rsistance Design value of the ultimate tenultime de traction du gosynth- sile strength of the geosynthetic
tique dans le sens transversal in the transverse direction of the
embankment
du remblai
MT 2
(N/m)
Rv ; d
Valeur de calcul de la rsistance Design value of the ground bearing resistance beneath a shaldu terrain sous une fondation
low foundation
superficielle
(m)
rk
(m)
rp
Rayon de linclusion
Inclusion radius
(m)
Aire de la semelle
Footing area
(m)
Sr
Degr de saturation
()
SRR
()
s
sc, sJ, sq
(m)
()
MLT2
(N)
Td
MLT2
(N)
Td
MT2
(N/m)
Td ; mesh
MT2
(N/m)
Td ; thrust
MT2
MT2
(N/m)
MT2
(N/m)
MT2
(N/m)
MT2
(N/m)
Rsultante horizontale du tor- Horizontal net force of the reacseur des ractions dveloppes tions developed in inclusions,
dans les inclusions pour la vrifi- for verification of ULS load-bearing capacity
cation de portance ELU
MLT2
(N)
MLT2
(N)
Td ; thrust ;
max
20T
geo
Tgeo, long
Tgeo, trans
TR
Notations
T
(N/m)
Notations
Tgeo, long
MT2
(N/m)
MT2
(N/m)
TR
Rsultante horizontale du tor- Horizontal net force of the reacseur des ractions dveloppes tions developed in inclusions,
dans les inclusions pour la vrifi- for verification of ULS load-bearing capacity
cation de portance ELU
MLT2
(N)
MLT2
(N)
Tv
Time factor
Degr de consolidation
ML1T2
(Pa)
LT1
(m/s)
Volume spcifique = (1 + e)
Specific volume = (1 + e)
VBS
Tgeo, trans
Vi
de la nappe gosynthtique
dans le sens longitudinal
()
(g/100 g
of soil)
MLT-2
(N)
Dplacement latral
Lateral displacement
(m)
MLT2
(N)
MLT2
(N)
MLT2
(N)
(m)
w (z)
Teneur en eau
Water content
wn
wL
Limite de liquidit
Liquid limit
()
wP
Limite de plasticit
Plastic limit
()
Tassement
Settlement
(m)
ys
Tassement du sol
Soil settlement
(m)
yp
Tassement de linclusion
Inclusion settlement
(m)
yM
Tassement la surface
du matelas
(m)
()
(-)21
Taux de couverture
()
Angular deformation
of a foundation
(rad)
ASIRI National
()Project
yp
Tassement de linclusion
Inclusion settlement
(m)
yM
Tassement la surface
du matelas
(m)
Taux de couverture
()
Angular deformation
of a foundation
(rad)
()
Dcc
()
()
Coefficient correcteur
pour valuer le module apparent du sol renforc
()
()
'
Foundation deflection
()
Dformation
Strain
()
H1ouHa
Dformation axiale
Axial strain
()
Hv
Dformation volumique
Volume strain
()
Icrit
()
Igeo
()
INC
()
Ipic
()
ISC
()
Angle de dilatance
Angle of dilatancy
()
Facteur de proportionnalit
entre les valeurs maximales
du dplacement horizontal et
vertical du sol autour dune
inclusion
()
*flu
()
22
Notations
Notations
()
Facteur de proportionnalit
entre les valeurs maximales
du dplacement horizontal et
vertical du sol autour dune
inclusion
()
*flu
()
*deg
()
*end
()
PROJET
ASIRI
J NATIONAL
Poids
volumique du sol
ML2T2
(kN/m3)
ML2T2
(kN/m3)
JC
()
JG
()
JM ; t
()
JQ
()
JR;h
()
JR ;d
()
JR ; f
()
JR ; v
()
Jb
()
Jc
Jcu
23
()
JR ; v
()
Jb
()
Jc
()
Jcu
()
Jgeo
Js
()
JI
()
JJ
()
Jr
ML2T2
(kN/m3)
Js
ML2T2
Notations
(kN/m3)
Jp
ML2T2
(kN/m3)
Jw
ML2T2
(kN/m3)
N*
Indice de dcompression
recompression dans modle
de comportement Soft soil =
2Cs /2,3 (1 + e)
Unloading-reloading index in
the constitutive soft soil model =
2Cs /2.3 (1 + e)
O*
Ocu
Facteur daugmentation de la
cohsion non draine cu en
fonction de laugmentation de
contrainte de consolidation
P*
Creep index used in the constitutive Soft Soil model = CDe /2.3
(1 + e)
Coefficient de Poisson
Poissons ratio
Contrainte totale
Total stress
ML1T2
(Pa)
Vc
Pression de confinement
Confinement stress
ML1T2
(Pa)
Vh
Contrainte horizontale
Horizontal stress
ML1T2
(Pa)
Vi
ML1T2
(Pa)
ML1T2
(Pa)
24
Notations
V
s
()
()
M1
1/m
()
Vh
Contrainte horizontale
Horizontal stress
ML1T2
(Pa)
Vi
ML1T2
(Pa)
Vs
ML1T2
(Pa)
Vc
Contrainte de consolidation
isotrope durant un essai
triaxial
ML1T2
(Pa)
Vv
ML1T2
(Pa)
Vv
ML-1T2
(Pa)
ML1T2
(Pa)
PROJET NATIONAL
ASIRIverticale effective
Contrainte
Vv
Vv, d
ML1T2
(Pa)
Vv, geo
ML1T2
(Pa)25
V 1, V 2,
V3
ML1T2
(Pa)
VV
V
Contraintes principales
effectives
ML-1T-2
(Pa)
V
ML1T2
(Pa)
VELS
ML1T2
(Pa)
Contrainte de cisaillement
Shear stress
ML1T2
(Pa)
Wcp
ML1T2
(Pa)
Rotation of a foundation
(rad)
[1, [2
Correlation factors for evaluating the results of inclusion static loading tests
()
[, [4
()
Inclination of a foundation
(rad)
Acronyms/abbreviations
ASIRI: Soil improvement by vertical rigid inclusions
CSTB: Scientific and Technical Resource Centre for the Building Industry
DTU: Unified Technical Document
DPM: Special Contract Documents
ELS: Serviceability Limit State (SLS)
ELU: Ultimate Limit State (ULS)
G11, G12, G2, G3, G4, G5: Standardized geotechnical missions (NF P 94500)
ICEDA: Active Waste Conditioning and Storage Facility (project conducted by Frances
EDF electric utility company at its Bugey site)
MV1, MV2, MV3: Simplified models for computing footings on rigid inclusions, subjected
to a vertical loading
MH1, MH2, MH3: Simplified models for computing footings on rigid inclusions, subjected
to a lateral loading
ma: Result of the ground slab computation on an equivalent homogenized soil (DTU 13.3,
Appendix C)
mb: Influence of rigid inclusions on a continuous ground slab
mc: Interaction between the rigid inclusions and the joints
OPM: Optimum value found by Modified Proctor compaction test
Soed(SJ): Computation result of a continuous ground slab resting on an equivalent homogenized soil
Soed(JT): Computation result of a ground slab with joints resting on an equivalent homogenized soil
IR(SJ): Computation result of a continuous ground slab resting on a network of rigid inclusions
IR(JT): Computation result of a ground slab with joints resting on a network of rigid inclusions
Summary
CHAPTER 1
19
The foundation on rigid inclusions approach is somewhat similar to the piled raft foundation
approach, in that it associates rigid vertical elements, i.e. inclusions, with a flexible cap layer,
yet without any rigid mechanical link placed between the two. From a purely theoretical
standpoint, the rigid inclusion concept remains pertinent once the conventional rigid linkage
between pile cap and piles disappears in the piled raft foundation configuration. In practice,
this transition from piled raft foundation to a foundation on rigid inclusions is reflected by a
combined geometric and mechanical discontinuity, along with the introduction of a relatively
thin mattress, generally a granular layer, called the load transfer platform.
This concept comprises various modes of interaction between: 1) the inclusions, each possibly
topped by a cap; 2) the load transfer platform directly supporting the foundation; and 3) the
foundation soil between the inclusions. The diagram in Figure 1.2 effectively summarizes this
range of interactions, with a differential settlement at the load transfer platform base, which
generates both a load transfer onto the inclusion caps via an arching effect taking place in the
load transfer platform and a negative friction along the inclusions over the upper part of the
soil between the inclusions. This load transfer by negative skin friction thus constitutes the
critical component of this concept and will be developed at length in Chapter 2.
20
21
As revealed in their mechanistic study (see Chapters 2 through 5), inclusions are intended to
carry a significant fraction of the loads applied to a soft soil volume. They must therefore be
designed to effectively withstand such loads; when given an appropriate cross-section,
inclusions are most often stopped in a ground layer exhibiting sufficient strength.
Typically not reinforced (especially in the presence of vertical loads), inclusions may
sometimes receive rebar additions in order to resist the eventual transverse forces being
exerted. This occurs, for example, when reinforcing the base of a tall embankment slope on a
soft or compressible soil, where sliding stability is a prerequisite.
Under this kind of scenario, in addition to placing rebar into the inclusions, it may be
requested to introduce above the pile caps, or inside the embankment body, other
reinforcements like horizontal geosynthetic sheets that absorb forces through traction and thus
limit vertical forces transmitted to the inclusions. Furthermore, the lack of a rigid bond
between inclusions and the supported structure, along with the absence of any group effect, is
indeed favorable in the event of seismic loading (Chapter 2).
3. PILE CAPS
The pile cap lies at the summit of the rigid inclusion. The surface area of its cross-section
divided by that of the reinforcement grid determines what is referred to as the "DUHD UDWLR.
This parameter, whose value often ranges between 2% and 10%, is an essential factor of
reinforcement efficiency.
In the majority of more common cases, the inclusion is simply embedded into the load
transfer platform; under certain conditions and for the purpose of raising this coverage rate,
the pile cap diameter may be increased or else the cap may be topped by a small slab.
The choice is available between uniform inclusions laid out in a relatively dense pattern and
elements sized with a diameter identical to the uniform inclusion but with greater spacing and
topped by a cap. Both cases offer the same cover rate.
Economic considerations however also play a role in the choice of cap type.
4. LOAD TRANSFER PLATFORM
The rigid inclusion concept implies that inclusion caps are not structurally bound to the
supported structure, as opposed to what is traditionally practiced for a foundation on piles
embedded into a tie raft.
In contrast, bonding is created by interspersing between the caps and structure a distribution
layer (or platform), on which the designed structure lies. The presence of this layer, called the
load transfer platform and most often composed of gravel, determines the specificity of the
rigid inclusion reinforcement technique.
Inclusion placement, combined with covering their caps by a load transfer platform,
ultimately leads to a composite or reinforced soil volume, which tends to be stronger and less
deformable than the initial soil volume, allowing the structure to lay on a shallow foundation.
22
A minimum load transfer platform thickness is necessary to allow for appropriate load
transfer between inclusions and soil, as well as to limit forces within the supported structure
(slabs, base plates, etc.). This thickness, often on the order of 40 to 80 cm, proves essential in
deriving an optimal design for the supported structure, particularly with the aim of reducing
bending moments in the slabs.
The load transfer platform may be composed of the following, depending on the application:
a simple layer of well-compacted granular material;
a layer of soil treated with hydraulic binder;
or a layer of soil reinforced by horizontal geosynthetic sheets.
In the case of a granular material layer, the objective consists of obtaining a high level of
compactness, which in turn yields a high modulus of deformation.
In the case of materials treated with hydraulic binder (lime or cement-lime mortar), the treated
layer must retain sufficient flexibility in order to avoid cracking; an extreme condition would
be described as a slab supported by deep foundations, whereby the principle of foundations on
inclusions would no longer apply.
A load transfer platform composed of a granular soil reinforced by geosynthetic sheets is
widespread in Anglo-Saxon countries. Geotextiles laid out in one or more horizontal sheets
within the load transfer platform are less commonly used in France, since they have been
deemed too easily deformable, especially when placed beneath slabs. The use of stiffer
geosynthetics, such as geogrids or reinforcing steel meshes, is more efficient given that
smaller deformations are sufficient to mobilize major forces. These horizontal reinforcements
can also be laid out in one or more sheets either inside the load transfer platform or at the base
of a homogeneous embankment, in which case they sit directly on the caps. For a given load
transfer platform thickness, the use of geosynthetics serves to lower the cover rate value.
The designer therefore has access to a wide array of potential solutions that would need to be
optimized as part of a comprehensive technical-economic approach aimed at improving the
design.
5. APPLICABLE SOILS
According to its underlying principle, the rigid inclusion system may be applied to all types of
soil conditions. In practice however, its economic benefit remains confined to soft or medium
soils, which are most often compressible, i.e. clay, silt or peat. Let's point out however that
peat and, more generally, all materials containing organic matter require special attention due
to the fact that they are subjected to secondary compression settlements.
The type of soil, which is often saturated, along with its index parameters, loading history and
mechanical properties are contributing factors in the choice of rigid inclusion production
mode, and some of these modes, like for piles, might not be advised.
The most widespread applications relate mainly to the limits of conventional soil
improvement techniques because of their inability to guarantee the necessary settlement
criteria or their requirement of a minimum quality threshold for the surrounding soil.
23
Reinforcement by rigid inclusions can also be effective for the purpose of building on former
waste storage sites. An analysis of such sites, in terms of chemical aggressiveness and
evolution, becomes critical when dealing with organic soils. Settlements, especially when
delayed, must be treated with special attention. The procedure for incorporating these
elements will be discussed in Chapter 6.
Under all circumstances, it is essential to recall that a foundation on rigid inclusions will
undergo settlement on the order of a few centimeters.
6. FIELDS OF APPLICATION
All structures, regardless of their nature, require both internal and external design in order to
ensure integrity throughout the construction period and then during the service life. This
integrity relies on whether the deformations, settlements, horizontal displacements and
distortions remain permissible, in recognizing the difficulty often involved when setting exact
thresholds.
European standardization also prescribes for geotechnical structures a specific deformation
design. Until recently, this step primarily revolved around: justifications at the point of failure
and the application of normal safety factors leading to relatively acceptable deformations for
the structures. Only very severe deformation conditions would motivate undertaking a
complete deformation design, which is often complex and ultimately results in choosing
higher safety factors than usually admitted. The foundation design, notably for compressible
soils, consists of identifying the most relevant solution that satisfies both stability and
deformation objectives while seeking to optimize construction costs and scheduling.
When the preliminary design of a structure reveals that the foundation soil in its current state
is incapable of carrying the loads transmitted by this structure without either excessive
settlement or incurring the risk of failure, then the designer must plan for a backup with deep
foundations or preliminary ground improvement or reinforcement.
The choice of solution thus depends on: the type of structure, type of applied loads (either
distributed or point), structure sensitivity to settlement, type of foundation soil, and
construction scheduling duration.
From a general standpoint, soil improvement solutions are more likely to be reserved for
structures with a large footprint and loads primarily of the distributed type.
As an example, a very tall building with a strong concentrated load on a poor quality
foundation soil will typically remain in the domain of more conventional deep foundations
with piles or diaphragm wall barrettes.
Conversely, many structures transmit distributed loads over large surface areas on the
foundation soil, as:
slabs and foundations of industrial and commercial buildings;
storage reservoirs (water, oil products or liquid chemicals), treatment plant basins and
retention facilities;
highway embankments or railway embankments for high-speed trains.
24
25
What are the specificities of these various types of structures for which rigid inclusion
solutions are commonly preferred?
For the first category, i.e. industrial and commercial building slabs, the problem to be solved
is basically one of limiting settlements, since the soil is typically capable of carrying these
distributed loads without any risk of failure.
For these structures, special attention is paid during load transfer platform design since the
slab structure tends to be thin and sensitive to the bending moments eventually created.
Furthermore, let's not overlook the fact that the operating load on the slab will not necessarily
be uniform (strips with fluctuating loads, traffic corridors, rack bases, etc.), all of which
serves to complicate the design step, as will be discussed in Chapter 5.
26
Let's also point out that should the slab foundation problem actually lead to a rigid inclusion
solution, then the same technology may be used as a complement underneath the structural
foundation footings.
For the second category, i.e. reservoirs, tanks and basins, the design criterion often stays
focused on settlements, whose magnitude is greater due to significantly higher stresses than
those found in the first category. Given the presence of load contributed by a liquid, the
operating stress will be perfectly uniform, except under the walls or skirts of tanks where
different localized reinforcements will be required.
Nonetheless, it must be noted that certain structures of the metal tank type are capable of
withstanding even greater settlements while remaining fully functional. For taller and more
heavily-loaded tanks, the problem of load-bearing capacity might also need to be addressed.
In the third and last category, i.e. embankments, these structures for the most part are loaded
in a permanent manner and by the embankment itself, with the extra operating load due to
road or rail traffic being relatively small. In this instance, it is difficult to identify a specific
structure, as opposed to the first two categories, and the embankment base is the target of a
specific treatment to ensure that the load transfer platform role is being fulfilled. For this
reason, full-scale testing undertaken within the scope of the ASIRI project has been divided
into two distinct full-scale experiments on inclusions: under slabs and under embankments.
For lower embankments, the design criterion will be based on settlements, while for taller
ones the stability problem will become predominant and generate the parasite horizontal
forces in the inclusions that need to be evaluated and taken into account.
7. A BRIEF HISTORY OF RIGID INCLUSIONS
Besides all available foundation soil improvement techniques, rigid inclusion reinforcement
offers a new and efficient technique. Yet in reality, this technique has already been used many
times over throughout history for building on difficult sites, such as marshy zones, without
necessarily a solid understanding of its mechanism and behavior.
It has longtime been observed that wood piles driven into the ground have served as
foundations for increasingly heavy structures over the ages. However, only when a sand layer
or stick bundle is positioned in between these piles and the structure can the thought be
entertained that rigid inclusion reinforcement is actually at work.
Noteworthy herein is the paper by Kerisel (1986), who cited an interesting publication by
Coles (1986) relative to the construction of Neolithic roads in England. The ingenious
underlying principle is as follows: long wood piles are driven at an angle in order to form a
succession of trestles, on top of which a traffic lane built on planks is placed, yet this
configuration can also comprise a line of vertical piles not aligned with the planks and placed
along a central axis, as shown to the right in Figure 1.4.
27
In the case described above, the planks are perforated with holes along the central axis to
enable the vertical piles to slide freely. While the trestles alone were in no way able to offer a
"foundation on rigid inclusions", the entire road foundation presented on the right-hand side
of the above figure provides a very clever example. The forces exerted on the traffic lane to
both sides of the central axis load the vertical piles with negative skin friction subsequent to a
soil settlement caused by the trestles. This mode of operations is similar to what was adopted
for certain buildings in Mexico City, as will be seen further below.
The modern period has led to technical and scientific improvements in the potential of these
foundation systems. A justification was provided with the study on negative skin friction
Zeevaert (1957) undertook in Mexico. Since Spains 16th-century conquest, Mexico City has
faced acute problems regarding the behavior of its edifices. A very thick layer of soft volcanic
clay (ash), which is both saturated and compressible, lies underneath the external loads and
experiences settlements up to several meters. These settlements were triggered and intensified
as of the beginning of the 20th century by the permanent pumping of water into deeper layers.
From the 1960's on, the technical literature provides many examples where both the negative
friction phenomenon and group effect are used advantageously, demonstrating that exterior
piles contribute more than piles located in the interior. The few original examples that follow
seem to have won only slight recognition.
A more subtle justification technique would be the so-called overlapping piles introduced in
Mexico by Girault (1969); Figure I.5 provides the corresponding diagram. The building rests
on piles A that do not reach as deep as the substratum but instead stop in the deformable
layers, in order to limit the substantial differential settlements occurring between the building
and the surrounding soil, which winds up being compacted. These building settlements
nonetheless require ensuring that any evolution diminishes over the time period following
construction. Short, type B inclusions, called piles for this purpose, are dropped into the
resistant layer and allow absorbing a fraction of the forces by means of the vault effect and by
mobilizing negative skin friction along their shaft. This set-up results in a notable decrease in
settlements and the appearance of significant forces in the type B inclusions.
28
Correa (1961), as related by Broms (1969) and Auvinet et al. (2006), designed buildings set
on an apron, with the combination resting on compressible soils undergoing consolidation.
Inside these soils were vertical piles lying on the substratum layer, the top of which crossed
the base of a hollow raft that had been designed as a box. This raft base was placed so as to
slide along the piles (see Fig. 1.6) which were acting like rigid inclusions. The uniform load
transmitted to the soil by the raft is then partially transferred onto the inclusions via negative
skin friction and the group effect. These inclusions have been designed relative to the soil
with a low safety factor and can be bonded to the raft without any technical difficulty should
it be sought to definitively stop the settlements.
29
In 1904 however, an engineer had proposed to support the Mexican Parliament palace
building on driven metal inclusions not bonded to the structure.
At present, Mexico City is the site of many construction projects featuring several thousand
non-reinforced concrete inclusions, where justifications have been conducted by applying a
finite element calculation method developed for this specific purpose.
In Japan, Okabe (1977) performed a full-scale study, on behalf of the Japanese Railways, of
groups of piles subjected to negative friction. The test site, containing 40 m of soft clay,
displayed an annual settlement of 10 cm due to water pumping in the depth. Isolated metal
piles, 60 cm in diameter, were able to bear up to 6,000 kN of undesirable forces due to
negative skin friction. Some heavier structures were supported on pile groups, which
themselves were surrounded by rigid inclusions independent of the pile group cap. The
recorded measurements confirmed transfer of the majority of parasite friction onto the
inclusions, with an unloading of piles bonded to the structure.
At the European level, the first papers on the subject of structures supported on rigid
inclusions relate to road or rail embankments. In this respect, Broms (1979) in Sweden
supplied abacuses to proceed with designing pile groups, with the use of expanded pile caps
being specified. Along these same lines, Rathmayer (1975) in Finland produced a series of
similar recommendations with applications for road embankments. Several embankments for
accessing bridges were built in Scotland in 1983 based on recommendations featuring the
simultaneous use of geotextile reinforcement sheets above expanded prefabricated pile caps.
Let's add the construction, for West Germany's railway system, beginning in 1976 of a 7 m
embankment on rigid inclusions topped by prefabricated caps, as part of a layout designed by
Smoltczyk (1976).
All observations and measurements performed during the construction of these various
structures have demonstrated the efficiency of reinforcement by rigid inclusions.
In France, the first publication on reinforcement by rigid inclusions, presented in Figure 1.7,
is credited to Gigan (1975), who used the technique without taking account of the mechanism.
Nonetheless, discussions held with the specialized contractor, as recounted by the author,
reveal the emergence of new ideas in the field of foundations. In the case studied, the focus
was to improve the characteristics of former embankments, clayey silts then sandy silts, by
means of driving cast-in-place piles without bonding them to the structures. This pile-driving
step actually led to very significantly enhancing the characteristics of in situ soils, which were
then deemed sufficient to accommodate a direct foundation. Designers however questioned
one another over the complementary role played by piles, which despite being isolated from
the base plates by a sand-gravel platform obviously performed like foundations on rigid
inclusions and not like conventional piles, as had been suggested.
30
Figure 1.7: Improvement in foundation soil characteristics thanks to the installation of piles (France).
The foundations presented in Figure 1.8 were designed in 1982 in Indonesia for reinforced
earth protection walls nearly 10 meters high by 6.5 m wide on a set of liquefied natural gas
retention reservoirs. These foundations are more technically sophisticated and rest on a 20 to
30-meter compressible clay layer. The walls are supported on open metal tubes 0.40 m in
diameter, with each one covered by a square metal plate 0.50 m to a side. The group of piles
is backfilled by a 1-m thick granular platform that had been horizontally stiffened by a
continuous structure of metal reinforcement strips merging and covering the driven pile caps.
While the rigid inclusions term had not been introduced at that time, this in fact was the
applied technique, though it was poorly understood 20 years ago. Simon and Schlosser (2006)
presented a paper on this case at an international congress on foundation soil reinforcement
held in Mexico City.
31
Two consecutive publications on the analytical design of rigid inclusion groups by Combarieu
(1988, 1990), built upon the findings of two highly detailed studies (theoretical, experimental
and bibliographical), were devoted to negative skin friction on piles. These works served to
boost popularity in France of the foundation soil reinforcement technique using rigid
inclusions.
The first study examined the construction of embankments on mediocre quality soils; it
provided a thoroughly justifying approach for the mesh pattern of a rigid inclusion groups and
for the resistant inclusion embedment, in addition to determining compressible soil settlement.
The second study related to calculating the inclusion groups for either large-sized flexible
structures or rigid shallow foundations. For these latter elements, a more targeted justification
had been proposed, dedicated to both load-bearing capacity and settlement, with this
justification being close to that developed for piled spread footing foundations.
These methods have been compared with the several existing European regulations as well as
with the experimental results of embankments on soft soils supported by rigid inclusions
(some of these results have already been cited). For rigid footings, the loading tests carried
out on a plate on sand reinforced by rigid inclusions, as conducted by Plumelle (1985) in a
large tank at the CEBTP testing facility, yielded results that match predictions fairly well.
The transition into practice occurred in 1989 at the time of an initial project, namely the
Carrre junction (Fig. 1.9), near the Lamentin-en-Martinique Airport (Combarieu et al., 1994).
These works were, technically speaking, conservative, with the objective of building bridge
access embankments consisting of Reinforced Earth abutments 6 to 7 m high. The footings of
these abutments were supported by HP piles crossing both the Reinforced Earth walls and in
situ soil.
The foundation soils were slightly sandy, saturated clays over a depth of 6 to 7 meters.
Without being treated, these soils would have led to a settlement of 40 cm. The 124 inclusions
were cast in place using a recovered metal tube and lost tip. Equipped with a central
reinforcement rod, these inclusions were all between 8 and 9 m long, with a 30 cm diameter
and 2.20 m spacing; they were subsequently covered by small, square prefabricated reinforced
concrete caps 0.80 m to a side. At the embankment base, a 10 cm thick concrete slab,
32
reinforced by a metallic latticework, was cast 20 cm above the pile caps. The settlements
observed at the end of the works period were equal to 5 cm, without any subsequent evolution.
More difficult projects ensued, with thicker and highly compressible soils that even contained
substantial peat layers. The outstanding problems became more complicated while at the same
time calculation methods underwent various changes, some significant examples of which
will be cited below.
In 1990, Simon and DApolito (1991) studied the foundations and access embankments of an
upper section over the A43 motorway, near Chambery in France (Fig. 1.10). This structure
was intended to double the existing facility with a foundation composed of heavily-loaded
piles on which any additional force impact needed to be prevented. Without special
precautions, the 7 meters of new embankment would have generated 80 cm of settlement and
an 11-cm horizontal displacement of the foundation soil, composed of 12 m of clayey soil.
The proposed solution consisted of foundation soil reinforcement by rigid inclusions, 37 cm
diameter and 15 m long, performed by using the soil mixing technique and supporting both
the new embankment and the shallow foundation of the new structure.
Figure 1.10: Abutment and embankment on rigid inclusions for the A43 motorway section,
Chambry (France).
Also during 1990, at Mandelieu on the A8 motorway (Bustamante and Gouvenot, 1991),
annual deformations of 15 cm reaching the surface and affecting the upper 15 m of 30 m high
marl embankments have been stopped using rigid inclusions. This significant highway
operations obstacle was in fact due to successive pavement thickness increases. Bored rigid
inclusions injected under high pressure over a height of approx. 18 m and then stopped in the
lower part of the embankment (built with high-quality materials) allowed absorbing by
negative skin friction the majority of the load due to the embankment's upper 15 m. All
movements were practically stopped since the residual deformation only measured 3 mm over
a 6-month period.
In 1999, the Nantes - Saint-Nazaire Port authority built a container dock on piles, in Montoirde-Bretagne on the banks of the Loire River (Combarieu and Frossard, 2003). The transition
between a slab on piles and the unstable riverbank was handled by floating rigid inclusions 30
meters long (Fig. 1.11) embedded into alluvial silts featuring 40-m power and average
characteristics at these depths.
Description, history, initial developments and launch of the national project
33
The inclusions, with a 3.50-m spacing and 54-cm diameter, were produced using a continuous
auger and composed of fiber-reinforced concrete. They were topped with 1.40 m u 1.40 m
prefabricated caps, which themselves were covered by a sheet of welded metallic latticework
laid out in the granular load transfer platform. The calculated settlement, in the absence of any
foundation soil treatment, was 1 m, and reduced to 20 cm with the rigid inclusions. In reality,
the measured settlements were between 5 and 10 cm.
Figure 1.11: Transition zone on inclusions between a dock on piles and the riverbank
(Montoir-de-Bretagne, France).
En 2000, it was proposed to found a closed frame made of reinforced concrete plus its
corresponding 6 m high access embankments on rigid inclusions, at the exit of Frances A15
motorway in Rouen, (Combarieu and Pioline, 2004). Use of this technique resulted in
successfully meeting the requirements of an imperative 4 month schedule for opening the new
lane to traffic. With 10 m of highly compressible soils, including peat, rigid inclusions 15 m
long and 40 cm in diameter, produced with a hollow auger and topped by caps, brought
settlements on the structure and its access ramps down to 4 cm in a homogeneous manner, a
feat that could not have been accomplished by deep foundations placed exclusively beneath
the concrete frame.
The ICEDA (French acronym for Conditioning Installation and Storage of Active Wastes)
project conducted in 2010 by France's electricity utility EDF at the Bugey site is a building
intended to temporarily store wastes from dismantled 1st-generation nuclear power plants and
the Super-Phenix reactor, plus wastes produced from sections of the pressurized water reactor
in operation. The predicted life span for this installation is 50 years. Its dimensions are on the
order of 120 m by 60-70 m, and average loads range from 140 kPa (under the space used to
store packages of waste) to 240 kPa (under the packaging workshop).
Prior to these works, the ICEDA project zone stratigraphy was as follows:
An average 5 m of sand-gravel alluvia offering high geotechnical quality;
35 to 55 m of silty clay;
A molasse substratum located at a depth of between approx. 40 to 60 meters (pitch of the
layer top on the order of 15%).
34
The settlements evaluated during the preliminary design phase, of roughly 25 cm, were
deemed excessive, and a solution to reinforce the compressible clay layer by means of rigid
inclusions was adopted. This solution comprised (Fig. 1.12):
292 reinforced concrete inclusions 1 meter in diameter, with heads at least 2 meters below
the underside of the rafts and tips reaching the weathered molasse (i.e. a length of between
~37 m and ~57 m), with spacing varying between 3 m and 6 m;
A load transfer platform 2 to 2.75 m thick, composed of site alluvia homogenized and
carefully compacted.
The building settlement calculation on the reinforced soil led to values on the order of 4 cm.
35
Figure 1.13: The Rion-Antirion Bridge - Principle of the foundation solution on inclusions (Greece).
These examples illustrate a few of the varied geotechnical problems that the new technique
involving rigid inclusions has been able to solve. While up until 2000 the building of
foundations using this process remained relatively infrequent, the number of such projects has
increased since then significantly. This concept is now being proposed by the majority of
contractors specialized in foundation works and soil treatment and moreover has been
accepted by structural inspection bodies and insurance companies, provided that the
contractors have been given a relevant set of specifications. Project Owners and Architects
alike have found in this technique an effective response to project deadline constraints. The
introduction of a slab and raft foundation has generated a number of attractive applications.
Alongside these advances, many computation models have been developed in France, by
either contractors or geotechnical design firms, entailing the use of analytical models relying
on the concept of negative friction or numerical models incorporating the finite element
method.
8. THE ASIRI NATIONAL PROJECT
As of the end of the 1990s, many projects could already be cited in which rigid inclusions
constituted the foundation system. Their use has only expanded since that time and the claims
rate recorded through use of this technique remains, at present, comparable to that of more
36
37
2002). The memoranda on these accomplishments then resulted in an IREX report published
in November 2002 (Brianon, 2002) offering a view of the current state of practices, by
highlighting outstanding project developments / experiments and relying on the results of a
survey conducted among specialized French firms.
The general program adopted by the National Project was presented on February 4th, 2005.
This R&D project was planned to span a 4 year period, with the first three years being
devoted to establishing and then implementing experimental and numerical protocols as well
as an analysis of the initial experimental data, and the final year spent synthesizing all data
collected, validating the various design methods and producing a document containing
recommendations.
The launch of this project was timed to coincide with the completion of the doctoral thesis
defended by Orianne Jenck, in November 2005 at the INSA de Lyon institution, dealing with
the physical and numerical modeling of the technique of reinforcement by rigid inclusions.
During the first Executive Committee meeting held on March 29th, 2005 at the FNTP offices
under the aegis of IREX, the National Project (abbreviated under the French acronym ASIRI,
for Soil Improvement by Rigid Inclusions) designated the following officers: President F.
Schlosser, Vice President O. Combarieu, and Scientific Director B. Simon.
Five working topics shape the scope of this project:
Topic 1: Full-scale experiments, hosted by C. Plumelle and L. Brianon (Geotechnical
Engineering Chair at the Cnam). This topic treats full-scale experimental campaigns in an
extremely thorough manner and solely to satisfy research needs.
Topic 2: Experimental sites, hosted by E. Haza (CER Rouen). This topic will focus on the
limited instrumentation built for actual project sites, with the aim of compiling results.
Topic 3: Laboratory testing, hosted by L Thorel (IFSSTAR Nantes). The scope here covers
all tests intended to characterize materials and develop centrifuge-based tests.
Topic 4: Numerical modeling, hosted by D. Dias and R. Kastner (INSA Lyon).
Topic 5: Recommendations, hosted by O. Combarieu.
38
References
Auvinet G., Rodriguez J.-F. Rigid inclusions in Mexico City soft soils: History and
perspectives. Symposium Rigid inclusions in difficult subsoil conditions, ISSMGE TC36,
Sociedad Mexicana de Mecanica de Suelos, UNAM, Mexico, 2006.
Brianon L. Renforcement des sols par inclusions rigides. eWDW GH Oart en France et
Otranger . Irex, opration du Rseau gnie civil et urbain, 2002.
Broms B.B. - Design of piles group with respect to negative skin friction; Specialty session
on negative skin friction and VHWWOHPHQWRISLOHGIRXQGDWLRQV 7th.ICSMFE, Mexico, 1969.
Broms B.B. - Problems and solutions to cRQVWUXFWLRQLQVRIWFOD\Proc. 6th Asian regional
conference on soil mechanics and foundation engineering, Singapore, vol. 2, 1979, p 2830.
Bustamante M., Gouvenot D. &RQIRUWHPHQW GXQ UHPEODL GH JUande hauteur par
injection . 10e confrence internationale de mcanique des sols et travaux de fondation,
Florence, mai 1991.
Coles J. - Sweet track to Glastonbury. Thames and Hudson, Londres, 1986.
Combarieu O. - Amlioration des sols par inclusions rigides verticales. Application
OpGLILFDWLRQ GHV UHPEODLV VXU VROV PpGLRFUHV . Revue franaise de gotechnique n 44,
1988, p. 57-79.
Combarieu O. Fondations superficielles sur sol amlior par inclusions rigides verticales .
Revue franaise de gotechnique n 53, 1990, p. 33-44.
Combarieu O., Frossard A. Amlioration des sols des berges de Loire par inclusions
ULJLGHVSRXUVWDELOLVHUOHVUHPEODLVGDFFqVGXQSRVWHjTXDL . 13e Congrs europen de
mcanique des sols et fondations, Prague, 2003.
Combarieu O., Gestin F., Pioline M. - Remblais sur sols amliors par inclusions rigides.
Premiers chantiers . Bulletin des laboratoires des ponts et chausses n 191, 1994, p. 5562.
Combarieu O., Pioline M. - &RQVWUXFWLRQ GXQ 3,&) HW GH VHV DFFqV VXU sol amlior par
inclusions rigides 6\PSRVLXPLQWHUQDWLRQDOVXUOamlioration des sols en place, ASEPGI, 9-10 sept. 2004, Paris.
Gigan J.-P. - &RQVROLGDWLRQGXQVROGHIRQGDWLRQSDUSLORWV . Bulletin des laboratoires des
ponts et chausses n 78, 1975, p. 12-16.
Girault P. $ QHZ W\SH RI SLOH IRXQGDWLRQ Proc. Conf. on deep foundations, Mexican
society of soils mechanics; Mexico, 1969, vol. 1.
IREX - Ple Comptence Sols. Utilisation des inclusions rigides pour le renforcement des sols
de fondaWLRQGouvrages et de remblais.eWXGHGHIDLVDELOLWpGun projet national, 2000.
Jenck O. - Le renforcement des sols par inclusions rigides verticales. Modlisation physique
et numrique. Thse de doctorat INSA Lyon, 2005.
Kerisel J. - Angleterre ; dcouYHUWH GXQH URXWH QpROLWKLTXH . Revue Archologia n 292,
1986, p 56-60.
Laurent Y. - Renforcement des massifs de fondation par inclusions rigides verticales, tude
bibliographique et numrique. Mmoire DEA INSA Lyon, 2002.
Okabe.T. Large negative skin frictionDQGIULFWLRQIUHHSLOHPHWKRGV 9th ICSFME, Tokyo,
vol. 1, 1977.
Pecker A. Le pont de Rion-Antirion en Grce, le dfi sismique . 550e confrence,
SURQRQFpHjO8QLYHUVLWpGHWRXVOHVVDYRLUVOHRFWREUH.
Plumelle C. - RenforcementGXQVROOkFKHSDULQFOXVLRQGHPLFURSLHX[ . Revue franaise
de gotechnique n 30, 1985, p.47-57.
Rathmayer H. Piled embankment supported by single pile caps . Proc. Istanbul
conference on soil mechanics and foundation engineering, 1975, 8 p.
Description, history, initial developments and launch of the national project
39
Simon B. - 8QH PpWKRGH LQWpJUpH SRXU GLPHQVLRQQHU OHV UpVHDX[ GLQFOXVLRns rigides en
dformation . XVe Congrs international de mcanique des sols et fondations, Istanbul,
vol. 2, 2000, p. 1007-1010.
6LPRQ % '$SROLWR 3 'RXEOHPHQW GXQ SDVVDJH VXSpULeur fond sur pieux ayant subi
des efforts parasites importants . Colloque international fondations profondes, ENPC,
Paris, mai 1991, p.285-292.
Simon B., Schlosser F. - Soil reinforcement by vertical stiff inclusions in France .
Symposium Rigid inclusions in difficult subsoil conditions, ISSMGE TC36, Sociedad
Mexicana de Mecanica de Suelos, UNAM, Mexico, 2006.
Smoltczyk - 3IlKOJUQGXQJ HLQHV (LVHQEDKQVGDPPV 6e Congrs europen de mcanique
des sols et fondations, Vienne, 1976, vol. 3.
Zeevaert L. Discussion on negative friction and reduction of SRLQWEHDULQJFDSDFLW\Proc.
4th.ICSMFE, 1957, vol. 3.
40
Summary
CHAPTER 2
Operating mechanisms
This chapter presents the operating mechanisms applicable to a group of inclusions, which
serves as an essential prerequisite to raising the notions of Design and Dimensioning in
subsequent chapters. These mechanisms, which pertain to load transfer and subsequent impact
on settlements, rely to a considerable extent on full-scale experiments, in addition to physical
and numerical models, conducted within the scope of the ASIRI National Project; however,
they also incorporate previous results, especially as regards the negative skin friction model
proposed by Combarieu (1974).
According to its underlying principle, the soil reinforcement technique using rigid inclusions
introduces a whole sequence of complex interactions, from the structure being supported
down to the substratum layer:
Interaction between the structure and the load transfer platform, depending on both platform
thickness and foundation stiffness;
Interaction between the transfer platform and the reinforced soil block composed of the soft
soil layer and rigid inclusions;
Interaction at the interface between the soft soil and the inclusions, where successively
negative then positive friction develops over the depth;
Lastly, interaction between inclusion tip and load-bearing soil.
Once these interaction mechanisms have been demonstrated, attention will focus on
explaining the negative skin friction model, as well as load transfer mechanisms and the
associated settlements and deformations. In order to trim the length of Chapters 3 and 5,
dedicated respectively to design considerations and justifications, the analytical formulations
related to negative skin friction and load transfer in the platform will be developed in this
chapter.
As a conclusion to this chapter, lessons will be drawn from full-scale field experiments, which
are considered an essential step towards producing an effective design, along with a review of
seismic loadings as set forth in the Guide on Soil Improvement and Reinforcement Processes
under Seismic Action, published with the support of the French Earthquake Engineering
Association (AFPS).
1. THE STRUCTURE - TRANSFER PLATFORM - INCLUSION HEAD
INTERACTION
1.1. Interaction between the Reinforced soil and the Load transfer platform
Using the set of notations selected for the ASIRI National Project framework, the load
transfer mechanism on inclusions within the granular platform can be quantified by means of
the following parameters (Fig. 2.1):
Operating mechanisms
41
Qp
W Q
and G
' Qp
'Q
(2.1)
In order to describe the transfer mechanisms, it would be useful to distinguish the case of an
embankment from that of a ground slab or shallow foundation.
1.1.1. Case of an embankment
The simplest case would be that of a load transfer platform composed of a granular material
inserted at the base of an embankment; this case has been studied in simplified form by
considering a uniform load exerted on the transfer platform. In using for this transfer platform
a three-dimensional model built with discrete elements, results demonstrate that, for a given
configuration, efficiency E increases with load until reaching a threshold value. This point
corresponds to the gradual mobilization of shear strength in the granular material, through
saturation. Moreover, calculations indicate the presence of tapered and relatively stationary
zones within the transfer platform on top of the inclusion heads, where shear concentrates at
the edge (Fig. 2.2, Chevalier et al., 2010). By this shear, load is transferred from peripheral
zones towards the inclusion heads, with only slight displacements associated. The opening of
cones is correlated with the peak friction angle of the granular material (or else with the
critical state friction angle should the surrounding soil be highly compressible).
1
42
Figure 2.2: Displacement fields (top view and vertical cross-section between rigid inclusions) for
various granular transfer platform thicknesses (hm = 0.5 m, left) and (hm = 1.0 m, right).
The ultimate load efficiency increases with transfer platform thickness hm until reaching a
threshold that defines a critical thickness. This thickness value may be interpreted as the
height from which the cones meet; it also corresponds to the complete development of stress
chains related to the most heavily-loaded grains in the granular medium that form like arches
inside the load transfer platform (Fig. 2.3, Chevalier et al., 2010).
The critical thickness depends on both the spacing between inclusions and mechanical
characteristics (deformation modulus and friction angle) of the load transfer platform. A
series of centrifuge experiments by Baudouin (2010) enables evaluating this thickness.
Among other findings, these tests demonstrated that the critical thickness is found for an hm/(s
- a) ratio value close to 2/3, with hm being the transfer platform thickness, and (s - a) the clear
span spacing between inclusions.
Figure 2.3: Stress chain networks inside the granular transfer platform (hm = 1.0 m).
Operating mechanisms
43
In continuous numerical models, this phenomenon corresponds to the full rotation of principal
stresses at the top of the transfer platform halfway between the inclusions.
Jenck's two-dimensional model (2005), which yields access to the displacement field in the
load transfer platform, highlighted that increasing the platform thickness mitigates the
differential settlements at the surface. This is illustrated in Figure 2.4, where the evolution of
settlements in the transfer platform is plotted, both in alignment with an inclusion and
between two of them. According to this two-dimensional model, a relative transfer platform
thickness hm/(s - a) equal to 1.3 prevents complete development of the arch, which is reflected
by a persistent differential settlement at the surface. A relative thickness hm/(s - a) equal to 2.0
would enable completing the arching development, as reflected by a quick equalization of
settlements within the load transfer platform. It should be pointed out that these values of 1.3
and 2.0 are taken only as examples due to the model limitations.
Figure 2.4: Vertical displacement of the two-dimensional model in alignment with an inclusion
and between two successive inclusions, during consolidation at an area ratio D = 22%.
From a general standpoint, Jenck (2005) and the series of physical models developed within
the scope of the ASIRI National Project (Jenck et al., 2006) suggest that compressibility of
the surrounding soil exerts no significant influence on efficiency E, as long as the soil is
capable of providing the required level of confinement around the inclusion heads. When this
confinement risks no longer being sufficient, the discrete models reveal the beneficial role
played by a geosynthetic reinforcement placed at the base of the transfer platform.
The following graphs (Fig. 2.5), extracted from Chevalier (2008), illustrate this role of
geosynthetic reinforcement in the case of no downdrag effect between the soft soil and
inclusions. They also display the evolution in efficiency E vs. applied loading with two values
of stiffness J for the geosynthetic and a surrounding soil of variable deformability. It can be
observed that the primary contribution of the geosynthetic becomes tangible for the most
highly-compressible supporting soils. In this case, the geosynthetic, provided it offers
adequate stiffness, is able to mobilize sufficient tension to allow confining the transfer
platform base.
44
Figure 2.5: Efficiency of the granular transfer platform E vs. surcharge q0 for various coefficients
of subgrade reaction; hm = 0.5 m without a ground slab (discrete 3D model).
Operating mechanisms
45
Figure 2.6: Evolution of boundary efficiencies vs. transfer platform thickness for the "embankment"
and "ground slab" cases (physical model run in a calibration chamber).
The following diagrams (Fig. 2.7) illustrate the load transfer mechanism underneath both an
embankment and a ground slab based on the results of a continuous 3D model using the Flac
software (Laurent, 2002).
,QWKHHPEDQNPHQW case (Fig. 2.7a), load transfer is driven by shear inside the embankment,
with a rotation of the principal stresses that reflects arching development.
,QWKHground slab case (Fig. 2.7b), load transfer is governed by slab stiffness and leads to a
stress concentration in the transfer platform columns situated on top of the inclusion. These
columns become preferred zones in the development of stress chain networks. The stability of
each column is provided by the confinement existing in the transfer platform as well as by the
underlying soil reaction. Insufficient confinement degrades system operations. This point
must always give rise to a specific verification in the ground slab case.
In both cases at the level of the transfer platform base, it is noticed that the settlement gradient
is localized in the immediate vicinity of inclusion heads, with settlement being relatively
uniform between inclusions.
46
Figure 2.7: Comparison of load transfer mechanisms underneath an embankment and ground slab
(continuous 3D model).
Operating mechanisms
47
Figure 2.8: Mobile tray set-up for centrifuge tests and simulation principle for soil settlement.
48
Figure 2.9: Evolution in the load applied at the inclusion head with soil settlement
(comparison between lime treatment and cement treatment) (Okyay, 2010).
Nearly all load is transmitted onto the heads of rigid inclusions. For higher loads, once the
transfer platform has been damaged, efficiency drops instantaneously. This phenomenon is
correlated with the transfer platforms fragile behavior. Given the considerable stiffness of
transfer platforms composed of cement-treated material, the stresses transmitted prior to
failure on the soft soil may be very small, or even nonexistent.
The load transfer platform behaves like a slab loaded in bending. If the transfer platform is
covered by a ground slab or raft, then the bending loadings will be distributed between this
structural element and the transfer platform in proportion with their respective stiffness and
according to the interface bonding condition.
Transfer platform behavior may be influenced by actual placement conditions. An ineffectual
treatment during reworking of the successive layers can considerably affect transfer platform
homogeneity as a result of inadequate adhesion.
1.1.4. Edge effect
At the edge of the structure, interactions between the load transfer platform and inclusion
heads are modified relative to conditions existing in the central part of the structure, due to the
following factors:
Asymmetry between the loading zone and the exterior unloaded zone, which necessarily
causes significant stress field variations in the transfer platform;
Limited extension of the load transfer platform around the inclusion;
The transfer platform's reduced lateral confinement capacity;
The transfer platform's vertical confinement limited to the structural footprint (ground slab
or footing);
A non-uniform stress distribution on the underside of a stiff foundation (stress peak at the
foundation edge).
These factors can induce specific failure mechanisms with, as a consequence, a lower load
limit capable of being applied at the inclusion head.
The continuous 3D model that served to justify the ICEDA project (as described in Chapter 1,
Section 7) illustrates that the stress field at the head of the peripheral row can be distinguished
Operating mechanisms
49
from that of the interior rows, which appears to be independent of position (Fig. 2.10). In this
project, the inclusions were 1 m in diameter, with 6 m spacing, and the load transfer platform
was 2 m thick. The present calculation assumes a 200 kPa load (including the weight of the
raft).
Figure 2.10: Example of a stress field at the periphery of the structure (ICEDA project).
In this case, which corresponds to a relatively thin transfer platform, nearly the same
stresses have been measured on the rigid inclusion as on the underside of the ground slab in
alignment with the inclusion: the bulk of load transfer thus takes place through the ground
slab;
As consequence, stresses measured under the ground slab are sharply lower between
inclusions than when aligned with them: the distribution of reactions is therefore nonuniform, a finding that must be taken into account when designing ground slabs.
Experimental and numerical studies have demonstrated that a thinner transfer platform
enhances the efficiency of rigid inclusions in the presence of a ground slab, inasmuch as the
ground slab will participate in the load transfer process. Compared to an infinitely thick
transfer platform, in which failure shear mechanisms cannot extend to the upper part of the
transfer platform, the presence of a ground slab on a thin transfer platform modifies how these
mechanisms develop through loading of the ground slab itself.
If a limited transfer platform thickness were associated with high mechanical characteristics
(modulus of compressibility, friction angle, cohesion), it would be possible to obtain optimal
efficiency, leading in turn to a maximum stress concentration on inclusion heads and therefore
a maximum decrease of settlements. In exchange, the ground slab would be more heavily
loaded in bending, as indicated in Figures 2.12 and 2.13 (Okyay, 2010). The characteristics of
both transfer platform and ground slab need to be defined so as to identify an optimum
between loadings in ground slab and settlements.
Operating mechanisms
51
Figure 2.13: Influence of the transfer platform friction angle and cohesion on bending moments in the
ground slab (Okyay, 2010; Okyay and Dias, 2010).
Ground slab: thickness = 20 cm;
Transfer platform: height = 60 cm and modulus E = 70 MPa, variable friction angle or cohesion;
Inclusions: diameter = 35 cm and spacing = 2 m.
The definition and control of the transfer platform component material, as well as its level of
compaction and aptitude for eventual treatment, are all keys to ensuring effective operations
of this transfer platform on a group of rigid inclusions. It would seem essential to define a
range of minimum and maximum values for the transfer platforms mechanical
characteristics, while acknowledging the placement conditions and their eventual evolution
(e.g. lime-treated soil, whose characteristics vary according to moisture conditions). The
52
upper values derived typically correspond to an upper bound of structural element loadings,
and lower values are associated with an upper bound of settlements.
The soil study executed in accordance with the sequence of geotechnical engineering missions
(see NF P 94-500 Standard, December 2006) is expected to establish a succession of soil and
fill layers, along with the values of parameters required as input for the calculation model
(Chapters 4 and 6).
Figure 2.14: Distinct zones to be considered for ground slab design and calculation.
The cases of point loads will also be determinant in the ground slab design process. In
practice, ground slab stiffness relative to the foundation base will imply a redistribution of
point forces on each mesh. The behavior of the foundation base (reinforced soil + transfer
platform) will remain controlled by the resultant of various reactions derived (by analogy with
Barr de Saint-9HQDQWs regularization principle for beams). Along these lines, the behavior
of the foundation base below the ground slab can be related to the behavior found below a
uniform loading with the same mesh resultant.
Under these conditions, the ground slab needs to be studied when exposed to a combination of
point loads at the ground slab surface, along with the reactions calculated on the underside.
Such reactions are estimated for an equivalent uniform loading applied to each mesh (see
Chapter 5: Justifications).
Operating mechanisms
53
54
Inclusions are mobilized not only by the load directly applied on their head, but also by the
friction of the surrounding soil, subjected to settlement under the loading applied by the
weight of the structure.
Over the upper part of the reinforced soil, the soil settles to a greater extent than the
inclusions, whereas at the base the inclusions settle more than the soil. A depth therefore
exists where both soil and inclusions settle to the same extent: this is known as the neutral
point. In the upper part, this effect takes the form of a negative friction loading the inclusion,
until reaching a maximum compressive load at the neutral point (Fig. 2.15).
This load is absorbed below the neutral point by means of positive friction and tip resistance.
The incorporation of soft soil action on the inclusions is complex since it involves a
combination of the following types of behavior:
Transfer of forces by means of friction from the inclusion towards the soil and away from
the soil;
Transfer of forces at the inclusion tip;
Soil settlement between inclusions;
Inclusion settlement by elastic shortening and tip penetration.
The main parameters to be integrated into the operating mechanism that describes this soilinclusion interaction are as follows:
Area ratio;
Soil overconsolidation ratio;
Deformability of the soft soil;
Deformability of the load-bearing soil;
Eventual creep rate of the soft soil (case of organic soils, peat, etc.);
Modulus of elasticity of the inclusion;
Soil-inclusion shaft friction, negative as well as positive, and tip resistance.
Operating mechanisms
55
All these parameters exert an influence on both the stress distribution and differential
settlements between the soil and rigid inclusions.
Remark: An evaluation of the shaft friction and tip resistance parameters must
take into account the type of inclusion and its installation technique, by
distinguishing those inclusions performed without soil displacementHJbored
piles) from those executed with soil displacementHJdriven steel pipe piles).
2.2. Negative skin friction
2.2.1. Underlying principles
Negative skin friction is produced when soil settlement along the rigid inclusion shaft exceeds
the vertical displacement of the rigid inclusion itself.
In contrast with the pile, rigid inclusions do not experience negative skin friction like a
parasitic force, but instead the negative friction serves to relieve the soft soil layer of the
residual stress existing at the transfer platform base. Such transfer onto the inclusions occurs
gradually with depth.
2.2.2. Neutral point
The equality between the rigid inclusion vertical displacement and the vertical soil settlement
along the rigid inclusion serves to define the depth hc of the neutral point N. Above this point
N, friction is negative, while below it turns positive (Fig. 2.15).
At its head, the rigid inclusion absorbs a load QP(0).
At point N, the rigid inclusion is subjected to this load QP(0) increased by the negative skin
friction force FN (FN = 0hc W(z).SD.dz) between the rigid inclusion head and point N.
The axial force in the rigid inclusion therefore rises between inclusion head and neutral point.
This total force [QP(0) + FN] must then be balanced by the positive friction force FP below
point N plus the tip force QP(L) (Fig. 2.15): QP(0) + FN = FP + QP(L).
In the vicinity of the inclusion shaft,VWUHVVv thus decreases with depth since a portion of the
ground weight is transmitted to the inclusion via the negative skin friction mobilized above
this given point: such an effect is known as the downdrag effect.
2.2.3. Mobilization over both time and depth
The maximum negative skin friction depends on:
Soil settlement (as a function of its compressibility and consolidation time);
Inclusion displacement (as a function of its rigidity and embedment stiffness);
Surcharges compared to the set of initial conditions;
Soil-inclusion interface conditions;
Boundary conditions (single inclusion vs. inclusion within a group).
56
The negative skin friction stress fn increases with relative soil-inclusion displacement Gy until
reaching a threshold value qs for relative displacement Gyc (on the order of a few millimeters),
hence a very small amount (Fig. 2.16). Given the consolidation process, this negative skin
friction fn is maximized in the long term and must be calculated on the basis of effective
stresses.
The methods typically implemented to evaluate the forces due to negative skin friction do not
explicitly introduce a mobilization law for these forces as a function of relative soil-pile
displacement; instead, these methods tend to be failure methods.
The relationship between effective vertical and horizontal stresses can be written as follows:
h .v (z)
(2.2)
with K being the earth pressure coefficient at the soil-pile boundary.
/HWV GHQRWH WKH soil-pile friction angle, which depends on both the type of pile and soil
characteristics. The shear stress due to negative skin friction is then expressed as:
fn(z) = K. v (z). Tan
(2.3.)
,QSUDFWLFHWKHSURGXFW.WDQLVFRQVLGHUHGDVDWHUPZLWKWZRLQVHSDUDEOHIDFWRUV
Negative skin friction may appear in cases other than surcharges on a soft soil, namely:
Natural or artificial soils (recent embankments) settling under their own weight;
Lowering of the water table;
Liquefiable soils or loose soils that may settle by either saturation or densification in the
presence of seismic effects.
A detailed description of the protocol for calculating negative skin friction around a pile (or
an inclusion), whether single or in a group, will be provided below (see Section 3 of this
chapter), with reference to the model proposed by Combarieu (1974, 1985).
2.3. Soil resistance mobilization
The resistant part of the inclusion located below the neutral point, also called the resisting
length, allows absorbing, through positive friction and tip resistance, the maximum inclusion
load at the neutral point.
Mobilization of the positive unit shaft friction depends on relative displacement between
inclusion and surrounding soil, should the inclusion displacement be greater than soil
displacement.
The resisting length of the inclusion will be calculated according to methods specifically
devised for deep foundations. The safety factors however will be deliberately selected at
Operating mechanisms
57
lower values (without applying a reduction factor to the creep load), given that a higher
displacement can be accepted for an inclusion than for a deep foundation.
2.4. Inclusion settlement
Loading tests have demonstrated that, like for piles, the vertical settlement of a single
inclusion bearing on a stiff layer is weak under a service load. Frank and Zhao (1982)
provided a method for calculating these settlements based on pressuremeter characteristics.
The settlement of an inclusion results from both its tip settlement and elastic shortening.
Elastic shortening of the various types of inclusions tends to be small in comparison with soft
soil settlement. Examples of characteristic deformation modulus values for inclusions are
given in Chapter 7 (Section 3.4).
As explained in Section 5 of this chapter, the settlement of a group of inclusions will be
calculated as the sum of two primary components: settlement of the actual reinforced soil
(including transfer platform settlement), combined with the settlement of non-reinforced deep
layers.
2.5. Loadings other than vertical and centered
2.5.1. Asymmetrical loading
In the case of large-sized foundations, the predominant vertical direction of loads to be
absorbed and the type of interaction mechanisms between soil and inclusions lead to primarily
compressive forces in the inclusions. Only those inclusions placed at the periphery of a loaded
zone would be able to accommodate additional bending loadings (i.e. the edge effect).
In the example of a small group of inclusions underneath a single footing (Fig. 2.17), the edge
effect that already exists in the presence of a vertical load becomes predominant when the
footing is exposed to a combined loading (i.e. transverse force and moment). A specific study
is needed to define the geometry and mechanical characteristics of the transfer platform that
yield a sufficient reduction in loadings in the inclusions so as to avoid, if applicable,
reinforcing the inclusions.
58
Under large-scale foundations, edge effects are to be taken into consideration, and this leads
to calculating the inclusions close to edges differently from those located in the central zones,
whose loadings tend more often to be uniform.
As an example, under tall embankments, bending loadings may appear at the heads of
inclusions placed below the slope, in addition to compressive loadings developed along the
embankment axis. This condition might justify the incorporation of reinforcement layers
(geosynthetics, geogrids) at the embankment base in order to reduce the horizontal
displacement and its influence on inclusion heads. The asymmetry of embankment loads leads
moreover to a non-uniform distribution of vertical forces within the inclusions.
2.5.2. Transverse loading
As opposed to footings placed on piles (where transverse forces are fully distributed at the
pile heads), the granular transfer platform absorbs a portion of deformations, which could
produce the beneficial impact of decreasing displacements at the inclusion heads.
The transverse forces exerted on the foundation are distributed, via the granular transfer
platform, between the inclusions and the soil.
Operating mechanisms
59
These lower forces applied at inclusion heads, when combined with the fact that inclusions
are free head, explain why the loadings induced in inclusions are typically significantly lower
than in the case of a footing installed on piles.
3. LOAD TRANSFER MECHANISMS
Load transfer mechanisms rely heavily on the model Combarieu (1974, 1985) proposed in
order to describe negative skin friction either around a single pile or on a group of piles.
In the case of reinforcement by rigid inclusions, this original model is only applicable to the
height of soft soil around the inclusions.
Combarieu subsequently proposed extending this model to the soil located above inclusion
heads in order to study load transfers within the platform, specifically in the following cases:
Embankments lying directly over the inclusions (piled embankments) (1988, 2007);
Load transfer platforms in the case of flexible ground slabs or rigid footings lying on soils
reinforced by inclusions (1990, 2007).
Other methods have also been proposed to study load transfer inside the platform, including
WKHGLIIXVLRQFRQH method.
3.1. Negative skin friction over the soft soil height
3.1.1. Single inclusion or single pile
The model was developed to calculate the negative skin friction acting on structural piles.
Conventional methods consist of estimating the critical height hc, which corresponds to the
inclusion length actually subjected to negative skin friction (Fig. 2.18), and then calculating
the intensity of this negative friction based on the following relation:
hc
FN
2 S rp
K tan G V'
( z , r p ) dz
(2.4)
in assuming WKDW WKH HIIHFWLYH YHUWLFDO VWUHVV v (z, rp) in the soil along the inclusion is not
disturbed by the presence of the inclusion itself and is identified with V1 (z), the effective
vertical stress in free field without inclusions.
60
Operating mechanisms
61
Combarieu (1985) went on to define a radial variation law for vertical stress at height z, by
introducing the notion of downdrag effect of soft soil around the inclusion (Fig. 2.19).
O
rp
V' v ( z , r ) V' v ( z ) V'1 ( z ) V' v ( z )1 e
(2.5)
where:
V1 (z) is the effective vertical stress in free field within the ground in the final state, as
calculated without taking into account the presence of the inclusion;
Vv (z) = Vv (z, rp) is the effective vertical stress at the inclusion contact in the final state,
while taking into account the downdrag effect;
Vv (z, r) is the actual effective vertical stress at a distance r from the inclusion axis, in
taking into account the disturbance due to the presence of the inclusion;
O is the so-called downdrag coefficient that characterizes the magnitude of the downdrag
effect around the inclusion.
The downdrag coefficient O is assigned the following values, all of which have been obtained
experimentally (Fig. 2.20).
62
1
0.5 0.25 K tan G
O
0.385 K tan G
if
if
if
K.tanG d 0.15
0.15 d K.tanG d 0.385
K.tanG t 0.385
The case O = 0, i.e. maximum downdrag effect, corresponds to the simplifying hypothesis of
zero reduction in vertical stresses due to the presence of inclusions.
The values of K.tanG, which are correlated with the type of inclusion or pile as well as with
the nature and characteristics of the surrounding soil, stem from full-scale experiments
conducted on deep foundations exposed to typically strong intensities of the final negative
skin friction that is necessary for a successful interpretation (only obtained after what are
often long consolidation periods).
Table 2.1 below lists the values of K.tanG proposed in 1985 for piles (which could also
constitute rigid inclusions), by focusing on those piles whose mode of execution is applicable
to current inclusion types of smaller dimensions. Table 2.2 provides the values of K.tanG
proposed in the NF P 94-262 Standard (Appendix H: Evaluation of negative skin friction on a
foundation element).
The various types of inclusions currently performed on sites have been introduced into this
table as well. It must nonetheless be noted that in justifying a group of inclusions, the major
part of forces is absorbed at the inclusion heads; forces absorbed in the soft soil by means of
negative skin friction are relatively weak, and the influence of the term K.tanG thus remains
limited.
Table 2.1: Recommended values of parameter K.tanG.
Piles or inclusions executed by displacement or under pressure,
in fine-grained soils with a liquid to soft consistency or in organic soils
*reduced to 0.15 for open elements
Piles or inclusions bored into fine-grained soils with a liquid to soft consistency
or into organic soils
*reduced to 0.10 for the lost casing method
Piles or inclusions executed by displacement or under pressure,
in fine-grained soils with firm to stiff consistency
*reduced to 0.20 for open elements
Piles or inclusions bored into fine-grained soils with firm to stiff consistency
*reduced to 0.15 for the lost casing method
Piles or inclusions executed by displacement or bored, in sands and gravels
0.20*
0.15*
0.30*
0.20*
0.35-1 (depending
on compaction
grade)
Organic soils
Soft soils
Firm / stiff soils
Very loose
Loose
Dense
Operating mechanisms
Bored piles
0.15
0.15
0.20
0.35
0.45
1.00
Driven piles
0.20
0.20
0.30
63
The principle of vertical force conservation in the model yields the following differential
equation, written using stressv (z) as the variable (Combarieu, 1985):
dV' v ( z )
P( O ) V' v ( z )
dz
with:
P( O )
dV'1 ( z )
dz
(2.6.)
O K tan G
1 O rp
(2.7)
The solution to this equation results in the general expression of effective vertical stressv(z)
at the soil-pile interface, over intervals where dV1(z)/dz can be assumed constant:
V' v ( z )
1 dV'1 ( z )
1 dV'1 ( z )
e P( O ) z V' v ( 0 )
P( O ) dz
P( O ) dz
(2.8)
In the special case of a homogeneous, saturated soft soil layer with groundwater at the surface
and subjected to a uniform surcharge q0, dV1(z)/dz is constant:
dV'1 ( z )
dz
J'
J'
P( O )z
e
q 0
P( O )
P( O )
(2.9)
and the height hc over which negative skin friction is exerted, i.e. from the inclusion head to
the neutral point as determined E\WKHFRQGLWLRQv (hc) = JKc, verifies the relation:
P (O )hc 1
e P ( O ) hc
q
P (O ) 0 1
J'
which, as a preference, is to be solved graphically.
If the calculated hc value is less than H (i.e. the soft soil thickness), then the negative skin
2 S rp K tan G
friction resultant is written as: FN
q0 ,
P (O )
meaning that proportionality exists between FN and q0.
If the calculated hc value is greater than H, then FN is to be applied over the entire soft soil
thickness H, with the negative skin friction resultant being expressed as:
2 S rp K tan G
FN
>J ' hc q0 V 'v (hc )@
P (O )
ZKHUHv (hc) is given by:
J'
J'
P (O )hc
V 'v (hc )
e
q0
P (O )
P (O )
It is shown that FN, which reaches its maximum value at the end of the consolidation period,
is very simply correlated with the degree of consolidation whenever hc is less than height H.
64
ab
S
Operating mechanisms
65
Figure 2.23 schematically illustrates the negative skin friction model being studied along with
its associated notations.
This analysis results in a differential equation identical to that in (2.6), which makes it
possible to determinev (z, rp) (Combarieu, 1985):
dV' v ( z , r p )
dz
with:
P( O , R )
or:
P( 0 , R )
dV'1 ( z )
dz
P( O , R ) V' v ( z , r p )
O
O R O
e
1 O 1
r
p
R
r
p
K tan G
rp
1
R rp
rp
K tan G
rp
(2.10)
if O z 0
(2.11)
if O 0
(2.12)
These relations are also given in the form of an abacus (Fig. 2.24).
66
dV'1 ( z )
(homogeneous at a unit weight)
dz
can be considered as constant, i.e. cases in which the profile V1 (z) can be linearized by layer.
Both the critical height determination and total negative skin friction force calculation are
carried out in the same manner as for a single inclusion. A number of software programs have
been developed for this purpose.
The vertical stress profileEHWZHHQLQFOXVLRQVv (z, r) is obtained by solving (2.10), which is
a first-order linear differential equation. This solution uses as input the vertical stress profile
V1 (z) that would be obtained without any inclusion and the values of the two coefficients
K.tanG and O, which serve to characterize the downdrag effect around the inclusions (by
which vertical stress at the inclusion contact v (z, rp) decreases).
The negative skin friction along the inclusion is expressed like for the single inclusion, with
(O) being replaced E\O, R/rp), where O) corresponds to the limit derived for R/rp =
In assuming, as is the case with older conventional methods, that the vertical stress1 (z) is
not disturbed by the presence of rigid elements, this would be equivalent to a maximum
downdrag effect O = 0, which does not prevent the group effect from intervening. Integrating
equation (2.10) yields the solution provided by Zeevaert in 1957, which constitutes a specific
solution of the proposed more general solution.
The average vertical stress v*(z) between inclusions is expressed, over the height where
negative skin friction is applied, as follows:
>
2 K tan G r p
r p 2 P( O ,
R
)
rp
(2.13)
The critical height hc is often determined by adopting the hypothesis that negative skin
friction only acts if the stress in contact with WKHLQFOXVLRQv (z, rp) is greater than the initial
stress Vv0 (z). The negative skin friction force acting on the inclusion can then be calculated
by the expression in (2.8) over the height hc or H.
Simon (2001) suggested an alternative whereby negative skin friction only develops upon
reaching a depth where the average vertical stress between inclusioQVv*(z) remains greater
than or equal to the initial stress Vv0 (z).
Combarieu's approach is based on ultimate stresses, where inclusions are only slightly
deformable, while the soil settlement is sufficient for shear stresses in contact with the
inclusions to reach their ultimate values W (z); these ultimate values are obtained for relative
displacements between soil and inclusion (or pile) on the order of 0.01 D (where D is the
element diameter). They stem from experimental tests on negative skin friction.
In order to verify method applicability, a study of deformations without inclusions must be
performed.
More specifically, this method produces:
The shear stress at all points on the inclusion shaft;
W (z) = K.tanG. v (z, rp);
The average vertical stressEHWZHHQLQFOXVLRQVv*(z), as used in the settlement calculation;
depth hc of the neutral point.
Operating mechanisms
67
1
2
FN ( R) FN (f)
3
3
FNe
2
1
FN ( R) FN (f)
3
3
Figure 2.25: Layout of an inclusion in a single file.
FNi
FN (R)
FNa
7
5
FN ( R ) FN (f) (a)o corner
12
12
FNe
5
1
FN ( R) FN (f)
6
6
(i)o interior
(e)o exterior
plane passing at this level ("neutral plane of equal settlement"), friction is positive and
contributes to increasing the load transmitted to the soil between inclusions by means of
reducing the force being supported by these specific inclusions. The resulting settlement of
the surrounding soil can therefore only be estimated in an approximate manner based on a set
of complementary hypotheses;
Soil stiffness under the inclusion tip and deep layer stiffness also prove to be major factors
since they determine the magnitude of differential settlement between the inclusion tip and
surrounding soil, thereby justifying the implicit hypothesis of Combarieus method, according
to which relative soil-inclusion displacements in a standard cross-section are of sufficient
magnitude for the negative skin friction to reach its ultimate values.
3.2. Load transfer in the platform
Among the various analytical methods available to evaluate load transfer in the platform, the
ASIRI National Project forwards the two following tested methods:
The fictitious inclusion method (developed by Combarieu), which offers the advantage of
providing a homogeneous approach consistent with the negative skin friction evaluation
method described above, based on the same set of principles (downdrag effect on fictitious
columns in the embankment);
The diffusion cone method, which entails an approach compliant with the mechanisms
exposed during the various experiments and modeling exercises conducted within the scope
of the ASIRI National Project (i.e. load transfer by diffusion above the inclusion), as detailed
in Section 1 of this chapter.
These two methods and their limitations will be discussed below.
3.2.1. The fictitious inclusion method
3.2.1.1. Description
Combarieu (1988) proposed modeling the phenomena involved in load transfer above the
inclusion heads by assimilation with the negative skin friction that would have developed, had
these inclusions been fictitiously extended over the same height hr as that where this negative
friction has been exerted (Fig. 2.27). Parameter hr is thus equal to the thickness HR of the
embankment or load transfer platform supporting a flexible structure subjected to a uniform
load of intensity q0.
In the special case of cohesionless materials, it is suggested to set the value of .WDQr = 1
r: embankment friction angle), yielding the value O = 0 for the downdrag coefficient.
The vertical stress in the soil between inclusions is then uniform and given by:
q S
with:
Pr
Jr
( 1 e P r hr ) q 0 e P r hr
Pr
(2.14)
2 rp K tan M r
R rp
where:
rp: inclusion radius,
R: equivalent mesh radius (or 2R: spacing between inclusions).
Operating mechanisms
69
In the special case of embankments made with cohesionless materials, this expression is the
same as that exposed by Terzaghi in his study on the silo effect for sands.
In its general form, r is also a function of O (directly related to K.tanG), the downdrag
coefficient introduced for negative skin friction and set equal to zero for a high-quality
embankment (granular material).
The expression yielding qs+ makes it possible to study the influence of the various geometric
parameters (rp/R, HR/R) on system efficiency. Moreover, in the calculation performed in
accordance with Combarieu's method, it has been verified that the load absorbed by an
inclusion is bounded by the load value corresponding to its zone of influence.
By applying Combarieu's model to the embankment with fictitious elements featuring the
same cross-section as the inclusions in soft soil, a comprehensive approach can be established
for the design of reinforcement by rigid inclusions, through setting an objective of minimizing
the stress applied on the soft soil.
From a general standpoint, it is sought to limit the stress qs+ on the soil between inclusions,
with the aim of limiting settlements. By setting an appropriate value for the stress reduction
rate (SRR) = qs+ / Jr.hr, it becomes possible to determine an acceptable mesh geometry for a
fixed inclusion radius rp. Each inclusion is then subjected to the following force at the head:
(FN + S rp2.Jr.hr), where FN is the negative skin friction developed over height hr.
3.2.1.2. Limitations and improvements introduced
The original method proposed by Combarieu in 1988 to treat the fictitious inclusion presented
the disadvantage of leading to stress reduction rate (SRR) increasing with embankment height
HR, which contradicts both the observations recorded on reduced-scale models and the
theoretical results established by Hewlett or Kempton (which indicate that this rate actually
reaches a threshold value vs. increasing embankment heights).
70
In reality, the method to be applied is based on the position of an upper plane of equal
settlement in the embankment, and not based on the upper surface position of the
embankment.
For the work carried out in the ASIRI National Project, the method has been modified
(Combarieu, 2007, 2008) as regards its extension to the calculation of forces passing through
the embankment above the inclusion heads. For this purpose, an empirical formulation has
been introduced to provide the active height ha between inclusion heads and the upper plane
of equal settlement (Fig. 2.27), vs. radius R of the equivalent circular mesh, i.e.:
ha
R
R
10 0.4(6 )
if
<6
rp
rp
rp
and
ha
rp
10
if
R
!6
rp
It is also proposed to evaluate the coefficient K.tanG in the embankment by means of the
following empirical relation:
K.tanG = 1.1 eE/E0
(2.15)
based on the deformation modulus E of the layer of thickness ha (with E0 = 50 MPa).
This modulus value E is equal to the Ev2 modulus measured at the 60-cm diameter plate,
which may also be estimated from correlations with other deformability measurements.
The residual stress between inclusions can now be written under these conditions:
qs
q( ha )
Jr
( 1 e P r ha ) J r ( hr ha )e P r ha
Pr
(2.16)
e P r ha as hr increases indefinitely.
Introduction of the plane of equal settlement causes a rise in qs+ and, consequently, a smaller
force at the inclusion head. Conversely, negative friction along the shaft, generated by qs+
using a consistent mode of calculation, subsequently rises. On the whole, for efficient groups
of inclusions, this situation leads, relative to the 1988 proposals, to just a slight drop in
efficiency.
In the case of ground slabs, the upper plane of equal settlement can be related to the ground
slab underside.
3.2.1.3. Evaluation
The fictitious inclusion model is capable of describing load transfer in the platform, provided
taking into account the specific points raised in the previous section (e.g. identification of an
upper plane of equal settlement). Chapter 3 will detail the requirements inherent in analytical
type calculation models.
Operating mechanisms
71
72
Based on the proposed geometry (Fig. 2.29), a determination can now be made of the share of
surcharge Qp and weight Wp of the granular layer being redirected towards the inclusions via
the load transfer zone. Both the efficiency E and capacity G of the granular transfer platform
relaying surcharges towards the inclusions can also be deduced in the case of square-section
inclusions or pile caps.
for hm d h*
Qp
Wp
Wp Qp
and
s J hm q
2
sa
2 tan T
for h m d h *
sa
(2.18)
2 tan T
(2.17)
(2.19)
For circular-section inclusions or pile caps, with diameter a, the following are obtained:
> a / 2
Wp
J S hm
3
Qp
q S a / 2 hm tan T
a / 2 hm tan T 2 a a / 2 hm tan T / 2
2
Wp Qp
s 2 J hm q
and
for hm d h*
G
sa
2 tan T
S a / 2 hm tan T
s2
(2.20)
(2.21)
(2.22)
Angle T is typically set equal to the peak friction angle Mp of the granular material in order to
assess the maximum efficiency capable of being mobilized for a given loading.
In comparison with numerical simulations using discrete elements run for a load transfer
platform of thickness hm = 1.0 m and 0.5 m, it is observed in Figure 2.30 that the proposed
analytical formulations serve to approximate, in a satisfactory manner, the range of load
Operating mechanisms
73
transfer mechanisms. For small granular layer displacements, the friction angle to be
considered in the analytical formulations is indeed the peak friction angle of the granular
material introduced. For greater deformations however, it could be assumed that particle
disorganization would require adopting smaller friction angles.
Figure 2.30: Comparison between load efficiencies E and G obtained by discrete numerical
simulations and analytical calculations (Chevalier et al., 2010)
4. BASIC MECHANISMS
4.1. Overall mechanism of reinforcement by rigid inclusions
/HWs recall that in comparison with foundations on piles or a reinforcement by flexible
inclusions of the stone columns type, a foundation soil reinforcement by rigid inclusions
features as its primary mechanism the distribution of loads between the inclusions and the
surrounding soft soil while minimizing settlements. This mechanism is also observed with
stone columns, though with the difference that stone columns produce equal settlements at the
column head and at the top of the surrounding soil.
Rigid inclusions thus serve to:
Channel the stresses due to embankments, loads on ground slabs or rafts, towards the deeper
stiff soil strata;
Reduce stresses in the soft soil in order to limit settlements;
Eventually shorten consolidation time by lowering stresses in the soft soil, depending on its
overconsolidation ratio.
4.2. Behavior of the single rigid inclusion
The single rigid inclusion functions like a small pile. During loading of the rigid inclusion
head, shaft friction is quickly mobilized and then the process occurs more slowly for the tip
until reaching the ultimate resistance conventionally set at a head displacement equal to onetenth the inclusion diameter (Fig. 2.31).
74
Figure 2.31: Mobilization of total and tip resistance for a rigid inclusion.
As regards embedding the inclusion in layers beneath the soft soil layers, it would be
necessary to differentiate two very distinct cases:
Domain no. 1: Inclusions are required for the stability and load-bearing capacity of the
structure. In this case, prescriptions from the Standard NF P 94-262 "Deep foundations" need
to be applied with, in general, a minimum embedment of 3 diameters for piles whose diameter
is less than 0.50 m. This lower embedment limit may be reduced to the minimum value of
0.50 m, provided the foundation contractor is able to guarantee execution of this pile tip
embedding operation within the load-bearing layer by either extracting samples, or
introducing a trepan or using a core sampler.
Domain no. 2: The rigid inclusions are primarily intended to decrease settlements and are
not required to justify the stability or load-bearing capacity of the structure. From a general
standpoint, a limitation will be imposed of a minimum one-diameter embedment in the loadbearing layer, provided a relevant evaluation has been performed of the load-bearing factor
(with respect to Standard NF P 94-262) along with a strict execution quality control (such as a
systematic recording of installation parameters), thus making it possible to guarantee
generating this minimum embedment for all inclusions. In the absence of a true load-bearing
layer, so-FDOOHG IORDWLQJ inclusions may be envisaged, albeit subjected to the condition of
specific justifications.
4.3. Behavior of a rigid inclusion within a group
Under the effect of soft soil settlement around grouped rigid inclusions, a negative skin
friction develops over the upper part of each inclusion; this friction increases with either the
embankment height or the surcharge. In this part, the vertical displacement of the neighboring
soil is greater than that of the inclusion itself. In contrast, over the lower part, the friction is
positive and moreover at each level, the vertical displacement of the rigid inclusion is greater
than that of the surrounding soil. A level of inclusion therefore exists at which the
displacements of both the inclusion and surrounding soil are identical; this level is referred to
as the QHXWUDO SODQH and corresponds to where the shear stress due to shaft friction equals
zero, and the normal force is maximized in the inclusion.
Two other neutral planes, located respectively above the inclusion heads and below the
inclusion tips, are typically defined as well (see Fig. 2.32):
Operating mechanisms
75
Figure 2.32: Position of the 3 neutral planes determined from the shear stress profile
along the inclusion.
The upper neutral plane is the most interesting to study, since above it the settlements are
uniform and, as such, the upper embankment surface will not be subjected to egg carton
type deformations (Fig. 2.33).
The lower neutral plane is located beneath the inclusions in the bearing soil, at a level where
soil settlement becomes uniform once again.
76
It thus proves necessary to introduce a thick enough transfer platform to avoid differential
displacements at the top of an embankment or hard points under a ground slab. An
insufficient transfer platform thickness leads, in the case of an embankment, to failure within
a soil volume above the inclusion head penetrating the surface (compare Figs. 2.35 and 2.36)
or alternatively, in the case of a ground slab, to high stresses inside it.
Operating mechanisms
77
Figure 2.37: Settlement curves for an overconsolidated soil both with and without reinforcement.
In addition to load efficiency E (as defined in Section 1.1 of this chapter), it is common
practice for engineers to consider a settlement efficiency Etass that specifically expresses the
capacity of the process to satisfy the settlement criteria required for the structure. This
efficiency is established via the expression:
Etass = 1 yM / y0
(2.23)
with:
yM: settlement of the soil reinforced by rigid inclusions, as evaluated or measured at the
surface of the load transfer platform;
y0: settlement of the virgin soil (i.e. without inclusions), as evaluated based on site
investigation data.
The reinforced soil settlement may be:
Either calculated based on data derived from site investigation results and the calculation
methods proposed in these Recommendations (see Chapters 3, 4 and 5);
Or measured using appropriate instrumentation implemented within the scope of the control
testing campaign.
The settlement of a shallow foundation such as an embankment, footing, ground slab or raft
on soil reinforced by rigid inclusions is defined as the sum of two components, i.e.:
The actual settlement experienced by the reinforced soil, including the settlement of the
load transfer platform;
The settlement of deeper, non-reinforced layers.
5.1. Settlement of reinforced soil
A reinforcement by rigid inclusions is reflected, and such is the intended purpose, by a
reduction (most often of large magnitude) of settlements over the entire reinforced soil height;
this reduction depends on both the reinforcement mesh and characteristics of the granular
transfer platform (thickness, deformability and eventual presence of a structural additive like
a geogrid).
These settlements could be lowered thanks to a substantial transfer of applied load onto the
inclusions that occurs at two levels, as already described in the previous sections, namely:
A load concentration at the inclusion head resulting from shear mechanisms in the transfer
platform and capable of leading, due to the induced stress rotation, to the creation of force
arches through which forces can be channeled, and possibly complemented by permanent
cohesion of the platform constituent material or by the membrane effect of geogrid(s) laid out
on the inclusion heads;
A negative friction effect of the soft soil acting on the inclusion, according to the set of
mechanisms analyzed by O. Combarieu, as governed by parameters including .WDQDQGO,
in combining the group and downdrag effects.
This load transfer yields a significant drop in the additional stress applied at the top of the soft
soil and inside it, hence leading to a very noticeable decrease in settlement, which after all is
the intended goal.
In recognizing that a sizable portion of the load is transferred onto the inclusions and
moreover that these inclusions display relatively high stiffness, final settlement reduction will
be even more pronounced if the inclusion tip rests on a load-bearing resistant layer such as a
stiff substratum or a layer of sand or gravel of adequate thickness.
Operating mechanisms
79
In this respect, it would be necessary to accurately distinguish the case of inclusions stopping
in the soft soil layer, which can be qualified as floating inclusions, from the case of inclusions
resting on a resistant layer.
With the aim of optimizing reinforcement efficiency in terms of settlement reduction, it is
highly recommended, in common cases, to allocate reinforcement over the entire soft soil
layer height and then embed the inclusions on a load-bearing level, while acknowledging that
relatively short floating inclusions would display a near-zero efficiency.
In the presence of organic soils (peat, household waste, etc.), the residual surcharge applied
on the soil could cause creep settlements and thereby limit the phenomenon of downdrag.
5.2. Settlement of deeper, unreinforced layers
Depending on the mechanism described above, inclusions play the role of deep-seated load
transfer elements. This surcharge, transferred at the level of the inclusion base, winds up
causing settlements in the underlying layers (remaining in a virgin state given their lack of
reinforcement), in the same way as settlements potentially occurring under a group of piles
(even though the transfer mechanism is substantially different in the case of piles) (Fig. 2.39).
Figure 2.39: Necessity of verification for the load transfer influence under the inclusion tip.
It should nonetheless be pointed out that these deep layers tend to be relatively
incompressible, meaning that settlements induced at the surface are small and often uniform
at the scale of the structure examined.
Yet this examination of settlement within the deeper layers must not be overlooked.
5.3. Geotechnical investigations
As regards site investigations, it will be necessary to plan a series of boreholes with suitable
laboratory or in situ testing to allow evaluating soil compressibility characteristics. This will
require reaching a sufficient depth, relative to the targeted reinforcement area and depth, and
with a sufficient number of borings to determine the potential variations in these
characteristics at the scale of the structure, with the aim of assessing both total and differential
settlements.
80
The reader is referred to Chapter 6 for a detailed description of the relevant boreholes and
tests.
5.4. Determinant parameters
As a complement to in situ soil compressibility characteristics, the following key parameters,
which will be showcased in the calculation methods described in Chapters 3, 4 and 5, are
determinant to settlement evaluations:
Granular transfer platform characteristics (thickness, deformability, friction angle, eventual
cohesion), including if applicable the characteristics of additives used as reinforcement layers;
Inclusion characteristics (head diameter and diameter of the standard cross-section,
deformation modulus, length compared to soft soil thickness);
The area ratio, i.e. ratio of the head cross-sectional area to the element mesh area.
Searching for the best combination of these parameters allows deriving an optimal design that
satisfies the settlement criteria.
6. LESSONS FROM THE FULL-SCALE EXPERIMENTS
As a complement to the description of fundamental mechanisms, it seems important to
highlight a number of observations or lessons from the two full-scale experiments (SaintOuen-OAumne and Chelles) conducted within the scope of the ASIRI National Project.
These lessons are deemed to be particularly useful in producing an effective design or
carrying out a realistic preliminary design of reinforcement by rigid inclusions during the
early project design stage.
6.1. Inclusion behavior
As for inclusion head loads and settlements, the Saint-Ouen-OAumne (Fig. 2.40) and Chelles
(Fig. 2.41) experiments demonstrate that the behavior of an inclusion underneath the structure
(embankment, ground slab) is the same as that of the tip of a single rigid inclusion subjected
to a loading test. This outcome reveals that the load exerted at the head of an inclusion is
close to that exerted in the vicinity of its tip.
Such a finding expresses the fact that the positive and negative friction resultants along the
inclusion shaft are more or less balanced, and that consequently inclusion efficiency is
basically governed by its load-bearing capacity at the tip. These results expose the need to
determine this capacity and then accurately model the inclusion tip behavior in order to obtain
realistic results at the top of a structure on soil reinforced by rigid inclusions.
This result can be considered as valid solely for groups of inclusions lying on a resistant
substratum (and in theory does not apply to the floating inclusions).
Operating mechanisms
81
82
inclusions may not be optimal, given that it is capable of constraining this friction-based
transfer mechanism.
The step of cutting off inclusions at the base of the granular transfer platform could reduce
load transfer in that the soil aligned with the inclusion might not be compacted like the rest of
the load transfer platform. In this respect, appropriate arrangements are required in order to
guarantee the homogeneity of transfer platform compaction (see Chapter 7: Execution).
6.3. Reinforcement layers (geosynthetic)
The Chelles experiment (2007) indicates that the state of stress above a granular transfer
platform reinforced by geosynthetic layers differs depending on the type and number of layers
in the configuration:
Case of the load transfer platform reinforced by a layer of geotextile, i.e. composed of a
rather deformable material: the geotextile deforms during embankment construction, and this
deformation is concentrated around the inclusion heads;
Case of the load transfer platform reinforced by a geogrid layer, i.e. composed of a less
deformable material: the geogrid deforms in a uniform pattern during installation of the
granular platform; afterwards, the deformation is also concentrated around the inclusion heads
during embankment construction. The geogrid tends to prestress the load transfer platform
prior to its loading by the embankment, which yields a certain amount of cohesion. The
interpretations drawn from the Chelles full-scale experiments served to quantify this
additional equivalent cohesion;
The number of layers introduced is also a key factor in that it contributes to increasing
transfer platform stiffness.
6.4. Settlement efficiency
The reinforcement by rigid inclusions significantly reduces settlement both at ground level
and throughout the soil depth.
In the Chelles instrumented test embankment, it has been observed that settlement occurs for
the most part during the construction stage, which therefore makes it necessary to impose
settlement criteria with reference to expected settlements after embankment construction and
not before.
The efficiency calculated for stresses at the level of inclusion heads is not always the most
pertinent design parameter; instead, the rationale should target reducing settlements. As an
illustration, in the case of the Chelles embankment, it can be remarked that a 20% stress
efficiency has yielded a 60% settlement reduction!
7. SEISMIC LOADINGS
The objective of a reinforcement step using rigid inclusions is to introduce sufficient
mechanical characteristics into the reinforced soil so as to allow building a shallow
foundation (or perhaps a raft or ground slab) or an embankment under fully acceptable
stability and settlement conditions.
Operating mechanisms
83
In the case where the reinforced soil is subjected to seismic loadings, two very distinct
domains need to be distinguished (the AFPS Soil improvement and reinforcement processes
under seismic action Technical Guide, 2011):
Domain no. 1: Inclusions are necessary for structural stability and for load-bearing capacity,
with verifications of the ultimate limit states (GEO). The next step requires guaranteeing the
same level of inclusion strength during an earthquake as that taken into account in design
calculations, thus justifying that these inclusions remain within the elastic domain
(justifications comparable to current rules in effect for piles).
Domain no. 2: The rigid inclusions are primarily intended to reduce settlements and are not
required to justify the structural stability or load-bearing capacity. The absence of structural
failure must then be justified when exposed to seismic action by neglecting either part or all
of the inclusions, in a way that is not detrimental to human safety in the event of structural
damage.
Remark: The situation may be encountered where the Project Owner imposes that
the structure still be in working order after the earthquake, i.e. without any
physical disorders and operational limitations appearing either during or after the
earthquake [EN 1998-1 Standard, Section 2.1 (1) P]; in this case, the inclusions
would enter into domain n 1.
The granular transfer platform must be able to perform the following functions:
Facilitate energy dissipation by means of sliding in the granular medium. Accordingly, the
load transfer platform plays the role of a fuse triggering a controllable sliding, rather than
footing rotation via a loss of load-bearing capacity. A hierarchy in the strength of various
system components is thus introduced, and this step naturally assumes the absence of a plastic
zone outside of the transfer platform. This finding implies that rigid inclusions do not fail and
therefore prove to be correctly designed or even overdesigned;
Reduce loadings in the inclusions (decrease in both shear force at the inclusion head and
lateral displacement of the soil over the upper part of the rigid inclusions);
Perform a filtering effect by avoiding any seismic loading amplification, or even in some
cases by reducing inertial forces. This reduction phenomenon is closely correlated with
energy dissipation in the load transfer platform by friction. The intensity of this phenomenon
depends on transfer platform thickness;
Ultimately, increase soil load-bearing capacity. The load transfer platform may be sized to
make the transition from domain no. 1 to domain n 2.
Both shear and normal stresses will develop in the rigid inclusions when exposed to kinematic
and inertial effects. In domain no. 1, where inclusions are necessary for structural stability
(and load-bearing capacity), it will be mandatory to verify that all inclusion sections remain
compressed under the induced loadings and moreover that normal and shear stresses continue
to be acceptable (refer to Chapter 5: Justifications).
The earthquake resistance calculation for foundation soils reinforced by rigid inclusions is
complicated by the very complex feature of the soil-inclusion-platform-foundationsuperstructure interaction under cyclic loading.
Incident waves induce deformations in the soil, which then transmits these deformations to
the foundations and subsequently to the structure. The interaction present is thus very strong
and it decomposes into a kinematic interaction and an inertial interaction:
84
The soil deformed by incident waves forces the foundations, and consequently the
superstructure, to follow its movement (i.e. a kinematic interaction). Even in the absence of
superstructure loads, foundations will exhibit different displacements than the free field soil
displacements due to differences in stiffness between foundation and soil;
The superstructure mobilized by the foundation movement will develop inertial forces that
serve to induce soil-foundation system forces (i.e. an inertial interaction).
The two mechanisms occur simultaneously; the soil-foundation-structure (the soil-structure
interaction) system response can be obtained by simply combining these mechanisms. The
need to take into account a combination of both mechanisms depends on: the type of soil, the
seismic zone, and the category of structure (as referred to in Section 5.3 of the AFPS Soil
improvement and reinforcement processes under seismic action Technical Guide, 2011).
In liquefiable zones, the liquefaction potential may be reduced by use of rigid inclusions
under the eventually combined effect of the following actions:
If inclusions are being used to increase density of the surrounding soil, e.g. by soil
displacement and/or vibration, then the gain in density following treatment would need to be
verified by in situ tests conducted between inclusions;
Should inclusions be used to lower shear stress in the soil, then the design must be carried
out according to Hashins homogenization method (1983), as detailed in the AFPS Soil
improvement and reinforcement processes under seismic action Technical Guide (2011). Any
method that consists of increasing average shear strength of the treated mesh by incorporating
inclusion strength would not be valid. This approach imposes verifying inclusion integrity
(i.e. both normal and shear stresses must remain acceptable). The practical application of this
design method typically only allows for a limited increase, on the order of 5% to 10%, in the
safety factor relative to liquefaction.
The AFPS Soil improvement and reinforcement processes under seismic action Technical
Guide (2011) will be consulted for:
A review of the appropriate analytical methods;
Indications on the role played by a group of inclusions in a potential soil liquefaction.
Operating mechanisms
85
References
Internal reports (ASIRI National Project)
Baudouin G. - Sols renforcs par inclusions rigides : modlisation physique en centrifugeuse
de remblais et de dallage. Thse de doctorat, universit de Nantes, 27/10/2010.
Brianon L. 5HQIRUFHPHQW GHV VROV SDU LQFOXVLRQV ULJLGHV eWDW GH Oart en France et
OpWUDQJHU,5(;RSpUDWLRQGX5pVHDXJpQLHFLYLOHWXUEDLQ, 2002.
Brianon L. -5DSSRUWILQDOGHOexprimentation de Saint-Ouen-O$XP{QH . Rapport 1-07-102 et annexes numriques, juillet 2007.
Brianon L. - 5DSSRUW ILQDO GH OH[SpULPHQWDWLRQ GH &KHOOHV . Rapport 2-08-1-05, juillet
2008.
Brianon L. Complment benchmark tranche 1 . Thmes 1 et 4. Rapport 1-07-1 et 4-02,
novembre 2007.
Chakroun S., Simon B. - Simulation du matriau Schneebeli dans Plaxis . Rapport 1-06-402, dcembre 2006.
Chevalier B. - tudes exprimentale et numrique des transferts de charge dans les matriaux
granulaires. Application aux renforcements de sols par inclusions rigides. Thse de
doctorat, universit Grenoble I Joseph-Fourier, 5/09/2008. http://tel.archivesouvertes.fr/docs/00/36/73/05/PDF/These_Bastien_Chevalier-version3.pdf.
Chevalier B., Combe G., Villard P. - Modlisation discrte : tude du report de charge .
Rapport 3-08-4-01, 2008.
Chevalier B., Combe G., Fantino R., Grange S., Villard P. - tude des mcanismes de report
de charge dans les matelas granulaires des remblais sur sol compressible renforc par
inclusions rigides . Rapport 4-10-4-02, juillet 2010.
Combarieu O. - Remblais sur sol mdiocre et inclusions rigides. Nouvelle approche du
dimensionnement . Rapport 1-07-5-01, 2007.
Dinh A.Q. - tude sur modle physique des mcanismes de transfert de charge dans les sols
renforcs par inclusions rigides. Application au dimensionnement. Thse de doctorat,
ENPC, 2009.
IREX Ple Comptence Sols. Utilisation des inclusions rigides pour le renforcement des
sols de fondation dRXYUDJHV HW GH UHPEODLV eWXGH GH IDLVDELOLWp GXQ SURMHW QDWLRQDO ,
2000.
Jenck O. - Le renforcement des sols compressibles par inclusions rigides verticales.
Modlisation physique et numrique. Thse de doctorat, INSA Lyon, 29/11/2005.
Jenck O., Dias D., Kastner R. - Exprimentations sur modle rduit bidimensionnel en
laboratoire. Modlisation physique de la plate-forme de transfert de charge . Rapport 106-3-01, octobre 2006.
Jenck O., Dias D., Kastner R. - Modlisation numrique bidimensionnelle du modle de
ODERUDWRLUH GpYHORSSp j O,16$ HW pWXGHV SDUDPpWULTXHV . Rapport 1-06-4-01, octobre
2006.
Jenck O., Dias D. - 0RGpOLVDWLRQ QXPpULTXH GLVFUqWH GXQ PRGqOH UpGXLW ELGLPHQVLRQQHO GH
matelas de transfert de charge granulaire difi sur inclusions rigides. Rapport 2-08-4-03,
octobre 2007.
Koscielny M., Brianon L., Dias D. Synthse benchmark tranche 1. Thmes 1 et 4 .
Rapport 1-07-1 et 4-01, septembre 2007.
Laurent Y. Renforcement des massifs de fondation par inclusions rigides verticales. tude
bibliographique et numrique. Mmoire de DEA, INSA Lyon, 17/09/2002.
86
Le Hello B. - Renforcement par gosynthtiques des remblais sur inclusions rigides. tude
exprimentale en vraie grandeur et analyse numrique. Thse de doctorat, universit
Grenoble I Joseph-Fourier, 26/06/2007.
Okyay U.S. - tude exprimentale et numrique des transferts de charge dans un massif
renforc par inclusions rigides. Application des cas de chargements statiques et
dynamiques. Thse de doctorat, INSA Lyon, 24/11/2010.
Rault G., Baudouin G., Thorel L. - Essais centrifugeuse FRPSRUWHPHQW GXQ JURXSH
lmentaire (3*3 inclusions) : conception des montages (1) . Rapport 2-07-3-05,
dcembre 2007.
Rault G., Baudouin G., Thorel L. - Essais centrifugeuse FRPSRUWHPHQW Gun groupe
lmentaire (3*3 inclusions) : ralisation des montages (2) . Rapport 2-08-3-06, mai
2008.
Rault G., Thorel L. - tude du transfert de charge par cisaillement. Dispositif de plateau
mobile. Essais de rception . Rapport 3-09-3-14, octobre 2009.
Thorel L., Rault G., Baudoin G. - &RPSRUWHPHQWGXQJURXSHpOpPHQWDLUH
LQFOXVLRQV :
VLPXODWLRQHQFHQWULIXJHXVHGXFKDUJHPHQWGXQUHPEODL . Rapport 2-07-3-08 et annexes,
novembre 200).
External references (publications and standards)
AFPS 3URFpGpVGDPpOLRUDWLRQHWGHUHQIRUFHPHQWGHVROVRXVDFWLRQVVLVPLTXHV , 2011.
Berthelot P., Besson C., Boucherie M., Carpinteiro L., Deryckere N., Frossard A., Glandy M.,
Pezot B., Poilpre C., Volcke J.-P. - Recommandations sur la conception, le calcul,
OH[pFXWLRQ HW OH FRQWU{OH GHV FRORQQHV EDOODVWpHV VRXV EkWLPHQWV HW RXYUDJHV VHQVLEOHV DX
tassement . Revue franaise de gotechnique n 111, 2004, p. 3 16.
British Standard BS 8006 - &RGH RI SUDFWLFH IRU VWUHQJWKHQHG UHLQIRUFHG VRLOV DQG RWKHr
ILOOVVHFWLRQGHVLJQRIHPEDQNPHQWVZLWKUHLQIRUFHGVRLOIRXQGDWion on poor ground,
1995, p. 98-121.
Cartier G. - Effet de groupe sur le frottement ngatif (essais en centrifugeuse). Mmoire de
DEA, LCPC Nantes, 11/07/1994.
Chevalier B., Villard, P., Combe G. - Investigation of load transfer mechanisms in
geotechnical earth structures with thin fill platforms reinforced by rigid inclusions.
International Journal of Geomechanics, doi:10.1061/(ASCE) GM.1943 5622, 0000083,
2011.
CLOUTERRE 1991 (+ Additif 2002) Recommandations pour la conception, le calcul,
OH[pFXWLRQHWOHFRQWU{OHGHVVRXWqQHPHQWVUpDOLVpVSDUFORXDJHGHVVROV.
Combarieu O. - (IIHW GDFFURFKDJH HW PpWKRGH GpYDOXDWLRQ GX IURWWHPHQW QpJDWLI .
Bulletin de liaison des laboratoires des ponts et chausses n 71, mai-juin 1974, p. 93107.
Combarieu O. - Frottement ngatif sur les pieux. Rapport de recherche des Laboratoires
des ponts et chausses n 136, octobre 1985, 151 p.
Combarieu O. - Amlioration des sols par inclusions rigides verticales. Application
OpGLILFDWLRQ GHV UHPEODLV VXU VROV PpGLRFUHV Revue franaise de gotechnique n 44,
1988, p. 57-79.
Combarieu O. &DOFXOGune fondation mixte .1RWHGLQIRUPDWLRQWHFKQLTXHPLQLVWqUHGH
OeTXLSHPHQWHWGX/RJHPHQW/&3C, 1988.
Combarieu O. - Fondations superficielles sur sol amlior par inclusions rigides
verticales . Revue franaise de gotechnique n 53, 1990, p. 33 44.
Operating mechanisms
87
Simon B. - 8QH PpWKRGH LQWpJUpH SRXU GLPHQVLRQQHU OHV UpVHDX[ GLQFOXVLRns rigides en
dformation . XVe Congrs international de mcanique des sols et fondations, Istanbul,
vol. 2, 2000, p. 1007-1010.
Simon B., Schlosser F. - Soil reinforcement by vertical stiff inclusions in France.
Symposium Rigid inclusions LQ GLIILFXOW VXEVRLO FRQGLWLRQV ISSMGE TC36, Sociedad
Mexicana de Mecanica de Suelos, UNAM, Mexico, 2006.
Swedish Geotechnical Society (SGS) /LPH DQG Lime Cement columns. Guide for design,
FRQVWUXFWLRQDQGFRQWURO6*)5HSRUW)RUVPDQ-XSGDWHG
Zeevaert L. Discussion on negative friction and reduction of point bearing capacity. Proc.
4th.ICSMFE, 1957, vol. 3.
Operating mechanisms
Summary
89
CHAPTER 3
Design methods
1. NOTATIONS
The forces and loads considered in the design methods are shown in the reference diagram in
Figure 3.1, where depths are measured from the inclusion head.
The force-reduction elements in the cross-section of an inclusion, denoted respectively Qp (for
axial force), Tp (shear force) and Mp (bending moment), can be identified by their respective
values:
Qp(z), Tp(z), Mp(z) in the cross-section of depth z below the head;
Qp(0), Tp(0), Mp(0) at the level of the inclusion head;
Qp(L), Tp(L), Mp(L)at the level of the inclusion base (length L).
When the computational model includes a fictitious column overlying the inclusion on the
height of the load transfer platform hR (or of the embankment), the forces applied at the top of
the fictitious column are respectively Qp(-hR), Tp(-hR) and Mp(-hR).
When taken into consideration, the reduction elements applicable to a horizontal section of
the volume of reinforced soil excluding the inclusions are designated using the same
convention, by substituting the index s for p, i.e. Qs, Ts and Ms.
Figure 3.1: Reference diagram adopted for identifying the force reduction elements
in the inclusion or soil.
Design methods
91
2. UNDERLYING PRINCIPLES
The design methods of a foundation on a soil reinforced by means of rigid inclusions are of
three distinct types:
Analytical models (Section 3);
Numerical models in continuous medium using either finite elements or finite differences
(Section 4);
+RPRJHQL]DWLRQ PRGHOV whose application remains complex and not yet widespread
(Section 5).
Each of these models may be used for both the ultimate limit state (ULS) and service limit
state (SLS) verifications provided it is capable of taking into account:
The behavior of an isolated inclusion loaded at the head, which is comparable to the
behavior estimated by implementation of the W w curves (curves linking the mobilized skin
friction at a given depth with the corresponding axial displacement) and q wb curves (point
pressure point displacement curve) proposed by Frank and Zhao (1982) as functions of the
pressuremeter modulus EM, the pile diameter and the type of soil;
The behavior of a structure, when applicable, in the absence of inclusions, in terms of both
settlement and bearing capacity, with predictions based on classical methods such as:
pressuremeter or penetrometer methods in the case of a shallow foundation; the oedometer
method and bearing capacity and punching calculation in the case of an embankment, raft or
pavement.
Regardless of the method employed (analytical or numerical), the first step consists of
evaluating the behavior of the structure without inclusions, which will allow determining the
contribution of inclusions relative to the settlement and bearing capacity/punching.
3. ANALYTICAL METHODS
Three methods were selected within the scope of the ASIRI project; all three are simple to
implement for solving some of the most common situations: grid of inclusions placed in the
middle of the reinforced zone and subjected to uniform vertical loads. Their main intent is to
calculate the average expected settlement after reinforcement and then estimate the maximum
load applied to the inclusions.
An extension to the case of vertically and horizontally-loaded footings was derived for each
of these selected analytical methods.
Such methods are not necessarily analytical in the strict sense of the term in that they typically
do not lead to results in the form of explicit analytical formulas; they usually need to be
solved numerically through the use, for example, of a spreadsheet or a solver.
These methods however are capable of estimating the distribution of forces between soil and
inclusions, settlements and horizontal displacements.
They do not provide the possibility of incorporating any geotextile reinforcement in the load
transfer platform.
92
Figure 3.2: Semi-empirical mobilization curve of the resistance at tip of the inclusion
(Frank and Zhao, 1982)
kq
mq E M
DP
(3.1)
93
The values mq, which depend on both the type of soil and the method of installation, are listed
in Table 3.1 below (Frank and Zhao, 1982).
Table 3.1: Values of mq (Frank and Zhao, 1982).
Soil
Silt, clay
Sands, gravels
Chalk
Marly limestone
Marl
Driven, closed
11
14
4.8
4.8
4.8
Type of pile
Driven, open
11
11
11
11
11
Bored
11
4.8
11
11
11
b) Skin friction
Skin friction mobilization is described at any depth by a transfer function, thus expressing the
GHSHQGHQFHRIVKHDUZLWKWKHUHODWLYHGLVSODFHPHQWZ]
The empirical curves established by Frank and Zhao (1982) for piles bored into fine-grained
or granular soils, based on interpretations from a large sample of load tests, prove well suited
for describing the interactions that develop along the full length of the inclusion. These
relationships may be applicable as an absolute value, for positive as well as negative friction.
Figure 3.3: Semi-empirical mobilization curve of shaft friction (Frank and Zhao, 1982).
kW
mW E M
DS
(3.2)
Driven, closed
2
3
0.8
0.8
0.8
Type of pile
Driven, open
2
2
2
2
2
Bored
2
0.8
2
2
2
Dp q p
mq E M
q
ln1
qp
(3.3)
Ds q s
W
ln1
mW E M q s
(3.4)
Design methods
95
Figure 3.5: Development of shear along the fictitious column overlying the inclusion.
96
where appropriate), P the inclusion perimeter, and finally Ap and As the respective crosssections of the inclusion and soil at the indicated level.
Both the inclusion settlement and average soil settlement satisfy the following equations:
dy p ( z )
dy s ( z )
Qp ( z )
dz
(3.7)
Qs ( z )
dz
As E s
(3.8)
Ap E p
where Ep(z) is the Youngs modulus of the inclusion material, and Es(z) the apparent modulus
of deformation of the soil at the applied load level, to be evaluated based on an oedometric
type of behavior (i.e. zero lateral deformation at the models outer vertical boundaries and at
the soil/inclusion interface, see Section 2.1.1.2 and Chapter 6, Note 6).
Friction mobilization is described at the level of each layer by a transfer function (see Section
ZKLFKH[SUHVVHVWKHGHSHQGHQFHRIVKHDUon the relative displacement w(z), defined
as the difference between inclusion displacement and soil displacement: yP(z)-ys(z):
W( z ) f y p ( z ) y s ( z )
(3.9)
3.1.2.3. Solution
The nonlinear system composed of differential equations (3.5) through (3.9) requires, for its
resolution, that the relevant boundary conditions be defined. Such conditions are partially of
the stiffness type, i.e. fixed values of the load/displacement ratios at both the head and base.
The resulting system can only be solved by means of iteration.
The boundary conditions at the head of the model are as follows:
Total load applied at the head of the model Q( hR);
The fraction of this load applied at the head of the model directly above the inclusion
EQP = Qp( hR)/Q( hR):
- in the case of an embankment subjected to a uniform load at the head, the factor EQP
constitutes the ratio of the inclusion section area to the grid area.
- in the case of a slab on grade or raft, the value of the factor EQP is established by
assuming that the surface settlements yp( hR) and ys( hR) are equal, which is justified by
the fact that deformations of the structural surface element (concrete slab or raft) are
negligible when compared to soil deformations.
Design methods
97
Figure 3.7: Boundary conditions at the top of an embankment or slab on grade model.
This calculation necessitates the update, at each iteration, of the average soil settlement
profile ys(z) according to Equations (3.5) through (3.9). Moreover, it provides the distribution
vs. depth of the settlement of the inclusion, the average soil settlement, the mobilized skin
friction, the axial force in the inclusion, and the vertical stress in the soil.
Figures 3.8 and 3.9 illustrate the application of this method to a slab on grade (or raft) and to
an embankment over the same soil profile. The settlements calculated using the analytical
method directly above the inclusion and in the surrounding soil are compared to those
obtained from an axisymmetric finite element model. In the embankment case, it is observed
that settlements become uniform in the embankment, at some height above the inclusion head.
98
Figure 3.9: Settlement profiles calculated for a slab (left) and an embankment (right).
3.1.3. Footings
3.1.3.1. Specificities and references
In the case of spread footings placed on rigid inclusions, the limited number of inclusions no
longer makes it possible to satisfy the hypothesis of a large number of adjacent identical grids.
Design methods
99
The interaction of the reinforced soil block below the footing with the surrounding nonreinforced soil domain must be taken into account.
Few experiments involving spread footings on rigid inclusions have been reported in the
technical literature. Nonetheless, two types of experiments will be cited herein, including a
vertical loading of the targeted spread footings:
Tests conducted in the CEBTP research facility's experimental tank specific to a spread
footing lying, through a layer of sand or crusher run, on a loose sand reinforced by micropiles
(Plumelle, 1985);
Tests conducted in a geotechnical centrifuge on spread footings featuring varied geometry
and lying, through a sandy layer, on sand reinforced by inclusions laid out in a very loose
mesh pattern (Bigot et al., 1988).
Lets also mention the lateral load tests on spread footings acting as a piled footing
foundation, conducted in 1970 at Bucknell University (United States), as reported by Borel
(2001).
The ASIRI Projects experimental program was solely focused on the behavior of large
structures: slabs on grade and embankments on rigid inclusions under vertical loading. No
full-scale experimentation or centrifuge tests of a spread footing on rigid inclusions had been
performed. ASIRI project data relative to these types of structures are input from the 3D
numerical modeling of spread footings using models validated by comparison to simulations
of large structures with experimental data collected on the same type of structures. These
numerical models included the cases of vertical or inclined footing loading.
The simplified models proposed could be evaluated by means of a comparison with 3D
numerical modeling results (ASIRI National Project report, 2011); they cover:
Simplified models MV1 through MV3 for a purely vertical loading of the footing;
Simplified models MH1 through MH3, which take into account a lateral footing loading
(inclined force or moment).
The use of these simplified models requires consistency verifications specific to each model,
thus making it possible to validate a posteriori whether or not the model is applicable to the
given situation.
Any model from the MV or MH series can be used interchangeably:
They yield an increasing volume of information (soil settlement, inclusion settlement, soil
stresses, loading inside the inclusion);
They are based on an increasing number of computation steps;
They require verification of a decreasing number of consistency tests, thus making it
possible to verify a posteriori that the hypotheses intrinsic to the model are indeed acceptable.
3.1.3.2. Simplified models for a purely vertical loading
3.1.3.2.1. MV1 analytical model
This model was first presented by Combarieu (1990).
It allows estimating the limit load QL and settlement y of the footing lying on rigid inclusions,
under a centered vertical load, by interpolating between the case of the footing acting alone
and the case of the footing acting as part of a composite foundation.
100
(3.10)
(3.11)
The transition from series (1), (2), (3) to seriesLQYROYHVVXEVWLWXWLQJWKHin situ
soil thickness hR by a load transfer platform. According to this approach, it is considered that
the extension of such a substitution is unlimited within a plane, which obviously constitutes
an optimistic approximation.
Design methods
101
This substitution improves each of the QL terms into QL terms. The improvement in footing
load-bearing capacity due to the mere presence of the load transfer platform (LTP) however
remains small as a result of the LTPs limited lateral extension. In most instances therefore,
this improvement can be neglected for the purpose of justifying the load-bearing capacity.
The results obtained using this calculation method are in agreement with the experimental
findings of both Plumelle (1985) and Bigot and Canepa (1988).
With respect to settlements, it is also well known how to calculate settlements corresponding
WRWKHWZRFDVHVDQGRUDQGZLWKDORDGWUDQVIHUSODWIRUP&DVHRULH
for the footing on inclusions, may be treated by introducing the notion of equivalent modulus
over the height H + hR, thereby simplifying the settlement calculation to that of a shallow
foundation on an equivalent multilayer. The estimated settlement y2 (or y2LVWKHQOHVVWKDQ
that of the footing on its own yet greater than that of the composite foundation, both of which
can be calculated.
Combarieus original method (1990) must be completed by two additional steps, intended to
calculate loadings within the inclusions.
Step 1
This first step consists of establishing the settlement profile ys(z) over the inclusion height, in
association with the settlement value at the head ys( hR). The applied load below the footing
(width: 2R) yields the soil settlement under the footing ys(-hR), by means of pressuremeter
theory, and at every point ys(z) using the method suggested by Combarieu (1988). The
principle here is to replace the pressuremeter modulus EM of each element layer of thickness
R that enters into the settlement calculation, according to pressuremeter theory, by a very
large value, which through the corresponding difference determines the contribution of this
layer to total settlement. The composition of these element settlements is then used to draw
the desired profile ys(z).
Step 2
The loading curve of a single deep foundation element comprising the inclusion itself and the
related load transfer platform prism on top of it, which is assumed to be embedded inside a
soil block subjected to an imposed settlement ys(z) (as derived during Step 1 above), allows
establishing the load Qp( hR) to be applied at the head of this column in order to obtain a
settlement equal to ys(-hr). In this particular calculation, the shaft friction at any point along
the inclusion (respectively the pressure under the tip) is correlated with the relative
displacement w(z) = [yp(z) ys(z)] using Frank and Zhaos laws for friction mobilization
102
(respectively, the point pressure). This calculation also determines the distribution of axial
forces Qp(z) in the inclusion.
In order to validate model consistency, it is necessary to verify that the stress found at the
inclusion head is compatible with the following (see Chapter 5, Section 4.2):
The shear strength characteristics of the load transfer platform;
The ultimate stress value at the inclusion head;
The load transfer platform geometry, taking into account edge effects as required.
This consistency test is critical to the model validation step:
When this test is not satisfactory, it becomes necessary to use a more detailed model;
When this test is satisfied, the results obtained can be used to proceed with the set of STR
and GEO verifications, as indicated in Chapter 5.
Remark: This method neglects the settlement of the load transfer platform related
to stress concentration on the inclusion head. These settlements however may not
necessarily be negligible.
3.1.3.2.2. MV2 analytical model
This model was presented by Glandy and Frossard (2002).
It is intended to study the behavior of a footing on inclusions subjected to a vertical load. This
problem can be broken down into two distinct domains:
1) The pile domain, composed of the inclusions and soil volumes that extend the pile
vertically until reaching the lower or upper model boundary. This domain comprises n
identical elements, assumed to exhibit comparable behavior;
2) The soil domain, which serves to complement the above pile domain.
Design methods
103
The interaction between the pile domain and the soil domain is assumed to be fully
described by the shear stress W that develops at their shared vertical boundary and by the
vertical stress q at the base of the pile domain; it is assumed that these interactions can be
described by the corresponding Frank and Zhao laws applied, considering the relative soilinclusion displacement at all points (instead of the absolute displacement for a pile).
Moreover, the method assumes that the settlement profile at the bottom of the footing is the
same for both domains.
This method requires an iterative computation in order to ensure the consistency of
calculations carried out in each domain, from the standpoint of settlements as well as
interaction forces. Moreover, it allows the estimation of a complete settlement profile along
the inclusion axis (pile domain) and in the soil domain underneath the footing, given an initial
hypothesis of total vertical load Q distribution between the soil domain (Qs fraction) and the
pile domains (n Qp fraction).
104
Based on the load Qs( hR), which is assumed to be transmitted to the soil under the footing,
the settlement profile ys (z) below the footing can be determined by associating the
pressuremeter method with the method suggested by Combarieu (1988), as discussed in the
previous section; this profile also provides the settlement of the soil domain at the level of the
inclusion tip ys(L).
Based on an assumed inclusion displacement relative to this settlement ys(L), the following
can be calculated:
End-bearing pressure mobilized at the tip,
and for each successive slice along the entire pile domain height (i.e. inclusion plus volume of
overlying load transfer platform):
The amount of mobilized skin friction, taking into account the relative inclusion-soil
displacement (or soil-soil displacement in the platform);
The normal stress;
The elastic compression of each slice and its cumulative settlement;
The load Qp(-hR) and the total settlement yp ( hRDWWKHWRSRIWKHSLOHGRPDLQ
An iterative calculation of the assumed displacement at the tip needs to be performed in order
to obtain the settlement compatibility at the top of both domains: ys ( hR) = yp ( hR).
Iterative calculations of the ratio of the load assumed to be applied to the soil domain is also
carried out so that this load gets summed with the nORDGVFDOFXODWHGDWWKHWRSRIWKHSLOH
domain, yielding a value equal to the total load applied to the foundation:
Q( hR) = n * Ap * Qp( hR) + Qs( hR) * (S n * Ap)
(3.12)
where Ap: area of the cross-section at the pile domain head;
S: area of the footing.
Figure 3.15 displays the application of the model to a simple case (Frossard, 2000), with the
calculation highlighting the various points to be verified.
Figure 3.15: Example of results derived from the MV2 model (Glandy and Frossard, 2000).
In order to validate model consistency, it is necessary to verify that the stress obtained at the
inclusion head is compatible with the load transfer platform shear strength characteristics and
moreover that the mobilized skin friction W above the neutral plane is indeed compatible with
the limiting values of negative skin friction (W .WDQGVv
This validation step requires that the vertical stress distribution in the soil domain be
estimated as follows:
Design methods
105
For a relatively thin load transfer platform compared to the footing width, it can be assumed
that the total load at the bottom of the load transfer platform is equal to the load on the
footing. According to this condition, it is possible to evaluate the stress in the soil domain at
WKH ORDG WUDQVIHU SODWIRUP EDVH IURP WKH ORDG LQ WKH SLOH GRPDLQ IRU WKH VDPH HOHYDWLRQ
This allows verifying that the stress obtained at the inclusion head is compatible with both the
load transfer platform shear strength characteristics, i.e. ultimate stress at the inclusion head
and the platform geometry;
The edge effect (Chapter 5, Section 4.2);
The consistency verification relative to the negative skin friction proves to be a difficult step
as it depends on the vertical stress distribution in the soil domain, in alignment with the
footing, which is not calculated in the model. All verifications with respect to negative skin
friction thus impose adopting an additional hypothesis to allow the estimation of the vertical
stresses at any point (e.g. diffusion of loads transmitted to the soil by the footing at a selected
angle E).
Such consistency tests are crucial to the validation of the model:
When these tests are unsatisfactory, it becomes necessary to use a more detailed model.
When these tests are satisfactory, the results obtained can be used to proceed with the set of
STR and GEO verifications, as indicated in Chapter 5.
Remarks: Friction mobilization laws must incorporate the group effect between
adjacent inclusions: for the same load at the head of the model, the shaft friction
mobilized along an isolated inclusion is greater than that developed on one of the
groups inclusions. To take into account this phenomenon, the friction mobilization
curve in the MV2 model would need to be corrected by lowering, for example, the
value of the coefficient mW defined by the Frank and Zhao curves.
This group effect is naturally modeled in the MV3 model (to be described below),
which uses an axisymmetric model of the inclusion within a given group.
Since the settlement profile generated around the inclusions does not include the
influence of forces being transmitted to the soil below the neutral plane, this model
may tend under certain conditions to underestimate the settlement of the structure
(Fig. 3.16).
Figure 3.16: Loading of the soil block exposed to the action of positive skin friction mobilized
under the neutral plane.
106
Step 1
A study of the behavior, under distributed vertical load, of a basic unit cell without any
interaction with the surrounding domain (i.e. case of a cell placed in the middle of a multiple
network of identical cells) serves to establish the elevation of the horizontal plane underneath
the inclusion tip where soil settlement is uniform (lower neutral plane). The average
Design methods
107
settlement derived between the upper cell face (below the footing), and this lower plane
allows calculating the apparent equivalent modulus of deformation E* of the unit cell under
vertical loading (Fig. 3.18).
Figure 3.18: a) Studied configuration; b) Step 1 of the calculation according to the MV3 model.
Step 2
A study of the vertical monolith with modulus E* assimilated to an isolated pile interacting
with the surrounding (non-reinforced) soil domain, exposed to a vertical force Q, determines
the profile ys(z) of the average monolith settlement, accounting for skin friction mobilization
on the monolith perimeter. The settlement recorded at the top ys( hR) remains less than the
settlement of the cell studied during Step 1, as a result of the stress diffusion through shaft
friction along the surrounding soil block (Fig. 3.19).
108
Remark: The friction taken into account at the monolith edge is a soil-soil type
friction; it may be taken as the undrained cohesion cu in clayey layers. In sandy
layers, the value adopted will be calculated from the values for pl
RUFDQGIRI
the targeted layer.
The models relevance depends on the limiting friction value considered along the
monolith perimeter. It is recommended to calibrate the value of this limiting
friction by applying a monolith model comparable to the footing loading case on
the soil presumed to be non-reinforced in order to verify that the resulting
settlement is in fact comparable to that obtained by pressuremeter theory a load
kpl*. In this case, the monolith has the same height as that used with the
inclusions and a modulus value equal to the modulus E of the soil and not the
modulus E* determined during Step 1.
Step 3
The loading curve of an inclusion assumed to be isolated (including the load transfer platform
prism displaying the same overlying cross-section) in a soil block subjected to an imposed
settlement ys(z) makes it possible to establish the load Qp( hR) applied at the head of this
column in order to obtain a settlement equal to ys( hR). According to this calculation, the
shaft friction at any point of the inclusion is correlated to the relative displacement [yp(z)
ys(z)] at the same depth. This load Qp( hR) then determines the distribution of axial forces
Qp(z) in the actual inclusion. The value at the inclusion head (i.e. depth hR under the base of
the monolith studied during Step 2) is denoted Qp(0).
Design methods
109
To validate model consistency, it is necessary to verify that the stress calculated at the
inclusion head is compatible with the load transfer platform shear strength characteristics and
moreover that the friction mobilized above the neutral plane is compatible with the limiting
values associated with the negative skin friction KtanGVv
Knowing the vertical stress at any depth in the monolith (Step 2) and the axial force in the pile
domain (Step 3) enables the calculation at any point of the average stress in the soil domain
and then the verification of the consistency of these results (load transfer platform and
negative skin friction).
Consistency tests are essential to validate this model:
When these tests are unsatisfactory, it becomes necessary to use a more detailed model.
When these tests are satisfactory, the results obtained can be used to proceed with the set of
STR and GEO verifications, as indicated in Chapter 5.
Remark: Group effects relative to shaft friction are automatically taken into
consideration by use of the axisymmetric model of an inclusion within a group, as
introduced during Step 1.
3.2. Other loads
3.2.1. Horizontal load and moment on footings
The interaction between soil and inclusion is determined through the use of a set of simplified
models identified below.
These models are applied in accordance with the three methods developed subsequently.
110
3.2.1.1. Common features of the simplified models that incorporate both a horizontal loading
and a moment
The loading configuration considered herein has been presented in Figure 3.17; it includes a
vertical force, a horizontal force and a moment.
Lateral forces on the foundation cause shear to develop in both the soil and the inclusions:
Shear generated in the soil introduces horizontal displacements of the soil block;
The inclusions, which are also affected by a portion of the forces transmitted by the footing,
interact with the soil subjected to this displacement field.
The models must account for these interactions to evaluate the stress distribution within the
inclusions.
The horizontal force at the head of an inclusion Tp(0) is bound by the following value since it
can only be transmitted by friction between the load transfer platform and the inclusion head:
T p 0
Q p 0
d min
T ( hR ), Q p 0 tan G
Q( hR )
(3.13)
Design methods
111
These subgrade reaction coefficients are estimated in a manner similar to an isolated pile (see
Appendix I of the NF P 94-262 Standard on deep foundations).
The following analytical models present different ways of evaluating the imposed
displacement field.
Remarks: In general, the same rules as for shallow foundations are applied on the
unreinforced soil regarding the determination of passive pressure and side friction
reactions on the lateral footing faces.
When these reactions are neglected, the horizontal force at the bottom of the
footing is equal to the horizontal force applied at the top of the footing.
3.2.1.2. Simplified models for horizontal loading and moment
3.2.1.2.1. MH1 analytical model
The shear transmitted in the soil can be neglected by assuming g(z) = 0. The upper bound of
the horizontal force potentially developed at the inclusion head is (see Section 3.2.1.1):
T p 0
Q p 0
min
T ( hR ), Q p 0 tan G
Q( hR )
(3.14)
The inclusions are only taken into account if they are strictly located in the compressed zone,
as shown in Figure 3.23.
Remark: The load transfer platform enables spreading shear forces generated by
T with depth. This spreading can be neglected without jeopardizing safety.
Figure 3.21: Inclusions taken into account relative to the lateral loadings.
Remark: Neglecting, during this step, both the group effect (which would imply
lower subgrade reaction coefficients) and the soil displacement g(z) induced by
the residual shear force (T 6Ti) is an acceptable simplification.
3.2.1.2.2. MH2 analytical model
The horizontal force applied at the footing is used for the purpose of calculating the horizontal
displacement of the footing v( hR) lying on an elastic mass.
For this task, it is possible to use the foundation impedance formulas at the surface of a semiinfinite elastic mass. As an illustration, for a circular foundation of radius R at the surface of
an isotropic homogeneous mass characterized by shear modulus G and Poissons ratio Q, the
stiffness Kx = T/v with respect to a horizontal force is given by (Pecker, 1984):
Kx = 8 GR/(2 Q)
(3.15)
Similar solutions for a two-layer system have been given by Gazetas (1990).
Based on the shear stress applied to the soil at the level of the inclusion head, the shear stress
W(z) distribution at any depth could be calculated from a diffusion angle E down to the bearing
layer (where shear strains are considered to be negligible).
Starting from the bearing substratum, the shear strain is integrated in order to derive the
lateral displacement profile as follows:
g(z) = 6Wz/G(z) dz
(3.16)
where G(z) is the shear modulus of the soil at depth z
The parameter E is calibrated to obtain a deformation at the top g(-hR) equal to the value
v(-hR) determined previously.
The resulting displacement profile is the targeted displacement field g(z).
Remark: When the soil is homogeneous, the tri-linear profile proposed by Borel
(2001) for a piled-raft foundation can also be adopted.
Table 3.3: Tri-linear approximation of relative displacement in the center of a stiff rectangular footing L
x B at the surface of a soil mass with height H (valid for H/B > 3).
Design methods
0
1
0.4
0.5
1
0.25
3
0
113
The subgrade reaction calculation incorporates this displacement field g(z) and the following
boundary conditions:
at the inclusion head:
Mp(0) = 0
(3.17)
and Tp(0) calculated in accordance with the recommendations set forth in Section 3.2.1.1.
More specifically, a loading envelope is produced by selecting the value for Tp(0) that yields a
displacement of the inclusion head equal to g(0) i.e. equal to the displacement of the
surrounding soil, provided Tp(0) remains less than the maximum value defined in Equation
(3.13);
at the inclusion tip:
Mp(L) = 0
(3.18)
and Tp(L) = 0
(3.19)
3.2.1.2.3. MH3 analytical model (monolithic method)
The MH3 model, introduced in conjunction with the MV3 model, has two successive
additional steps (Simon, 2010; Fig. 3.23).
Remark: The considered monolith corresponds to only the compressed part under
the footing.
This approach is similar to that employed to study an isolated pile subjected to transverse
loading, as depicted by a force T and a bending moment M applied at the head and/or a
displacement imposed on the surrounding soil g(z), on elastic-plastic springs (p = p(v) < plim).
The limited monolith length-to-width ratio and its orthotropic nature however require taking
the shear deformations of the pile into account.
The simple model of a thin beam, commonly used for piles, tends to be inappropriate in that
case.
114
Step 4
The monolith with an equivalent modulus E*, which can ultimately be extended to the footing
volume (modulus Eb), is assimilated with a horizontally-loaded pile interacting with the
external soil block via elastic springs p(v). The horizontal load considered herein is
characterized by the load system [T, M] applied to the footing. The calculation establishes a
transverse displacement profile g(z) for the monolith, which is subjected to the loading action;
it also determines the rotation Z( hR) of the footing (Fig. 3.24).
Design methods
115
116
Axial forces in the inclusions depend on the cell position with respect to the axis of rotation of
the footing. The footing rotation Z(-hR) (as calculated during Step 4) actually determines
settlement along the axis of the cell placed a distance d from the axis of rotation, as follows:
yp(-hR) = Z(-hR).d
(3.23)
The associated axial forces in the inclusion placed at the center of this cell can then be
estimated by assimilating them with the axial forces found under a uniform vertical loading of
the cell that yields the same settlement yp(-hR) = ys( hR). This step is performed by means of
a specific calculation linking Steps 2 and 3.
The forces Qp(z), Tp(z) and Mp(z) obtained according to the various load cases must be
combined in order to verify the stresses in the inclusions.
Remark: The various inclusions placed beneath the footing are subjected to
vertical forces that in theory differ from one another due to the footing rotation.
The horizontal force applied at the top of each inclusion must remain compatible
with the associated vertical force (according to the formula in 3.13).
3.2.2. Loadings under asymmetric conditions
Inclusions may be subjected to additional bending stresses under certain kinds of situations.
These loads may be imposed on peripheral inclusions installed below embankment slopes, or
on inclusions placed beneath a mechanically-stabilized embankment used as bridge abutment.
They may also be those imposed by asymmetric load conditions on an industrial slab or
during an earthquake.
In the case of inclusions placed under an embankment slope, a g(z) type of approach proves
possible. The maximum displacement gmax of the profile needs to be correlated with the
maximum settlement ys(0) in the intermediate sections following treatment. The
proportionality factor * = gmax/ys(0), as well as the shape of the g(z) curve, may be chosen
following the observations recorded on instrumented embankments (see the NF P 94 262
Standard on deep foundation applications).
The bending stresses inside the inclusions may be calculated similarly to piles placed in a soil
subjected to a displacement g(z).
Remark: Since inclusions only typically represent a small volume of reinforced
soil, it is permitted to consider that free displacement of the soil block between
inclusions is not being modified by the inclusions.
In the case of earthquakes, the g(z) curve to be assessed is the one defined by the free field
soil displacement, which depends on the shear wave velocity Vs profile (AFPS, 2011).
4. NUMERICAL MODELS USING FINITE ELEMENTS OR FINITE DIFFERENCES
The finite element and finite difference methods are the most widespread for resolving
mechanical equations involving continuous media.
Both methods rely on a discretization of the model under study. Based on constitutive models
of both soils and the various component materials found in this model, these methods allow
Design methods
117
simulating the behavior of the discretized medium in terms of stresses and strains under the
effect of loads.
The accuracy of a simulation depends on the relevance and quality of the soil constitutive
models adopted and the discretization strategy selected for the studied soil block.
Moreover, these types of methods lead to a verification of bearing capacity and stability
criteria for the modeled zone.
From the users standpoint, no noteworthy differences distinguish the two methods.
For purposes of illustration, a 2D finite element mesh model will be presented below
(Fig. 3.26) along with a 3D model of a reinforced soil block beneath a footing (Fig. 3.27).
The result obtained through these methods yields an approximated solution whose accuracy
depends on:
The model constitutive laws of materials and interfaces;
Discretization, resulting in a meshing pattern that must be finer where the strain field
variations are greatest;
The type of elements adopted (number of nodes) and interpolation laws incorporated into
each element (linear vs. quadratic);
The use of interfaces between structural elements and the soil in order to allow the
integration of soil/structure interaction phenomena;
The boundary conditions.
118
Design methods
119
is no longer acceptable when adding the rigid inclusions. The 3D inclusion must then be
transformed into an equivalent 2D plate (Fig. 3.29).
A slab on grade may also demonstrate a behavior closer to a 3D representation than a 2D
behavior, in which case the model strays significantly from reality since the computational
space remains two-dimensional.
Figure 3.29: Model of the embankment and soil reinforced by rigid inclusions in plane deformation.
120
121
It might be useful to test the sensitivity of the results by studying how they are affected by the
choice of boundaries further away from the inclusions (larger model).
4.3.2. Boundary conditions
In an axisymmetric model, the horizontal displacement is by definition equal to zero on the
model axis and the outer edge of the model.
122
In the case of a plane strain model, more often than not, a zero horizontal displacement is
imposed at the vertical boundaries of the model. When the problem is symmetrical
(Fig. 3.31), e.g. in plane deformation, the horizontal displacement is by definition equal to
zero on the vertical plane of symmetry.
For both models, a zero vertical and horizontal displacement is imposed on the lower
horizontal boundary.
4.3.3. Axisymmetric model of a unit cell
4.3.3.1. Embankment
Far from the toe of the embankment (Fig. 3.33), the following axisymmetric model may be
considered (Fig. 3.34).
Figure 3.34:
Basic cell.
ab
S
Design methods
(3.24)
123
Figure 3.36:
Unit cell
124
Figure 3.37: Geometry-based criteria for the plane model of an embankment over inclusions.
Design methods
125
It is necessary to examine the amount of settlements caused by this uniform stress in the soil
volume located below this neutral plane of equal settlement.
If these settlements are non-negligible, their estimation may require a specific model distinct
from the axisymmetric model of the unit cell:
Simply extending the unit cell model is most definitely a pessimistic approach since no
lateral diffusion of the load would be allowed to intervene. The resulting calculated settlement
would thus be an upper bound of the total settlement under the structure (Fig. 3.40);
One option consists of estimating the settlements of a fictitious foundation of the same
dimensions as the structure itself, which has been lowered to the level of the neutral plane
positioned below the inclusion tips and loaded with the average stress of the structure. The
total surface settlement is therefore the sum of the unit cell settlement (with the model limited
to the neutral plane beneath the tip of the inclusion) plus the settlement of the lowered
fictitious foundation (Fig. 3.41);
126
Figure 3.42: Homogenized equivalent volume interacting with the non-reinforced soil,
for the purpose of estimating the settlements of the underlying bearing layer.
127
128
In order to select the 5 relevant parameters (see Table 3.4), the most appropriate test is the
triaxial test.
Table 3.4: Mohr-Coulomb parameters.
Parameter
E (Youngs modulus)
Q3RLVVRQ
VUDWLR
MDQJOHRIIULFWLRQ
FFRKHVLRQ
\ (angle of dilation)
Unit
kPa, MPa
Degrees
kPa
Degrees
Means of determination
Triaxial, pressuremeter (EM)
0.2 < Q
Triaxial, correlations?
Triaxial, correlations
\ M 30
This model, which describes (using a limited number of parameters) the soil behavior up to
failure, can still prove overly simplistic in some cases:
For excavation during which the soil mass is unloaded; according to this model, the
unload/reload modulus is equal to the loading virgin modulus, which is an unrealistic
assumption, resulting in an excessive calculated heave at the bottom of the excavation when
compared to experimental data;
For slightly over-consolidated soft soils whose apparent deformation modulus depends on
the stress increment.
The initial stress state of the soil, which depends on both its geologic and anthropogenic
history, must be specified by oedometer testing at various levels in order to determine the
degree of over-consolidation (ROC) vs. depth.
The stresses from the structure, the embankment, slab on grade loads, tanks and reservoirs,
etc. will be used to determine the precise domain of stresses where the various layers will be
loaded, i.e. normally consolidated domain vs. over-consolidated domain (Fig. 3.44).
)RUVRLOVLQWKHQRUPDOO\FRQVROLGDWHGGRPDLQF| 0; the angle of friction will be set at the
critical state MNC = Mcrit (on the yield surface) and will thus be identical to the peak friction
angle.
For soils with stresses in the over-consolidated domain, if the strain is limited and less than
that of the elastic limit (i.e. below the yield surface), then the values will be set at the peak,
where MSC = MpeakDQGFSC Fpeak. In the over-consolidated domain, stress paths end up on
Design methods
129
the limit state ellipse; the ends of these paths are therefore not aligned, though the curve is
usually linearly interpolated in order to calculate MpeakDQGFpeak. It can be noted on Figure
3.44 that the determination of MpeakDQGFpeak requires experience and is less reliable than in
the normally consolidated domain.
SuggHVWHG YDOXHV RI F DQG M, based on correlations as a function of Ip, are provided in
Chapter 6.
4.4.2.2. Elasto-plastic model with strain hardening (modified Cam-clay)
This model is well adapted for soils subjected to primary consolidation phenomena.
The concepts of elastic limit state and critical state are introduced on the basis of:
Isotropic consolidation tests that define the elastic limit points (Fig. 3.45);
Triaxial tests that define the critical states and critical state line T 0Sin the Cambridge
D[HV^ST`)LJZLWK
6 sin M '
q
p'
3 sin M '
(3.25)
130
So, in order to determine the 5 main parameters of the model, a series of oedometer tests and
triaxial tests need to be conducted (see Table 3.5 for further details).
The Soft Soil model has been derived from the modified Cam-clay model.
Table 3.5: "Soft Soil" model parameters.
Parameter
* (compression index)
* (decompression-recompression index)
MDQJOHRIIULFWLRQ
FFRKHVLRQ
\ (angle of dilation)
Units
Degrees
kPa
Degrees
Means of determination
Oedometer: Cc / 2.3 (1 + e)
Oedometer: 2Cs / 2.3 (1 + e)
Triaxial (or correlations?)
Triaxial (or correlations?)
\ M 30
4.4.2.3. Elasto-plastic model with strain hardening (modified Cam-clay) and creep
This model incorporates soil viscosity (Fig. 3.47) and is particularly applicable to soft organic
clay and peat for which creep (secondary consolidation) cannot be neglected when compared
to Terzaghi's primary consolidation.
Design methods
131
(3.26)
(3.27)
Parameter
* (compression index)
* (decompression-recompression index)
P* (creep index)
Units
MDQJOHRIIULFWLRQ
FFRKHVLRQ
\ (angle of dilation)
Degrees
kPa
Degrees
Means of determination
Oedometer: Cc/2.3 (1 + e)
Oedometer: 2Cs/2.3 (1 + e)
Long-term oedometer: CDe / 2.3
(1 + e)
Triaxial (or correlations?)
Triaxial (or correlations?)
\ M 30
132
Second-order models are currently the most advanced for the purpose of simulating the
behavior of both stiff and soft soils. As an example, let's describe the hyperbolic constitutive
model which is an elasto-plastic type model with strain hardening (so-called Hardening Soil
Model); it features an initial nonlinear behavior, with a plastic threshold, an unloading
modulus greater than the virgin loading modulus, and both compression and shear hardening
(Fig. 3.48). This model also allows describing the dilatancy prior to failure.
The deviatoric curves q - axial strain H1 from the triaxial tests are approximated to hyperbolas,
i.e.:
q
H1
H
1
1
Ei q a
(3.28)
Figure 3.48: Relation between stress q and strain H1 in the hyperbolic model.
V'
kp a 3
pa
(3.29)
where k and m are parameters determined based on drained triaxial tests and pa is the
reference pressure, typically taken as atmospheric pressure; k depends on the type of soil and
its density. For sand material, m is on the order of 0.5, while for clay m is closer to 1.
The failure threshold qf is determined from the Mohr-Coulomb criterion; the deviatorhyperbolic strain curve is truncated accordingly (Fig. 3.48). Rf is the ratio of qf / qa ; its value
generally lies between 0.8 and 0.9 and in the Plaxis software has been set by default at 0.9.
In order to completely describe this model, a total of 10 parameters are necessary. A triaxial
testing campaign should be conducted, in theory, with an unloading-reloading loop so as to
measure Eurref. Oedometric tests are also recommended. (Table 3.7).
Design methods
133
Parameter
E50ref
Eoedref
M
Eurref
Qur
pref
Rf
MDQJOHRIIULFWLRQ
FFRKHVLRQ
\ (angle of dilation)
Units
kPa
kPa
kPa
kPa
Degrees
kPa
Degrees
Means of determination
Triaxial CD, with a 'V measurement
Oedometer
Triaxial, correlations?
Triaxial, in unloading (3 u E50ref)
Triaxial, in unloading (0.2)
(100 kPa)
Triaxial (0.9)
Triaxial, correlations?
Triaxial, correlations?
\ M 30
The soil parameter determination protocol is provided as part of Chapter 6, Soil Surveys.
Remark: In order to calibrate the parameters in this model, results from drained
triaxial tests and/or oedometer tests are required.
4.4.3. Hydraulic characteristics of soils
The soil permeability values (both horizontal and vertical permeability) allow taking
consolidation into account. The selection of these values requires that the evolution of the
permeability with the load from the structure is known. Lets recall at this point the
relationship between the coefficient of vertical consolidation cv and kv:
cv
k v E oed
Jw
(3.30)
This formula can also be expressed as a function of Cc, in the normally consolidated domain
and, by extension, in the over-consolidated domain as a function of Cs, for a given vertical
stress VDVFORVHDVSRVVLEOHWRWKHYHUWLFDOVWUHVVIURPWKHVWUXFWXUHEHLQJFRQVLGHUHG
k 1 e
(3.31) in the elastic domain: VVp
cv
2 .3 v
V'
Cs J w
k 1 e
(3.32) in the plastic domain: V!Vp
cv
2 .3 v
V'
Cc J w
The value of this coefficient depends on the vertical stress VDVZHOODVRQWKHFRHIILFLHQWRI
vertical permeability kv (expressed in m/s), and soil compressibility Cc/(1 + e0).
The variation in permeability kv is directly derived from variation of the void ratio e (Fig.
3.49) by the following relation:
k
e e0
lg v
ck
k v0
'e
134
c k 'lg k
(3.33)
Figure 3.49: Curve depicting the variation in coefficient of permeability kv vs. void ratio,
thus enabling the determination of parameters ck and log kv0.
Benchmark values and correlations between these parameters are discussed in Chapter 6.
4.4.4. Characteristics of the inclusion/soil interfaces
These numerical calculation tools propose various types of interfaces for simulating the soil /
inclusion interaction. The constitutive models typically used for interfaces are of the elastoplastic type.
The elastic part allows modeling a gradual mobilization of shear with strain.
As for the plastic part, two techniques are used:
Either a reduction of MDQGFLVDSSOLHG
Or a fictitious soil M ZLWK D QRQ]HUR FRKHVLRQ IRU VLPXODWLQJ FRQVWDQW IULFWLRQ
F Ts, in compliance with the limiting values of shaft friction set forth by the French
standard for deep foundations.
Other means are also available to handle this phenomenon, for example by refining the mesh
around the contact zone. In this approach, it is more difficult to incorporate the limiting shaft
friction values qs, which also depend on the techniques of execution of the inclusions.
In order to fully represent this interaction and define the set of model parameters, data from in
situ axial load tests conducted on an isolated inclusion may also be used.
4.4.5. Characteristics of the rigid inclusions
A linear elastic model is typically adopted for the inclusions.
The value of the long-term modulus is generally used in the design calculations. The shortterm modulus value is only used for short-term exceptional situations (shocks, impact loads or
seismic actions).
The Poisson's ratio value is typically set equal to Q = 0.2.
The diameter of the inclusions in the model is the nominal diameter defined in Chapter 7,
which depends on the execution technique, while the length selected is the theoretical value.
Design methods
135
Design methods
137
138
Figure
Figure3.52:
3.52:Extension
Extensionofofthe
theplastic
plasticzones
zonesunder
under7
m7ofmembankment,
revealing
a
confined
failure
bulb
of embankment, revealing a confined failure
above
the
inclusion
head.
bulb above the inclusion head.
5. HOMOGENIZATION-BASED METHODS
5.1. Simple homogenization method
One homogenization method applicable in cases when loads are exclusively vertical consists
of modeling beforehand the behavior of a typical representative unit cell including a typical
inclusion along with its associated soil and load transfer platform volumes. This unit cell is
generally modeled using an axisymmetric model.
The results of this model are then compared to those obtained with a second model featuring
the same dimensions, where the soil and inclusion are replaced by a unique and homogeneous
material, with characteristics chosen so as to yield results similar to the first model.
This equivalence is typically achieved with regards to settlement, which allows defining the
apparent modulus E* of the equivalent homogeneous material that provides an average
equivalent settlement at the surface of the model.
The properties derived may then be used in a more comprehensive model that encompasses
some or all of the unit cells. This approach reduces the complexity of the model while
describing the complete behavior of the structure. This approach is only valid for the specific
modeled loading case.
Let's point out that it is not possible for rigid inclusions to use the same homogenization
procedure as that applied when designing stone column reinforcement projects, which states:
Design methods
139
Figure 3.53: Principle behind the biphasic modeling of a soil reinforced by rigid inclusions.
Such a modeling approach leads to a description of stresses (i.e. internal forces) relative to
each of the phases assessed separately (Fig. 3.54): standard stresses for the matrix phase
representing the soil; axial force, shear force and bending moment densities for the
reinforcement phase, which is modeled as a continuous distribution of beams.
140
In the case of a soil layer reinforced by a uniform distribution of vertical rigid inclusions,
subjected to a uniform and vertical load (Fig. 3.55), the biphasic model is equivalent to a
simplified approach developed in the Taspie + computation software (Cuira and Simon,
2009). According to this configuration, forces inside the inclusions are actually simplified to
the axial loading, whereas the interaction forces I and p are purely vertical. The interaction
laws can thus be directly derived from the "W-w" curves classically established for piles.
Far from being limited to this unique situation, the biphasic model may in fact be easily
integrated into a finite element computation code, and this software can solve a very broad
array of situations. The main advantage of such a modeling technique, compared to a more
conventional design method that models the soil and inclusions as two geometrically distinct
elements, is twofold. With regards for example to a numerical simulation of the problem
using either the finite element or finite difference technique, the conventional approach
consists of separately discretizing the inclusions and soil, resulting in a mesh size significantly
smaller than the size of the inclusions. Since the problem also happens to be threedimensional, such a numerical procedure requires considerable mesh preparation time and
Design methods
141
computation time and is therefore not compatible with the goal of a quick design of the
solution.
As an example of the superior performance of the biphasic model, Figure 3.56 shows the
surface settlement profiles of a reinforced soil layer subjected to the weight of an
embankment, as evaluated based on a 2D biphasic simulation and a comprehensive 3D
numerical calculation whose implementation was far more complex than the 2D biphasic
calculation (Hassen et al., 2009). The biphasic model was incorporated into the CESARLCPC software package (Bourgeois et al., 2006), with the constitutive laws of the individual
phases, as well as the interaction laws, assumed to be elasto-plastic.
Figure 3.56: Comparison between a biphasic (2D) calculation and a 3D numerical simulation
(extracted from Hassen et al., 2009).
142
References
AFPS - Amlioration et renforcement des sols sous actions sismiques . Groupe de travail,
2011.
Bigot G., Canepa Y. - Fondations de btiments. Utilisation des techniques de traitement et
GDPpOLRUDWLRQGHVVROV . Rapport interne LPC, 1998, p. 37 - 41.
Borel S. - Comportement et dimensionnement des fondations mixtes. tudes et recherches
LPC, GT 73, 2001.
Bourgeois E., Rospars C., Humbert P., Buhan (de) P. - A multi-phase model for finite
HOHPHQW DQDO\VLV RI WUDFWLRQ IRUFHV LQ EROWV XVHG LQ WKH UHLQIRUFHPHQW RI WXQQHO ZDOOV
Proc. Num. Meth. Geotech. Eng., Schweiger (ed.), Taylor & Francis Group, London,
2006, p. 341-346.
BS8006 - Code of Practice for Strengthened/reinforced soils and other fills, Section 8,
Design of embankments with reinforced soil foundations on poor ground
Cartiaux F.-B., Gellee A., Buhan (de) P., Hassen G. Modlisation multiphasique
DSSOLTXpHDXFDOFXOGRXYUDJHVHQVROVUHQIRUFpVSDUinclusions rigides . Revue franaise
de gotechnique, n 118, 2007, p. 43-52.
Combarieu O. - &DOFXOGXQHIRQGDWLRQPL[WHVHPHOOHSLHX[VRXVFKDUJHYHUWLFDOHFHQWUpH .
Note dinformation technique LCPC, 1988, 15 p.
Combarieu O. - Fondations superficielles sur sol amlior par inclusions rigides
verticales . Revue franaise de gotechnique, n 53, 1990, p. 33-44.
Cuira F., Simon B. - 'HX[RXWLOVVLPSOHVSRXUWUDLWHUGHVLQWHUDFWLRQVFRPSOH[HVGXQPDVVLI
renforc par inclusions rigides . Proc. 17th ICSMGE, Alexandrie, M. Hamza et al.
(Eds.), IOS Press, 2009, p. 1163-1166.
Dupla J.-C., Canou J., Dinh A.Q. - Caractrisation des graves utilises sur les plots
exprimentaux de Saint-Ouen-O$XP{QHHW&KHOOHV . Rapport ASIRI 1.07.3.02, 2007.
Frank R., Zhao S.R. (VWLPDWLRQ SDU OHV SDUDPqWUHV SUHVVLRPpWULTXHV GH OHQIRQFHPHQW
sous charge axiale des pieux fors dans les sols fins . Bulletin de liaison des laboratoires
ponts et chausses, n 119, mai-juin 1982, p. 17-24.
Design methods
143
144
Summary
Chapter 4
Design considerations
1. CHOICE OF IMPROVEMENT METHOD
The purpose of this chapter is to draw up an inventory of items that would justify choosing a
rigid inclusion-based solution.
The uniqueness of a reinforcement scheme using vertical rigid inclusions is for the soil, in
relative proportion to its own strength, to bear a portion of the loads generated by the
structure, with the bulk of these loads being transmitted by various mechanisms to the rigid
inclusions. This process applies to foundations as well as to earthen structures.
The basic criterion inherent in this set-up is the relative displacement of the inclusion with
respect to the soil: for the system to operate optimally, soil settlement at the level of the
inclusion head must be significantly greater than the displacement of the inclusion itself, with
a granular load transfer layer intended to absorb the resulting differential movement.
Generally speaking, the choice guiding a project towards a given type of soil improvement or
reinforcement solution is, in most cases, based on an economic rationale taking into account
traditional deep foundation solutions, associated with rigid and expensive structural designs
(e.g. a solution of structural slab with a distributed load).
This rationale generally comes first when designing the foundations for a heavily-loaded
industrial slab on surface areas extending several thousand square meters.
Various soil improvement and soil reinforcement processes may be considered for eventual
selection within the scope of a slab-on-grade project featuring typical characteristics and/or a
shallow type foundation system that meets both the absolute and differential settlement
criteria.
The choice of methodology is primarily dictated by the following criteria:
Data concerning the soil,
- type of compressible soil and its geomechanical characteristics,
- thickness of the compressible layer,
- presence of an aquifer;
Data concerning the structure,
- purpose and geometry of the intended structure (type and loading conditions),
- operating constraints,
- allowable residual strains and settlements;
Data concerning works execution conditions,
- environmental context,
- interface with the various construction phases (earthworks, paving, structural frame, etc.),
- works schedule.
Design considerations
145
2. INPUT DATA
Depending on the set of criteria indicated above, it would be advised to specify the following
items, above all else, in order to determine the "value" of a rigid inclusion solution.
2.1. Data relative to the soil
One or more standard lithostratigraphic cross-sections;
Compressibility, eventual overconsolidation profile; these data must be used to derive an
estimated range of final total settlement without reinforcement;
Water table elevation(s);
Data relative to both the consolidation rate and permeability of the various layers;
Soil strength profile: load-bearing capacity vs. foundation elevation;
Compressible thickness heterogeneity;
Lateral confinement capacity;
Organic matter content;
Presence of a creep-sensitive layer;
Onsite seismic conditions.
2.2. Data relative to the structure and its operations
Layout plan and footprint of the project structure;
Applied loads:
- permanent or temporary loads;
- distributed or point loads;
Eventual zoning considerations designating the extent of high-load areas and their intensity
(whether maximum or minimum);
Operating constraints;
Displacement and strain criteria;
Type of structure being supported;
Technical guidelines to implement.
2.3. Data relative to works execution conditions
Works schedule and milestones;
Eventual aggressiveness of the environment with respect to concrete;
Presence of cobbles creating potential obstructions to penetration;
Water table elevation (substitution method feasible or not );
Water table variations (seasonal, long-term);
Type of excavated materials and conditions for reuse.
3. PROJECT-SPECIFIC CRITERIA
The project design necessitates that the deformation criteria to be met during serviceability
limit state (SLS) verifications must have been defined and approved by the Project Owner,
i.e.: absolute and/or differential settlements.
Remark: The criteria to be taken into account for ultimate limit state (ULS)
verifications are listed in the most current rules and regulations as well as in the
present recommendations.
In the absence of prescriptions set forth by the Project Owner, it is possible to refer to one of
the following documents:
146
Annex H (informative) of Eurocode 7-1 (NF EN 1997-1 Standard), which gives the limiting
values of structural deformation and foundation movement;
Standard NF P 11.213 (DTU 13.3 reference), which indicates the limiting values of absolute
and differential settlements for concrete ground slabs;
Design of shallow and deep foundations (Frank, 1999), which provides a review of limiting
values of settlements applicable to buildings and civil engineering structures;
The RFF Railway Operator's Technical Guide;
Technical guidelines specific to tanks and reservoirs;
Guidelines relative to road pavements (surface evenness).
As a complement, a number of deformation-related consequences must also be taken into
consideration with respect to:
Maintaining the slopes for all utility networks;
Operating criteria specific to some equipment (vibrating machines).
In all cases, it is necessary to distinguish between:
Cumulative settlements since the beginning of the building construction;
Settlements occurring during the eventual proof loading test;
Settlements undergone by services and equipment;
Settlements recorded on the structure itself as well as those induced on neighboring
structures.
Experience has shown that a solution of reinforcement using rigid inclusions can lead to a
reduction in settlements by a factor of between 2 and 10 compared to settlements of nonreinforced soil.
4. USE LIMITATIONS AND FIELDS OF APPLICATION
Any foundation project on soil reinforced by rigid inclusions must consider the full array of
foundation components: inclusions, load transfer platform, and structural elements. The
design must therefore focus on all of these components and cannot be broken down into
several distinct functions that would successively treat each component individually. Herein
lies a major characteristic of all foundation projects built on a soil that has been reinforced by
rigid inclusions. As a result, the Engineer, who oversees and coordinates the design process
and the execution of the works, plays a truly vital role in projects featuring a rigid inclusion
solution.
4.1. Favorable factors
The factors that favor selecting a rigid inclusion strategy consist of:
Projects with settlement criteria on finished structures that are not overly stringent;
The following objectives:
- reducing overall compressibility of the foundation soil,
- increasing load-bearing capacity,
- limiting interactions with neighboring structures,
- allowing the use of a slab-on-grade instead of structural reinforced slabs,
- shortening the construction schedule (given the fact that less of the load gets transferred to
the soil, thus activating a more efficient consolidation phenomenon: reduced compressibility
if remaining in the overconsolidated range, plus a higher consolidation coefficient for lower
stress states),
Design considerations
147
Allocating an over-width of the reinforced zone beyond the identified structural footprint;
Inserting reinforcement layers (steel wire mesh or geogrid) immediately above the inclusion
heads.
Moreover, it is sometimes necessary to design a gradual transition zone placed between the
treated and untreated zones (adaptation of the grid of installation and/or inclusion length
between a reinforced and unreinforced platform zone, or between the embankment at the
periphery of a bridge abutment and an intermediate embankment section).
For some structures, edge effects might result in treating a larger footprint than the net
footprint of the structure.
5.3. Special case of a very thick embankment installed on a consolidating soil
Caution must be exercised with respect to the initial state of consolidation of the in situ soils,
especially when the site has been backfilled (whether recently or in the past) and the
consolidation of lower layers is not yet complete: the settlements experienced by the structure
will not exclusively depend on the structural loads, but also on a continuation of excess pore
pressure dissipation. The underconsolidated nature of the layers can only be revealed through
oedometric testing, piezocone tests or pore pressure cells.
The case where the soil profile contains layers capable of exhibiting creep (peat or organic
layers) falls in the same category as the previous case since creep settlement does not directly
depend on loads applied at the surface.
Another case would be one where project grading work requires a major preliminary
backfilling process. Under these circumstances, it is preferable to install the inclusions prior
to the embankment and then account for this load in the inclusion design.
The introduction of reinforcement layers at the base of the embankment can, in some
instances, offer a desirable option.
5.4. Choosing among the various rigid inclusion processes
During the preliminary project design stage, the type of inclusion (i.e. with or without soil
displacement, driven, etc.) is typically not taken into consideration.
An estimation of the load transmitted to the inclusion gives the order of magnitude of the
targeted strength for a given diameter.
5.5. Impacts on Slab-on-grade design
The behavior of a foundation installed on reinforced soil constitutes a complex problem of
soil-structure interaction between several localized elements with different characteristics and
behavior (soil, concrete).
The parameters involved in these interactions sometimes oppose one another on settlements
or loadings in the structure:
On the one hand, a stiffer load distribution platform induces smaller settlements, while
generating stronger bending moment and shear force values within the slab;
Design considerations
149
On the other hand, a more flexible load transfer platform creates larger settlements with less
demand on the slab.
The mechanical characteristics of the load transfer platform layer (deformation modulus,
angle of friction, cohesion) must be chosen so as to cover the expected dispersion in these
values, by focusing on the high or low results, depending on the type of verifications where
these values will be used.
For the load transfer platform layer made of lime- or cement-treated soil, it is indeed
necessary to take into account the eventual evolution in mechanical characteristics depending
on the type of binder, its dosage and moisture variations. For these materials, the influence of
loading type (e.g. cyclical nature) also needs to be incorporated.
6. DEVELOPMENT STEPS IN A CONSTRUCTION PROJECT
The design of a structure on a soil reinforced by rigid inclusions entails a step-by-step
approach: preliminary design using simple methods, followed by the detailed project design
and execution planning. This design process relies on a sequence of the following
geotechnical engineering missions, as defined in the NF P 94-500 Standard:
Preliminary site-specific geotechnical study (G11);
Preliminary geotechnical design (G12);
Detailed geotechnical project design (G2);
Geotechnical design and monitoring of project execution (G3);
Geotechnical supervision of the execution phase (G4).
Remark: Two documents (Syntec Ingnierie, 2009 and 2010) explain the links
between the geotechnical engineering missions specified in the NF P 94-500
Standard and the Engineers missions according to the MOP Law for
constructing buildings or infrastructure.
Another document, entitled Geotechnical considerations in the design and
execution of concrete slabs and released in June 2011 by Syntec/Unesi/Coprec,
emphasizes the importance of managing the interfaces among the various project
contractors in reference to the NF P 94-500 Standard.
6.1. Preliminary site geotechnical study (mission G11)
This study is conducted prior to having a precise definition of the project. By referring to
information contained on the geological map, risk prevention plans and data obtained from
the BRGM Office (Geology and Mining) (particularly underground cavities, shrinkageswelling hazards of clays) and DRIRE Office (Environment) (mining, underground
quarrying), the geotechnical engineer must indicate the major risks identified. He also
establishes a program of geotechnical investigations for the purpose of quantifying these risks
and, depending on their nature, recommends a set of feasible technical solutions.
6.2. Geotechnical study of the preliminary project design (mission G12)
Conducted in conjunction with the preliminary project design, this G12 mission is intended to
mitigate the major geotechnical risks identified during the previous study.
With respect to each of the risks encountered, the geotechnical engineer is to:
150
Define one or more geotechnical models depending on the degree of homogeneity of the
site and indicate, if possible at this stage, the depth at which the substratum may be
considered non-deformable; these models are tied to the NGF ground level benchmark
system;
Use these models to examine the set of feasible solutions adapted to the design loads and
estimate the resulting settlements;
Indicate, should the settlements obtained not comply with serviceability criteria, the types of
soil improvements potentially proposed by emphasizing the specificities of each improvement
type (e.g. the hard point effect of rigid inclusions);
Identify the measure(s) to adopt or consider adopting, especially in the presence of soils
sensitive to shrinkage-swelling phenomena;
Establish, if deemed necessary, a program of geotechnical investigations with the specific
aim of determining a geotechnical model that either integrates in-plane heterogeneity or more
narrowly targets the case of soils exhibiting complex behavior.
6.3. Services performed within the scope of the Engineers mission
6.3.1. Geotechnical study of the detailed project design (mission G2)
This G2 geotechnical mission, assigned as part of the Engineers scope of works, lays out a
basic project solution encompassing the inclusions, the load transfer platform and the
structure.
This study, paid for by the Project Owner, is critical to defining the final structural design.
This study is performed based on the Project Owners set of specifications defining the design
loads and acceptable values of deformations.
The geotechnical model(s) derived must be developed to a sufficient level of detail.
The soil deformation modulus values assigned to each layer are calculated as a function of:
Duration of load application;
Load intensity (stress-strain dependency);
Water content variations, causing modifications in consistency and/or suction.
The G2 geotechnical mission comprises:
A design at both the serviceability limit state (SLS) and ultimate limit state (ULS)
(including GEO and STR verifications) for the inclusions and load transfer platform;
Based on the estimation of the settlement of the reinforced soil, definition of the equivalent
homogenized soil profile;
The means and methods to be used by the concrete slab contractor to incorporate eventual
additional loadings due to the hard point effect of inclusions.
Lastly, this mission is to indicate the construction steps to be implemented (particularly in the
case of soils sensitive to desiccation phenomena) and, in the event of in-plane heterogeneities,
it must define the appropriate geotechnical models taking these heterogeneities into
consideration.
This study also facilitates the preparation of the construction project bid documents.
Design considerations
151
References
Frank R. - Calcul des fondations superficielles et profondes 7HFKQLTXH GH OIngnieur,
Presses des Ponts, 1999.
Loi MOP - Loi n 85-704 du 12 juillet 1985 modifie relative la matrise douvrage
publique et ses rapports avec la matrise dXYUHSULYpH, 1985.
NF P 11-213 DTU 13.3 - Dallages - Conception, calcul et excution, 2007.
NF P 94-500 - Missions dingnierie gotechnique, Classification et spcifications, 2006.
Syntec Ingnierie - Synchronisation des missions dingnierie gotechnique et de matrise
dXYUHSRXUODFRQVWUXFWLRQGHEkWLPHQWV. ditions Syntec-Ingnierie, 2009.
Syntec Ingnierie - Synchronisation des missions dingnierie gotechnique et de matrise
dXYUHSRXUODFRQVWUXFWLRQdinfrastructures. ditions Syntec-Ingnierie, 2010.
Syntec/Unesi/Coprec - La gotechnique dans la conception et la ralisation des dallages en
bton, 2011.
Design considerations
Summary
153
CHAPTER 5
Justifications
1. GENERAL FRAMEWORK FOR PRESENTING JUSTIFICATIONS
The justifications associated with structures built on soil reinforced by means of rigid
inclusions are conducted in strict compliance with Eurocode 7. This step requires verifying
both the serviceability limit states (SLS) and the ultimate limit states (ULS).
The structures need to be differentiated depending on whether they are embankments,
pavements and foundation of structures (slabs on grade, rafts, footings) supported on rigid
inclusions. Such structures all share the characteristic of relying on the in situ soil for a
portion of the support of the structural loads, with the bulk of this support being transmitted
via various mechanisms to the rigid inclusions. In the case of foundations, the load transfer is
produced by avoiding any direct contact between the structure and the inclusions thanks to the
installation of a continuous load transfer platform.
The present recommendations do not apply to footings that lie directly over the inclusions.
In the remainder of this document, a distinFWLRQ ZLOO EH PDGH EHWZHHQ WKH 1st domain
corresponding to the case of inclusions required for stability of the structure and the 2nd
domain corresponding to the case of inclusions not required to ensure the stability of the
structure, but instead whose objective is primarily to reduce the settlements.
1.1. Serviceability limit states (SLS)
The limit states that need to be verified are related to either the vertical or horizontal
movement.
1.1.1. Set-up
The behavior of structures lying on soil reinforced by rigid inclusions can only be fully
understood with a detailed study of the interaction between the soil and the various structural
elements (foundations, inclusions). This interaction entails considering the shear mechanisms
at work within the load transfer platform, along the shaft and under the tip of the inclusion.
In practice, the justification of these structures requires the implementation of a computational
model capable of incorporating these various mechanisms as well as ensuring the
compatibility of the deformations required to mobilize each individual mechanism. This step
signifies that the justification systematically includes the calculation of displacements under
service loadings.
The verification with respect to serviceability limit states, to be emphasized herein, places the
focus on deformations as described by Eurocode 7-1 in 2.4.8 (1):
Ed < Cd
(5.1)
Ed: design value of the effect of actions;
Cd: limiting design value of the effect of an action.
Justifications
155
Note 1: The appearance and overall functionality of the structure may be altered
whenever the calculated deflection ymax of a beam, slab or bracket subjected to
quasi-permanent loads exceeds L/250, where L represents the span. Practically
speaking, the deformation after construction ymax is typically limited to L/500 for
quasi permanent loads.
The foundation movement components that need to be considered are primarily the surface
settlement values (average value yave or maximum value ymax) or relative settlement (or
differential settlement: ymax - ymin), either with or without the ground slab, for comparison
with allowable values.
The pertinent settlement criteria are defined by the Project Owner in conjunction with the
Engineer.
Remark: Depending on the nature of the structure bearing on the ground (absence
or not of a slab on grade, relative stiffness of this slab) and on the type and
sensitivity of any surcharge (road, storage), there might be major differences in
the allowable values.
These verifications are quite obviously complemented by verifying that the forces in the
structural elements, most notably the inclusions, are compatible with the design resistance of
the material of the different components.
156
0RUHRYHU LI WKH load transfer SODWIRUP VWUHQJWK criterion is not included in the SLS
computational model, it becomes necessary to verify after the fact the consistency of the
calculated values of the vertical stresses between the soil and the inclusion heads relative to
this shear strength.
1.1.3. Alternative SLS verification
1.1.3.1. Inclusions used to reduce settlement (Domain 2)
The alternative verification with respect to the soil consists of the verification that a
sufficiently low fraction of the ground strength has been mobilized to keep deformations
within the required serviceability limits [EC7-1: 2.4.8 (4)]. Such verification is typically
unnecessary whenever the inclusions strictly serve to reduce settlements, since the settlements
are always calculated (as opposed to the procedure for deep foundations, for which the
displacements are not systematically calculated and may be omitted). It is therefore
unnecessary to verify the SLS forces within the inclusions relative to creep values.
This is also true under the condition that the computational model is capable of accurately
representing the gradual mobilization of both the shaft friction and point resistance along the
entire inclusion, until a point close to failure.
Numerical models in a continuous medium must include interfaces between the soil and the
inclusions, with interaction laws matching collected experimental data (limiting value qs of
the soil-inclusion friction in accordance with the type of soil and selected installation
technique, gradual mobilization of friction with settlement of the element). One approach for
verifying that this behavior is included in the model consists of modeling the case of a direct
load at the top of an inclusion and then comparing the results obtained to the loaddisplacement curve determined using the W-w law method proposed by Frank and Zhao (1982)
or Combarieu (parabolic law, 1988).
In the case of a simplified model, the relevance of the model must be verified with respect to:
Load transfer platform shear and deformation characteristics;
Vertical stress applied to the soil between inclusions.
The model, applied to the case of loading of a rigid shallow foundation without an inclusion,
must also yield results comparable with the settlement estimations obtained by the most
appropriate method between the pressuremeter method or the oedometric method.
1.1.3.2. Inclusions required for the bearing capacity (Domain 1)
When inclusions prove necessary to justify the ULS stability, the conditions set forth in the
'HHS IRXQGDWLRQV DQG 6KDOORZ IRXQGDWLRQV national Standards are applicable. It then
becomes necessary to verify at SLS that:
The maximum load in the inclusion does not exceed the design value of the critical creep
load in compression Rc,cr;d under the neutral plane (Article 14.2.1, NF P 94 262 Standard,
Deep foundations);
The stress applied on the soil at the level of inclusion heads (top of the inclusion) does not
exceed the SLS limiting value (Chapter 13, NF P 94 261 Standard, Shallow foundations).
Justifications
157
(5.2)
Incorporation of these coefficients [1, [2, [3 and [4 becomes necessary if the load-bearing
capacity verificationLVFRQGXFWHGDFFRUGLQJWRWKHPRGHOSLOH method;
These coefficients are not applicable should the verification be conducted based on the
JURXQGPRGHO method.
These recommendations tend to favor the XVHRIWKHJURXQGPRGHO method.
In Domain 2, since justification of the stability of the structure has been performed without
taking inclusions into account, these coefficients are not used.
Moreover, model coefficients JR;d, defined in the national Application Standards for piles (NF
P 94 262) and shallow foundations (NF P 94 261), must be acknowledged; these coefficients
serve to define the characteristic resistance values of the ground Rk based on either
pressuremeter or penetrometer tests.
Remarks: According to Design Approach 2, the choice of partial resistance factors
is complicated by the fact that the assessed foundations cannot be considered in the
category of either shallow foundations or deep foundations. Hence, Section 7 of the
EC7 code which focuses on deep foundations, stipulates that the set of provisions it
describes should not be applied to the design of piles that are intended for
settlement reduction; such an indication directly targets the case of rigid inclusions
placed beneath structural foundations when strictly used to reduce settlements. The
distinction proposed in this document between Domain 1 (inclusions required for
ground bearing stability) and Domain 2 (not required for ground bearing stability)
addresses this specific observation.
According to the national Appendix EC7-1, the Design Approach 3 (DA3) can also
be adopted for the numerical analyses of soil-structure interaction.
1.2.2. Special case of embankments on rigid inclusions (Design Approach 3)
In the special case of embankments on rigid inclusions and for persistent and transient
situations, Design Approach 3 should be selected for the overall or combined stability
verification step, which is consistent with the choice made for reinforced soil structures (NF P
94 270 Application Standard).
The combined stability verification of a reinforced soil structure must be carried out by
considering a sufficient number of overall stability failure surfaces that intercept and/or run
parallel to at least one of the reinforcing elements.
The load transfer calculation considers the internal embankment equilibrium and therefore
must be carried out with Design Approach 2.
Design Approach 3 introduces the following combination for persistent and transient
situations:
$
RU$A05
The symbol * denotes actions from the structure;
The symbol ^ denotes geotechnical actions.
According to this approach, the partial factors are applied to actions or to the effects of actions
from the structure and to ground strength parameters.
Justifications
159
All permanent actions (originating from the soil) are weighted by a factor JG = 1.0, while the
unfavorable variable actions (transmitted through the soil) are weighted by a factor JQ = 1.3.
The M2 partial factors are as follows:
JI = JF = 1.25; Jcu = 1.4; JJ = 1.0
The set of R3 partial resistance factors are all equal to1.0.
2. *(2 LIMIT STATE VERIFICATION MODELS
Experience has shown that for common structures, these verifications may be conducted using
the simplified models described in Section 2.1.
The more detailed models presented in Section 2.2 are typically only justified for exceptional
structures.
2.1. Simplified models
7KH VLPSOLILHG HQYHORSH models are intended to replace the more sophisticated models
described below, in order to demonstrate that the fundamental relation Ed < Rd is indeed
satisfied.
Simplified models lack the capacity to introduce more advanced laws (for soil and/or
interfaces) or to establish relevant failure modes for the considered load case [e.g. difficulty in
handling the failure of a footing exposed to loading (Q, T, M)].
These models are applicable to check both the SLS and ULS of footings, rafts, ground slabs,
embankments or pavements.
A fully elastic model without an interface is at first glance insufficient; at the very least,
interface laws of the type Frank and Zhao would need to be introduced.
2.1.1. Simplified models for ULS verifications
Three approaches of increasing complexity are proposed herein in order to evaluate stability
(Fig. 5.2). It is merely necessary to demonstrate that the stability criterion is satisfied with any
one of the approaches.
Remark: If the stability criterion can be verified using Approach (i), then it is
necessarily verified with approaches of a higher level.
160
161
The approach thus strictly consists of performing ULS (GEO and STR) structural
verifications (on shallow foundations or embankments) while neglecting the presence of
inclusions.
Remark: The envelope model (1) is critical to identify which domain to associate
with the inclusions:
Domain 1: inclusions required for stability;
Domain 2: inclusions used to reduce settlements and not required for stability.
For a shallow foundation (Fig. 5.3), these verifications are carried out according to Design
Approach 2 by following the set of detailed conditions explained in the corresponding
Application Standard (NF P 94 261, being drafted as of January 2011):
Unfavorable permanent actions are weighted by JG = 1.35, and unfavorable variable actions
by JQ = 1.5 (the alternative combination of 1.0 G + 0.0 Q tends not to be relevant);
Bearing capacity is verified using the partial resistance factor JR;v = 1.4;
Sliding is verified by using the partial resistance factor JR;h = 1.1;
A model factor JR;d is introduced into the computation of each of these resistances.
162
F = 1.00
163
QR = 6 Qp(0)
TR = 6 Tp(0)
MR = 6 Qp(0) di
where di is the lever arm of inclusion i relative to the center of the footing.
(5.3)
165
if the computational model is of the simplified type, then the forces transmitted at
the inclusion head will satisfy the consistency conditions explained in Chapter III,
Section 2.2.4.2.
The forces calculated in the inclusions Qp(I) and Tp(I) represent a global reaction (QR, TR,
MR), evaluated similarly to the design loading (Qd, Hd, Md).
QR = 6 Qp(0)
TR = 6 Tp(0)
(5.5)
MR = 6 Qp(0) di + 6 Tp(0) hi
di and hi are the vertical and horizontal distances between the point of intersection of the
failure surface with inclusion i and the point where design loads (Qd, Hd, Md) are defined.
The bearing capacity verification Ed < Rd is thus verified:
By adopting for Ed the combination of the design loading (Qd , Td , Md) and of the reaction (QR , -TR , MR);
By setting for Rd the design value of the ground bearing resistance below a shallow
foundation with the same surface area as the footing, as evaluated according to Standard NF P
94 261.
This approach requires searching for the most critical failure surface as well as the
contribution of the inclusions at the level of their intersection with this surface.
Remarks: The verification against sliding failure is always performed in
accordance with envelope model (2) procedures (i.e. by selecting for QR and TR
the values calculated at the inclusion head).
In the case of a footing subject to a primarily vertical load on a homogeneous soil,
it is acceptable to adopt the geometry of the Prandtl diagram as the critical failure
surface.
Yield design theory (Salenon, 1983) offers the most comprehensive theoretical framework
for evaluating the safety of structures lying on soil reinforced by rigid inclusions; this method
is similar to the method used for soil nailed structures.
For these structures, a more restricted part of the theory will be applied, by only considering:
Its kinematic approach;
The Mohr-Coulomb failure criterion;
The motion of rigid blocks, as delimited by a succession of logarithmic spiral arcs with the
same pole and with a parameter equal to the angle of friction of each layer.
It is possible to take the bending strength of inclusions into consideration in addition to their
compressive strength. Stability may then be evaluated by only considering the contribution of
axial forces in the inclusions or else the combination of axial forces and shear forces. Practice
has shown that a significant stability improvement is generally obtained with axial forces,
whereas the additional benefit derived by adding shear forces is more limited: neglecting the
shear contribution therefore does not exert much influence and remains a safe simplification.
The set of multicriteria (Schlosser, 1983) establishes the domain of possible values for the
force within each inclusion [QP(I), Tp(I)] (Fig. 5.7).
166
Figure 5.7: Examples of the stability envelope domain [Qp(I), Tp(I)] of an inclusion
depending on whether the shear force is taken into account or neglected.
The contribution of each inclusion (Qp(I), Tp(I)) is determined by seeking the maximum
participation to the resistant moment depending on the orientation of the rigid block
displacement velocity vector (Fig. 5.8).
Justifications
167
(Inclusions taken into account in the evaluation of load transmitted to the soil and
by their interaction with the failure surface)
Figure 5.9: Envelope approach 3 applied to the case of a footing
168
Figure 5.10: Decomposition of the problem into two juxtaposed passive and active limit states of
equilibrium.
For embankments, these calculations are carried out using Approach 3 (see Section 0), which
relies on the Yield design exterior approach:
All permanent actions (i.e. originating from the soil) are weighted by a factor JG = 1.0, and
unfavorable variable actions (transmitted by the soil) by a factor JQ = 1.3;
The characteristic soil parameters are assigned the following partial factors:
JI JF = 1.25; Jcu = 1.4; JJ = 1.0;
By analogy with the combined stability verification of reinforced soil structures (NF P 94
270 Standard), a model partial factor JR;d must be introduced into this calculation. The value
of this factor is equal to 1.1 when the structures are relatively insensitive to deformations,
without otherwise influencing any of the other justifications required for serviceability limit
states. It would be necessary to adopt higher values for structures that show great sensitivity
to deformations.
The values for the contribution of each inclusion (QP(I), TP(I)) are computed based on a
strategy that relies on the specifications of Standard NF P 94 270 for nailed soils, as these
structures had also been verified according to Approach 3: the coefficient of dataset M2 for
the soil-inclusion interaction resistance is thus set at: JM,f = 1.1.
Compared to nailed soil structures, structures built on rigid inclusions also require taking into
account the ultimate stresses at both the inclusion head and tip (these contributions are
assumed to be zero in the case of nailed soils due to the small diameter of the nails).
For the inclusions, these contributions are combined with the forces developed by shaft
friction over either the inner part of the studied block or its outer part (Fig. 5.11).
The same value of coefficient JM,f of dataset M2 is adopted for the contributions at the head
and at the tip.
To ensure continuity with the NF P 94 262 Standard (for deep foundations), it is
recommended to perform the calculations according to the "ground model" method, along
with the application of the associated model factor JR,d on the contribution of the inclusions
(shaft friction, head and tip). The values of this factor depend on the inclusion installation
Justifications
169
method, as well as on the type and mode of interpretation of the geotechnical data (i.e. with or
without statistical analysis).
Remark: This is equivalent to applying a reduction coefficient JM,f x JR,d = 1.1 JR,d
on forces Qp(I), Tp(I), Mp(I).
The computation of forces mobilized in the inclusions must also take into account the limit
design resistance of the inclusion material, as explained in Section 3.1 (concrete inclusions) or
3.2 (steel inclusions).
Figure 5.11: Application of the failure calculation to the case of an embankment built on rigid
inclusions, for which only the axial contribution has been taken into account.
Remarks: In the example provided in Figure 5.11, where only the axial
contribution of the inclusion has been taken into account, the inclusions placed to
the left of the center of rotation contribute via a normal compressive force, while
those placed to the right may be working in tension. This maximum resisting
contribution in either compression or tension varies depending on the targeted
failure surface, meaning that the contribution cannot therefore be chosen ahead of
time; moreover, it must necessarily be established for each failure surface by
applying the principle of maximum plastic work. Only the Yield Design approach
allows applying this principle without any approximations.
The design values of the forces in either tension or compression must be
compatible with the STR verifications of the constitutive material of the inclusions.
2.1.2. Simplified models for SLS verifications
All the models described in Chapter III can also be used for SLS verifications.
If some of the strength criteria have not been explicitly introduced in these models, they still
need to be verified a posteriori. Depending on the specific models used, this verification may,
for example, entail:
Punching of the inclusion head into the load transfer platform;
Vertical stress applied at the soil surface between the inclusions with regards to its
allowable value for the specific situation considered
The ratio of negative friction (W/Vv) mobilized relative to the limit value KtanG, specific to
each layer;
Positive friction values relative to the limit friction value qs, associated with both the type of
soil and inclusion installation technique;
170
The value of the stress below the inclusion tip relative to the limit value qp, associated with
both the type of soil and inclusion installation technique;
Maximum stress values in the materials (inclusion, ground slab, reinforcement, etc.) relative
to the allowable SLS values.
2.2. Detailed numerical models
The most comprehensive computational model is a 3D model using finite elements or finite
differences in a continuous medium, through the use of advanced behavior constitutive laws
for the various layers that take into account the range of soil-structure interactions via
interface elements. These laws must be capable of describing the behavior in terms of
stresses/strains until nearing failure.
The introduction of interfaces between inclusions and the ground is essential for two reasons:
It authorizes the sliding of soil against the inclusions;
It allows defining a limit value for soil/inclusion friction in order to reflect the influence of
the method of installation of the inclusions, which would have been neglected in the
computational model (e.g. lateral displacement).
The significant rotation of the principal stresses in the load transfer platform, coupled with the
need to describe a nonlinear behavior highly dependent on the stress state, leads for the
granular material typically used in the platform to the selection of an advanced behavior law
(2nd-order model or nonlinear elasticity).
The 1st-order model associating linear elasticity with the Mohr-Coulomb failure criterion
might not be sufficient to assess ULS situations, even though it generally proves acceptable to
check SLS situations.
The (soil-inclusion) interface laws also need to introduce a failure criterion.
It is preferable to introduce either a Cam-Clay or hardening soil type model (i.e. one that
incorporates an isotropic strain hardening mechanism) for slightly overconsolidated
compressible soils. Only these models are actually capable of reproducing the oedometric
type behavior of soils while distinguishing deformations before and after the preconsolidation
pressure.
In the other cases, in particular for the bearing layer, the requirement or not of advanced laws
must be examined on a case-by-case basis.
For the inclusion component material, the adopted law must necessarily introduce the
material-specific failure criterion (an elastic model without a failure criterion is unacceptable).
The ability of the model to yield representative results up to near failure must be evaluated by
assessing:
The direct load case of an isolated inclusion, and then comparing this simulation:
- to results of a direct load test, if available (with the failure load being set as the load
yielding a settlement at the top equal to one-tenth of the diameter D/10);
- otherwise, to the prediction obtained by using the Frank and Zhao type transfer curves;
The loading of a rigid footing at the model surface (not reinforced by the inclusions) in
order to compare:
Justifications
171
- under SLS loads, the calculated settlement with the settlement calculated using the most
appropriate method (either the pressuremeter or oedometric method);
- under larger loads, the values for bearing capacity or sliding resistance using
conventional methods.
Such models can obviously also be used for SLS verifications.
2.2.1. Detailed numerical models for ULS verifications
2.2.1.1. Case of foundations built on inclusions
2.2.1.1.1. Underlying principle
If Design Approach 2 is selected, then the model allows establishing the applied load displacement curve of a characteristic point of the foundation, over an interval greatly beyond
the service load.
The shape of this theoretical load-displacement curve is examined to see if it reveals a limit
value Rc beyond which an increase in the ratio between displacements and applied forces is
noticeable. In the event such a threshold is defined, the value of the design strength Rd is:
Rd = Rc/JR
(5.6)
It is proposed herein to adopt the value JR = 1.4 for the partial resistance factor.
If the failure mechanism is associated with a sliding plane on the underside of the foundation,
then the value JR can be reduced to 1.1 in agreement with the conditions applicable to shallow
foundations.
Remark: Should the detailed models rely on a greater number of parameters, it is
important to recall that the use of this method is acceptable only if the tests
discussed in Section 2.2 have all been performed.
2.2.1.1.2. Application example
An equivalent approach has been adopted to justify the ICEDA project, which was an
exceptional structure subjected to some very specific requirements. This project was
described in the historical review presented in Chapter 1 of this guide.
A static load test area containing 9 inclusions has served to calibrate the constitutive laws of a
numerical model representing the test section, using a 3D finite element model.
A basic unit cell model using these constitutive laws was built in order to study the overall
behavior of the foundation, with special emphasis on the settlement / applied load relation.
A parametric study conducted on this model has enabled the determination of the sensitivity
of the settlement to successive variations of practically all the parameters at increasing applied
loads.
Figure 5.12 illustrates one of the results obtained: it depicts the settlement vs. load behavior of
a raft, for a given mesh and for the reference case (blue curve), as well as for the improved
172
characteristics of compressible layer (red curve) and degraded characteristics of this layer
(green curve). The settlement curve vs. loading indicates a change in behavior around an
applied load of 360 kPa on the raft.
In order to design the grid of inclusions for the ICEDA project under static loads, the forces in
the inclusions corresponding to a load of 360 kPa on the raft were factored using the JR
coefficients of 1.0 at ULS and 1.4 at SLS, in order to define the maximum allowable forces.
Once the ultimate value for the stress on the raft was defined, the grid of inclusions was
designed with respect to the SLS such that:
The calculated settlements remain less than the allowable settlement values;
The integrity of the load transfer platform is verified: absence of zones where the shear
strain (evaluated as the second invariant of the deviatoric strain tensor) would be greater than
or equal to 5% and moreover would:
- continuously connect the inclusion heads to the raft;
- continuously connect the inclusion heads to the natural soil, at the periphery of the
buildings;
- continuously connect two contiguous inclusions.
The structural integrity of the inclusions is verified, through the conventional methods
described in the EC7 draft "Deep foundations" Standard.
Remark: Calibration of the ultimate resistance value must be the same as for the
case of the footing on non-reinforced soil.
2.2.1.2. Case of embankments built on rigid inclusions
In this case, the verification step is carried out according to Approach 3.
Justifications
173
The ULS verification takes the form of a specific calculation during which the parameters for
WKH VWUHQJWKRIWKHVRLOVDUHGHJUDGHG compared to their characteristic values applying the
M2 set of partial factors:
JI JF= 1.25
Jcu = 1.4
JJ = 1.0
The M2 partial factors must be complemented by the factor applicable to the soil-inclusion
lateral interaction strength JM;f.
As presented in Section 2.1.1.3.3 and in order to ensure continuity with Standard NF P 94 262
(deep foundations), it is proposed to adopt, for the factor JM;f applicable to the shaft friction
resistance, the composite value (1.1 JR,d) where JR,d is the partial factor associated with the
"ground model" method in Standard NF P 94 262. The values of this factor depend on the
inclusion installation technique, as well as on the type and mode of interpretation of the
geotechnical data (with or without statistical analysis).
As opposed to the shaft friction, the interaction at the head or base of the inclusion does not
require, for this calculation, the introduction of a specific M2 coefficient. This interaction is
correctly described by the input parameters for the soils provided that the inclusions are
modeled as volume elements and not as "beam" elements (i.e. without thickness).
Remark: It should be recalled that for the inclusion, the behavior law must
necessarily introduce a failure criterion for the material in the inclusion (an elastic
model without any failure criterion is unacceptable).
The ULS verification therefore requires a specific calculation step that consists of ensuring
that an equilibrium has actually been reached when all M2 factors are applied to the soil
parameters, inclusion material and shaft friction.
The fact that the calculation performed using these new parameters actually converges on a
solution provides the desired verification.
Remark: It is to be noted that the displacements obtained have no physical
significance.
2.2.2. Detailed numerical models for SLS verifications (settlements)
The purpose here is to verify that Ed< Cd:
Ed and Cd are settlements (or deformations);
Cd is the allowable settlement (or deformation) defined by the Project Owner.
This advanced model is employed without factoring the strengths. Since the model takes into
account the full set of material failure criteria, the existence of localized plastic zones in the
soil is acceptable as long as the SLS settlement criterion remains satisfied.
Remark: In the case of the exceptional ICEDA structure, the previous criterion was
supplemented by a criterion limiting the development of a free plastic flow zone:
the absence of a continuous shear band (defined by a shear1 above 5%) connecting
either the inclusion heads to the raft, or the inclusion heads to the natural soil, at the
periphery of buildings, or two contiguous inclusions.
1
174
2.3. Specific calculation models based on the Yield design theory for ULS verifications
These models apply the Yield Design theory (Salenon, 1983) within its general framework:
Static and kinematic approaches;
Displacement field not reduced to a field of rigid blocks.
These models are capable of processing any type of loading cases (Q, T, M) by determining
the associated failure mode.
Remark: These models are to be used under exceptional circumstances and must
be calibrated by comparison with other approaches, as illustrated in the following
section.
2.3.1. Application example
This approach was applied to the exceptional structure of the foundations for the RionAntirion bridge, as described in Chapter 1 (Pecker, 1998).
An upper bound on the foundation capacity was established using the Yield Design Kinematic
approach by introducing the strength contribution of the inclusions into mechanisms such as
the one presented in Figure 5.13. Given the foundation caisson dimensions (diameter: 90 m),
this approach was developed exclusively in the form of a 2D plane strain model.
These mechanisms can be defined by 5 geometric parameters (Z or D, O, G P and H DV
identified in Fig. 5.13). These mechanisms juxtapose the motions of rigid blocks and blocks
under deformation fields compatible with the displacements of neighboring blocks.
The minimization step over the space defined by these geometric parameters (representing all
possible geometries of these various failure mechanisms) serves to establish the value of the
smallest load capable of leading to failure.
The theory indicates that this value constitutes an upper bound of the actual failure load: any
load greater than or equal to this value necessarily causes the failure of the system.
Justifications
175
The set of ultimate values obtained allows defining a surface within the space (Q, T, M) that
encompasses all loads that the system is able to withstand. Any loading represented by a point
located outside this surface causes failure.
A cross-section of this surface in a plane corresponding to a constant vertical force value Q
gives the shape of this limit surface (Fig. 5.14). This example illustrates the increase in
capacity obtained by installing inclusions compared to a situation with the same foundation
on non-reinforced soil.
The inclusions have served to raise the values of Tmax and Mmax that can be supported for a
given Q value. The vertical part of the limit curve with inclusions represents the
configurations in which failure occurs by sliding on the foundation's lower side.
The results from the Yield design calculations have proven to be consistent with those yielded
by finite element numerical models as well as physical models tested in a centrifuge. In this
special case, the general validation principle for detailed models presented in Section 2.2 has
thus also been respected.
2.3.2. Verification principle
The formal inequality Ed < Rd therefore consists of verifying that the point representing the
load Ed is within the volume Rd delimited by the limit surface:
Within this framework, Design Approach 2 consists of establishing the limit surface within
the Q, T, M space based on non-weighted soil parameter values, then positioning the
representative loading point obtained by applying factors JG = 1.35 or 1.0 and JQ = 1.5 or 0 to
the characteristic values Q, T, M;
Design Approach 3 consists of establishing the limit surface in the Q, T, M space based on
the soil parameters modified using the set of M2 partial factors:
JI JF = 1.25
Jcu = 1.4
JJ = 1.0
and then locating the representative loading point (characteristic values Q, T, M assigned the
A2 partial factors: JG = 1.0 and JQ = 1.3 or 0).
The distance from the representative loading point to the domain boundary characterizes the
VDIHW\UHVHUYH.
176
f*
f (t )
C
Min D cc k 3 ck ; D cc ck
; D cc max
JC
JC
JC
(5.7)
with:
Dcc: coefficient depending on the presence or absence of steel reinforcement (reinforced =
1, non-reinforced = 0.8);
JC: partial coefficient with a value equal to 1.5 at the fundamental ULS and 1.2 at the
accidental ULS;
Justifications
177
fck*: characteristic value of the compressive strength of the concrete, grout or mortar in the
inclusion, as determined based on the following formula:
1
(5.8)
f ck
inf fck (t ); Cmax ; fck
k1k2
fck: characteristic value of the compressive strength measured on cylinders at 28 days;
fck(t): characteristic value of the compressive strength measured on cylinders at time t;
Cmax: maximum compressive strength value taking into account the required consistency of
the fresh concrete, grout or mortar, depending on the technique used, as shown in Table
5.1;
k1: value depending on both the drilling method and slenderness ratio, as listed in Table
5.1.
Table 5.1: Assigned values of Cmax and the coefficient k1.
Case
1
2
3
4
5
Execution mode
Drilled inclusions with soil
extraction
Drilled inclusions using a hollow
auger with soil extraction
Drilled inclusions using a hollow
auger with soil displacement
Inclusions either vibratory driven
or cast in place
Incorporation of a binder with
the soil (treated soil columns, jet
grouting, etc.)
Cmax
(MPa)
k1
35
1.3
30
1.4
35
1.3
35
1.3
(*)
(**)
For a soil treated by jet grouting or with a tool that does not guarantee a homogeneous section
geometry, the k1 value is to be determined on a case-by-case basis k1 > 1.5.
k1 may be decreased by 0.1, only for drilled inclusions when the composition of the ground
layers guarantees stability of the outer borehole walls or when the inclusion is cased and
concreted in the dry (a guarantee of borehole wall stability must be demonstrated using the
procedure outlined in the EN 1536 bored pile execution Standard):
k2 depends on the slenderness ratio:
k2 = 1.05 for inclusions whose ratio of smallest dimension d to length is less than 1/20;
k2 = 1.3-d/2 for inclusions whose smallest dimension is less than 0.60 m;
k2 = 1.35-d/2 for inclusions combining the two previous conditions.
k3 depends on the type of control performed, as specified in Table 5.2 below.
178
k3 values
Domain 1
(inclusions
required for
stability)
Domain 2
(inclusions not
required for
stability)
Without test
With
reflection or
impedance
tests (a)
With quality
tests
With loadbearing
capacity tests
With reinforced
control tests (b)
0.75
**
1.2
1.4
0.65
0.85
1.4
1.5
1.7
The coefficients in Table 5.2 cannot be combined: for instance, when both load-bearing
capacity tests and reflection or impedance tests are performed, the k3 coefficient value equals
1.2 for Domain 1 and 1.5 for Domain 2.
/HWs recall the following definitions (see Chapter 8: Controls):
quality tests: static loading test conducted at the maximum value, as defined in Section
2.2.1.2 of Chapter 8;
load-bearing capacity test: static loading test conducted at the maximum value, as defined in
Section 2.2.1.3 of Chapter 8;
reinforced control test: as intended in the Eurocode 7 National Application Standard NF P
94 262 for Deep foundations, whenever applicable.
Example: /HWVFRQVLGHUDsmall-sized project with 1500 ml inclusions and without
any load-bearing test. Domain 1 inclusions will be controlled by reflection or
impedance tests; they are assigned a coefficient k3 = 0.75. Domain 2 inclusions
receive a coefficient k3 = 0.65 if not controlled by reflection or impedance tests and
a coefficient k3 = 0.85 otherwise.
3.1.1.2. ULS design compressive strength
At the ULS, the maximum compressive stress is limited to the computed value fcd, while the
average compressive stress on the compressed part of the cross-section is limited to a fixed
value of 7 MPa.
3.1.1.3. SLS design compressive strength
At the SLS, the maximum compressive stress in the concrete is limited to Min (0.6 k3 fck*, 0.6
fck) and the average compressive stress on the compressed part of the cross-section is limited
to (0.3 k3 fck*).
The value of k3 is defined in Table 5.2.
Justifications
179
180
Figure 5.15: Evolution of design shear strength vs. ULS normal compressive stress.
Justifications
181
182
The slab on grade is intended to be covered with a skid resistant layer either directly on the
slab or via a self-leveling product.
According to the DTU 13.3 guidelines and in order to fulfill the non-brittle condition in
tension, the minimum section area of steel reinforcement is 0.4% in each direction.
3.4.1.2.2. Tensile limit state for non-reinforced concrete in bending
The calculated tensile stress in bending at the serviceability limit state under the most
unfavorable action combinations listed in DTU 13.3 must verify the following condition for a
non-reinforced slab on grade:
SLS = 6.M/h Ic28
(5.9)
(M: moment per linear meter)
As an example, for a concrete compressive strength fc28 = 25 MPa, the limit tensile stress
HTXDOVSLS = 1.8 MPa.
It is also possible to refer to the limit splitting tensile strength value, ftsplitting28, at 28 days. In
this case, the condition to be satisfied becomes:
SLS = 6.M/h Itsplitting28
(5.10)
For a given concrete strength class (e.g. C25/30), depending on the concrete mix design and
the type of materials used, the characteristic splitting tensile strength can vary significantly,
HJ WR 03D UHVXOWLQJLQ D OLPLW VWUHVVSLS varying between 1.8 and 2.7 MPa. Splitting
tensile tests are required.
Note: The initial approach stemming from compressive strength has a higher
margin of safety since it takes into account the uncertainties between a concrete's
compressive strength and its tensile strength. In reality, no mathematical
relationship exists between these two magnitudes.
3.4.1.2.3. Bending tensile limit state of the concrete with metal fiber additives.
Steel fiber-reinforced concrete slabs on grade are assimilated with non-reinforced slabs for the
entire set of conditions listed in DTU 13.3. The specificities of these processes are governed
by technical memoranda issued by the CSTB Building Research Center. The main
exemptions are: the tensile stress strength of the composite concrete, concrete characteristics,
distances between interruptionsEHWZHHQSRXUVLQWKHFDVHRIa slab on grade without sawed
MRLQW, or more commonly called DMRLQWIUHHVODERQJUDGH), and aspects of the construction
method.
The orders of magnitude of the limit tensile stresses evaluated by these technical memoranda,
for fiber concentrations of between 20 and 40 kg/m3, currently stand at:
in the vicinity of joints (panel corner and edge): between 1.8 and 3.5 MPa;
over the main part of the panel: between 2.5 and 5 MPa.
7KHGHVLJQHUs attention must be drawn to the fact that all technical assessments of fibers are
based on the post-cracking behavior of the material; this behavior may lead to conflicts with
the requirement for no cracking, hence making the slab unfit for its final purpose.
Justifications
183
184
Justifications
185
Figure 5.17: Examples of modeled situations in the main part of the zone for point loads.
Figure 5.18: Examples of modeled situations in the edge zone for point loads.
The slab on grade supported on soil reinforced by inclusions is a structure addressed in the
DTU 13.3 document, yet its design over its main part requires incorporating the specific
nature of a non-uniform distribution of reactions on the ground slab lower face.
As such, the simplified method for evaluating deformations and loads, as explained in
Appendices C3.1 and C4.1.4 to C4.1.7 of DTU 13.3, proves to be insufficient.
The present recommendations propose two distinct methods based on an interpretation of the
full-scale experiments and the analysis of a large number of detailed, three-dimensional
computations (ASIRI, 2011):
An envelope PHWKRGEDVHGRQWKHFRQFHSWRIDGGLWLRQDOPRPHQWV;
An alternative method based on differentiated subgrade reaction coefficients.
Each of these methods serves to estimate the loadings of mechanical origin (related to loads).
The effects of shrinkage and thermal gradient will then be added (along with the appropriate
combination coefficients) to these loadings.
186
JT >
SJ
Soed
@ >
IRJT
Soed
IRSJ
@
IRJT Soed
IR
Soed
SJ
JT
SJ
ma
mb
(5.11)
mc
Uniformly-distributed surcharge, alternating loaded and non-loaded bands, superposition of point loads, etc.
Justifications
187
3.4.3.1.2. 'HWDLOHGVWXG\RIWKHWHUPVPDPEDQGmc
7KHPD term: Results of calculation of a slab on an equivalent homogenized soil
The soil reinforcement is represented by a homogenized soil profile, for which each layer is
charaFWHUL]HGE\DQHTXLYDOHQW<RXQJs modulus ESDVVRFLDWHGZLWKD3RLVVRQs ratio Q.
This is the case adopted when designing an industrial slab in accordance with the method
described in DTU 13.3 (Part 1, Appendix C), which allows taking into account the interaction
between applied loads and joints, based on their respective positions.
In this special case, during the design phase of the slab, it becomes necessary to perform the
analysis at the panel edge and at the corner in systematically considering the FXUOHG case as
well as the case where this curling effect is cancelled out by the subsequent to differential
shrinkage attenuation (load Qs equals 0 in the Appendix C design method) because this last
case may become the critical design case. It is therefore necessary to run two calculations in
SDUDOOHOFXUOHGZLWKRXWWKHLQIOXHQFHRIWKHLQFOXVLRQVDQGQRQ-FXUOHGZLWKLQIOXHQFHRI
the inclusions.
Remark: In the case of a uniformly-distributed load, as per Article 5.1.1.1 of Part
1, DTU 13.3, an isolated concentrated load with an intensity equivalent to the
distributed load value applied on a 1 m total area (with an impact area
corresponding to a 5 MPa pressure) must be considered in order to evaluate corner
and edge zone loadings.
The homogenized soil profile is selected from the study of a basic unit cell (inclusion and the
corresponding surrounding soil volume, load transfer platform and joint-free slab) subjected
to an equivalent uniformly-distributed load q0. This basic cell is modeled with a constrained
condition at the edge (zero lateral deformation). Any of the methods described in Chapter 3
may be used here.
The value of the load q0 should be representative of the average load case for the condition
being studied, on an area (see Fig. 5.19):
Corresponding to several grids of inclusions, in order to ensure that the behavior of these
inclusions is compatible with the zero lateral displacement boundary condition at the
periphery of the model;
In proportion with the reinforced soil thickness.
188
Settlement y at the model surface is used to define the equivalent oedometric modulus Eoedo*
over the total model height H:
Eoedo* = q0 H/y
(5.12)
This oedometric modulus can be split into at least three moduli corresponding to a three-layer
model bearing on a soil substratum assumed to be non-deformable:
Load transfer platform: height H1, modulus Eoedo1;
Reinforced soil: height H2, modulus Eoedo2;
Soil layer underlying the reinforced soil: height H3, modulus Eoedo3.
Since Eoedo1 and Eoedo3 are known values, the value Eoedo2 can be calculated from the following
formula:
H/Eoedo* = H1/Eoedo1 + H2/Eoedo2 +H3/Eoedo3
(5.13)
Justifications
189
Remarks:
H1 is the load transfer platform thickness (between the inclusion head and the
bottom of the slab);
H1 + H2 may be assimilated, indifferently, either as the sum of the load transfer
platform thickness + inclusions length, or as the distance separating the bottom of
the slab from the lower plane of equal settlement, located below the inclusion tip
(Fig. 5.21);
Selecting the thickness H2 as the compressible layer thickness, the inclusion
length or the depth to the lower plane of equal settlement placed (located slightly
below the tip of the inclusions) has no noticeable impact provided the value of the
equivalent modulus Eoedo* over the total model height remains unchanged. This
option however must be maintained for all of the calculations;
The design according to Appendix C of DTU 13.3 implicitly assigns an infinite
thickness to the deepest soil layer. Moreover, the geotechnical model typically
assumes that an uncompressible substratum has been reached at a given depth. It is
necessary to ensure the consistency of these two models, especially by verifying
that the two profiles considered yield identical settlements. This step therefore
leads to incorporating a layer with a very high modulus value into the
computational model derived from DTU13.3 at the base of the model, which in
turn simulates the incompressible substratum defined by the geotechnical engineer
(Fig. 5.20).
190
For calculations that rely on Appendix C of DTU 13.3 and are performed in a conventional
For calculations that rely on Appendix C of DTU 13.3 and are performed in a conventional
manner using the value Q = 0.35 for the soil, the equivalent oedometric moduli Eoedo* are
manner using the value Q = 0.35 for the soil, the equivalent oedometric moduli Eoedo
* are
FRQYHUWHGLQWRHTXLYDOHQW<RXQJs moduli E*:
FRQYHUWHGLQWRHTXLYDOHQW<RXQJs
moduli E*:
E* = Eoedo* * (1 +Q) (1 - 2Q)/(1-Q)
(5.14)
E* = Eoedo
(1 +Q) (1 - 2Q)/(1-Q)
(5.14)
Remark: The E* modulus corresponds to the E modulus listed in Appendix C of
Remark: The E* modulus corresponds to the Es smodulus listed in Appendix C of
DTU 13.3.
DTU 13.3.
The equivalent modulus values to be used in the specific slab on grade calculations must be
The equivalent modulus values to be used in the specific slab on grade calculations must be
evaluated on the basis of residual settlements experienced from the construction of the slab.
evaluated on the basis of residual settlements experienced from the construction of the slab.
The goal here consists of defining these modulus values for consistency with the slab on
The goal here consists of defining these modulus values for consistency with the slab on
grade, which only considers the loads directly applied to the slab.
grade, which only considers the loads directly applied to the slab.
Example:
Example:
x Hypotheses:
x Hypotheses:
/HWs take the example of an industrial slab on a fill layer with thickness h = 1.80 m and unit
/HWs take the example
of an industrial slab on a fill layer with thickness h = 1.80 m and unit
weight J = 20 kN/m3 3 and built on a soil previously reinforced by inclusions.
weight J = 20 kN/m and built on a soil previously reinforced by inclusions.
The average surcharge load is 30 kPa (including the weight of the slab).
The average surcharge load is 30 kPa (including the weight of the slab).
If the analysis of in situ soil consolidation rate allows the estimation of the percentage of
If the analysis of in situ soil consolidation rate allows the estimation of the percentage of
settlement reached under the embankment weight and prior to construction of the slab, then
settlement reached under the embankment weight and prior to construction of the slab, then
the equivalent modulus to enter into the ground slab design may be evaluated as follows.
the equivalent modulus to enter into the ground slab design may be evaluated as follows.
x Results of the basic unit cell calculation:
x Results of the basic unit cell calculation:
The calculation for a basic cell with inclusions incorporating all loads leads to the following
The calculation for a basic cell with inclusions incorporating all loads leads to the following
results:
results:
1) Total settlement under embankment weight q = 1.8 u 20 = 36 kPa: 'h = 1.4 cm;
1) Total settlement under embankment weight q1 1= 1.8 u 20 = 36 kPa: 'h1 1= 1.4 cm;
Justifications
191
192
Figure 5.22: Definition of the Mupper and Mlower moments calculated in a slab on grade on inclusions.
,Q SUDFWLFH WKH PD moments calculated from the homogenized model are therefore
increased by Mupper when they result in tension on the upper fiber and by Mlower when they
result in tension in the lower fiber.
Remarks:
If for the same structure, several separate load cases exist, it would be necessary
to conduct a basic unit cell calculation for each load case in order to determine the
corresponding additional moments;
For a load q0 with a limited extent, it is also possible to consider that the
moments Mupper and Mlower are proportional to the load, with all other parameters
being held constant;
The corrHFWLYH PE term is not applicable to moments calculated according to
Sections C4.2.3 and C4.2.4 of Appendix C, DTU13.3 whenever the edge (or
FRUQHU LV FXUOHG )LJ 5.23). The moments calculated with these methods only
pertain to the raised part of the panel which is not influenced by the inclusions.
Nevertheless, it may be necessary to study the influence of the load sharing
between soil and inclusions on the determination of value q0 used in the
determinationRIWKHPE term.
Justifications
193
On the other hand, it is also necessary to study the case where this curling of the
slab disappears subsequent to the attenuation of differential shrinkage (i.e. with
load Qs equal to 0, as per Appendix C, DTU13.3), and in that case, the corrective
PE term needs to be used.
7KHPF term: Interaction between the rigid inclusions and the joints
The presence of a joint in the slab produces the following effects:
No bending moment at the location of the joint;
The bending moment diagram is shifted in the vicinity of the joint.
In practice, the set of studies carried out within the scope of the National Project have led to
the conclusion that at the location of a joint and regardless of the type of loading, an upper
ERXQGYDOXHRIWKHFRUUHFWLYHPF term (i.e. specific to the interaction of the rigid inclusions
with the joints) may beFKRVHQDVWKHRSSRVLWHRIWKHPE term, which was established from
the basic unit cell calcuODWLRQ UHVXOWV 7KH FRUUHFWLYH PF term (maximum moment offset)
thus lies within the interval {-Mlower, -Mupper} (reminder: Mupper and Mlower are both algebraic
values).
Between the joints, this interval decreases with the distance to the joints exclusively. It is
equal to zero at a point on the ground slab sufficiently far from any joint. The overall shape is
presented in Figure 5.24.
194
Justifications
195
FRUQHU LV FXUOHG )LJ 5.23); the moments calculated with these methods only
pertain to the raised part of the panel which is not influenced by the inclusions;
On the other hand, it is also necessary to study the case where this curling of the
slab disappears subsequent to the attenuation of differential shrinkage (i.e. with
load Qs equal to 0, as per Appendix C, DTU13.3), and in that case, the corrective
PF term needs to be used.
The edges of the slab on grade (boundary of the structure) have a similar effect at
WKH MRLQWV7KH FRUUHFWLYHPF term must therefore also be taken into account in
DGGLWLRQWRWKHPE term.
Special case of continuous slabs RQJUDGHZLWKRXWVDZHGMRLQWVMRLQW-free slab on
JUDGH)
The design of slabs on grade with joints (typical spacing of 5 to 6 m between joints), which
involves the combined use of the three terms PD, PE and PF, leads to an envelope of
moment in all cases.
The special case of a slab made of reinforced concrete or with the addition of steel fibers and
without sawed joints (typical spacing of 25 to 35 m herein referred WRDVMRLQW-free slab on
JUDGH) can be desigQHG ZLWKRXW DFFRXQWLQJ IRU WKH PF term except in the vicinity of
pouring joint interruptions and the edge of the structure.
In this case, it has been observed in certain unique configurations (back-to-back racks and
alternation of loaded and non-loaded bands) that the moment obtained, when solely focusing
RQ WKH PD DQG PE terms, might not constitute a strict upper bound of the calculated
moment through a comprehensive 3D model.
This is illustrated in Figures 5.27 and 5.28, which compare for the geometry displayed the
PE values deduced from FLAC 3D calculations with the values of the moments Mupper and
Mlower calculated from the basic unit cell computation:
In the case of loaded and non-loaded bands (Fig. WKH PE values calculated by
FLAC 3D in the longitudinal direction of the loaded strip exceed the negative moment Mlower
value at the intersection of the loaded zone axis and the inclusion axis;
In the case of lRDGLQJYLDDUDFNV\VWHPWKHPE values calculated in FLAC 3D in the rack
direction exceed the Mlower value at location where the back-to-back racks are aligned with the
inclusions.
Consequently, it would be useful to apply a model coefficient JR;d = 1.75 on the corrective
PE term relative to the lower fiber over the typical part of the slab [Mupper; 1.75 u Mlower] in
all load cases (e.g. localized distributed loads, stationary or mobile point loads), with the
exception of the uniformly-distributed load case.
196
Figure 5.27: Comparison of "mb" values (along the Y direction) from FLAC 3D calculations
with the Mupper and Mlower values derived from a basic unit cell computation
(succession of loaded/non-loaded bands).
Figure 5.28: Comparison of "mb" values (along the Y direction) deduced from FLAC 3D computations
with the Mupper and Mlower values derived from a basic cell computation (loading via the rack bases).
3.4.3.1.3. Summary
A slab on grade supported by a network of rigid inclusions may be designed by adding the
results of the following calculation cases:
ma, which is the output from a model with joints lying on an equivalent homogenized
soil;
PE derived from the influence of rigid inclusions on the bending moment of a joint-free
slab on grade, with this influence being bounded by [Mupper; Mlower] (note: both Mupper and
Mlower are algebraic values);
PF generated from the interaction between rigid inclusions and joints, bounded by
[ Mlower; Mupper].
Justifications
197
In the general slab on grade case, the influence of the rigid inclusions interacting with the slab
joints always remains bounded by:
[Mupper - Mlower; -(Mupper - Mlower)]
)RUD JLYHQORDGFDVHWKHPEDQGPF terms are independent of the location of the load
with respect to the joints, as opposed toWKHPD term.
With better knowledge of the relative positions of inclusions and joints, it is possible to refine
the above interval (see Section 3.4.3.1.2 and Fig. 5.24).
When the design explicitly takes into account the relative position of loads, joints and
inclusions, these hypotheses must be listed in the contractual clauses in a detailed manner at
the design stage.
The design of slabs on grade with joints (typical spacing of 5 to 6 m between joints), which
involves the combined use of the three termVPDPEDQGPFOHDGVWRDQHQYHORSHRI
moment values in all cases.
In the special case of a slab on grade without sawed joints (i.e. joint-free slab on grade), the
PF term is equal to zero except in the vicinity of edges of the structure and pouring joints
interruptions. It is then necessary to apply a model coefficLHQWRIRQWKHFRUUHFWLYHPE
term for only the lower fiber of the slab [Mupper; 1.75 u Mlower] for all load cases, with the
exception of the uniformly-distributed load case.
3.4.3.2. Alternative method: subgrade reaction coefficient
According to this method, the ultimate objective is to define a computational model limited to
the slab itself, the interaction of the slab and the foundation being modeled by springs with
properties estimated for a given loading level.
The first step is to study a basic unit cell of ground improvement under an equivalent uniform
load of the same intensity as the load case being studied. Afterwards, a calibration step is
performed on the previous results in order to determine a distribution of subgrade reaction
coefficients. In a final step, these coefficients are incorporated into a model of a plate on
elastic supports for the purpose of estimating the moments in the slab under the project loads
(point and/or distributed loads).
198
This method consists of studying a basic unit cell subjected to an equivalent uniformlydistributed load q0 under the same conditions as those presented Fig. 5.29.
The axisymmetric model of this basic cell has a radius R.
This calculation gives the distribution of the vertical stress under the slab (Fig. 5.30) as well
as the slab bending moment diagram.
Figure 5.30: Vertical stress distribution characteristic directly under the slab.
199
with the corresponding FLAC 3D calculations; this rk value represents a diffusion cone with
an approx. angle of 6h/10v from the top of each inclusion.
Figure 5.33: Influence of the parameter rk on the moments Mupper and Mlower.
Justifications
201
202
Justifications
203
204
Point loads (i.e. loads due to traffic or storage) allow designing the pavement using properties
obtained at the top of the subgrade layer (as determined by plate load tests). These
characteristics depend on both the compaction and the thickness of the subgrade layer, but
tend to be only slightly influenced by the presence of inclusions.
In contrast, the pavement surface settlements must be calculated using a model that
incorporates inclusions.
Lawson (2000) proposed charts for designing the geosynthetic layer stiffness as a function of
acceptable differential settlements at the embankment surface (Fig. 5.38). These results may
be used for SLS situations, yet they cannot replace the settlement analysis performed using
conventional methods.
Justifications
205
Figure 5.38: Charts for geosynthetic design based on allowable differential settlement
(Lawson, 2000).
by Hewlett and Randolph (1988), and subsequently adopted in the BS8006 Standard (2010).
The Hewlett-Randolph method will be presented herein for illustration purposes: it is limited
to providing the stress distribution on the geosynthetics layer without taking into account
either the settlements or the inclusion friction phenomena.
In contrast, the method used to calculate the tension in the various layers, selected by these
Recommendations, is based on the BS8006 Standard Method (2010).
3.7.2. Function of geosynthetic layers
The reinforcement by geosynthetics can have two distinct and complementary objectives:
Ensuring the transmission of forces stemming from the weight of the embankment located
outside of the diffusion cones towards the inclusion heads (so-called stress approach), which
influences the optimal thickness of the load transfer platform;
AbsorbingWKHODWHUDOIRUFHVIURPWKHHPEDQNPHQW
These functions serve to determine the limit states needed to be verified.
3.7.3. Computational approach for the ULS verifications
Consistent with the EC7 NF P 94 270 Application Standard (2009) dedicated to reinforced
embankments:
Verifications pertaining to the geosynthetic reinforcement function at the base of the load
transfer platform (i.e. soil support between inclusions, lateral thrust, anchorage) are conducted
in accordance with the Design Approach 2 as intended in Eurocode 7;
For the general stability study of an embankment installed on compressible soil, as
performed with failure surfaces intercepting the geosynthetic reinforcement, the Design
Approach 3 must be employed.
3.7.675 verification
The tension Td calculated in the horizontal geosynthetic reinforcements must be less than the
computed long-term strength value Rt;d.
This tension is derived from the characteristic short-term strength value Rt;k (as measured
according to the NF EN ISO 10319 Standard), isochronous creep curves (established as per
the NF EN 13431 Standard) and the appropriate partial coefficients that serve to take into
account long-term product behavior:
Rt ;d
Rt ;k
J geo
This verification must be carried out in a ULS situation and, if necessary, a SLS situation (see
Section 3.7.5.4).
The safety coefficient on the geosynthetic material Jgeo can be expressed as:
Jgeo = JM;t . *creep . *deg . *dam
JM;t partial material factor set equal to 1.25 (as per the NF P 94 270 Standard).
The values of *creep , *deg and *dam are specific to each product and supplied by the
individual producer. Default values have been proposed in the NF P 94 270 Standard:
Justifications
207
*creep
partial coefficient correlated with geosynthetic behavior vs. time. Application
of this coefficient allows, over the entire structural use period, taking into consideration the
influence of creep on the tensile strength of geosynthetic reinforcements as well as limiting
structural deformations.
*creep =1/Ucreep, as defined in the NF P 94270 Standard, Appendix F.4.3.
*deg partial coefficient correlated with aging of the geosynthetic products, e.g. by means of
hydrolysis or oxidation, given the product's environmental conditions.
*deg =1/ Udeg, as defined in the NF P 94270 Standard, Appendix F.4.4.
*dam partial coefficient corresponding to the geosynthetic reinforcement damage produced
by their installation along with embankment compaction.
*dam =1/ Udam, as defined in the NF P 94270 Standard, Appendix F.4.2.
Remarks: The NF P 94270 Standard serves to define the partial coefficient Ucreep
for the SLS and ULS limit states, by taking into account both the creep failure
criterion and allowable creep deformation criterion.
The partial coefficient *dam corresponding to damage may vary with respect to the
deformation. At the SLS, the value corresponding to the geosynthetic service
deformation would be a suitable choice.
The other partial factors (JM;t and *deg) are identical at both the SLS and ULS.
3.7.5. Tensile forces under vertical loading
3.7.5.1. Review of notations
HM:
s:
a:
embankment height,
center-to-center spacing between inclusion,
inclusion head dimension (assuming a square inclusion head).
In the case of a circular inclusion head of diameter D, the equivalent square dimension is set
such that:
a
S D2
| 0 .9 D
4
(5.18)
These methods make it possible to calculate the average vertical stress Vv,geo applied to the
soil between the inclusion heads.
As an example of other potentially effective methods, Hewlett and Randolph's (1988) will
now be presented.
x Hewlett and Randolph method:
Figure 5.40: Schematic diagram of the Hewlett and Randolph method (extracted from BS8006-2010).
This method considers hemispherical domes for the purpose of calculating load transfer
efficiency, by focusing on failure at the key of vaults or at the inclusion heads (Fig. 5.40). For
thin embankments (relative to the inclusion spacing), stability at the keystone typically
governs the mechanism, whereas for thicker embankments, stability at the inclusion head
level becomes predominant.
This method calculates the average vertical stress applied to the geosynthetic by evaluating
the efficiency of load transfer towards the inclusions, in considering system failure either at
the top of the arch (i.e. efficiency EV) or at the level of the inclusion heads (efficiency ET).
The selected efficiency is the lower of the two, thus maximizing the load transferred to the
geosynthetic reinforcement.
Emin = Min (EV, ET)
x Efficiency EV:
Ev
a 2
1 1 A AB C
s
Justifications
(5.19)
209
2( K 1 )
p
a
1 s
s 2K p 2
2 H 2 K p 3
s a 2 K p 2
2 H 2 K p 3
(5.20)
(5.21)
(5.22)
x Efficiency ET:
E
1 E
ET
E
(5.23)
a K p
a
1 K p
1
s
s
a
K p 1 1
s
2Kp
(5.24)
The vertical load to be absorbed by the geosynthetic between two adjacent inclusions equals:
Vv, geo = s2/(s2 - a2) Vv (1 - Emin)
(5.25)
with: Vv = JJ Jr H + JG g + JQ q
(5.26)
HM: embankment height above the geosynthetic
Jr:
unit weight of the embankment
g:
permanent excess vertical load
q:
variable excess vertical load
JJ:
partial factor on the soil unit weight
JG: partial factor on the permanent actions
JQ: partial factor on the variable actions
3.7.5.3. Tensile force Td;mesh in the geosynthetic
The main principle of this method assumes that the entire vertical load acting between two
inclusions is aEVRUEHGE\WKHPHPEUDQHHIIHFW, resulting in tension in the geosynthetic and
in considering that no load is being transmitted to the soil between the inclusions.
The BS8006:2010 Standard gives the relation between the tension in the geosynthetic T and
both WT, the vertical load between two inclusions, and the geosynthetic deformation Hd:
( s a ) s V v ,geo
( s a)WT
1
1
Td ;mesh
1
=
1
(force per linear meter)
(5.27)
2a
6H d
2a
6H d
with: WT Vv, geo (force per linear meter)
This equation with two unknowns Td;mesh and d can be solved either by using the forcedeformation law of geosynthetics (with introduction of the tensile stiffness of the
geosynthetic) or by taking into account a maximum deformation criterion (called the
reference deformation).
3.7.5.4. Reference deformation
It is necessary to consider the deformation of the geosynthetic to calculate the mobilized
tension T d;mesh.
210
In a ULS situation, the reference geosynthetic deformation is the one selected by the BS8006
Standard (2010), i.e.: Hd = 6%.
For a load transfer platform thickness HM > 0.7 (s-a), the reference deformation must not
exceed this maximum long-term value of 6%.
For thinner embankments (e.g. when the condition HM > 0.7 (s-a) has not been satisfied) and
for which the load transfer is not guaranteed, this reference deformation must be lowered to
values of less than 3% in order to reduce the risk of differential deformation at the load
transfer platform surface. In this case, verification in an SLS situation becomes necessary to
ensure that the differential deformation at the load transfer platform surface is acceptable.
It must be noticed that the value of the partial coefficient Jcreep also depends on the choice of
the reference deformation.
The reference geosynthetic deformation must include the initial deformation as well as the
creep deformation: the initial geosynthetic deformation is necessary to develop reaction
forces. Moreover, the long-term deformation (due to creep) must be limited in order to ensure
that localized, time-delayed deformations do not occur on the embankment surface.
The post-construction creep deformation of the geosynthetic must be reduced to a minimum
so as to prevent localized differential deformations at the embankment surface, while not
exceeding 2% over the structural life cycle. This deformation is determined on the basis of
indications provided in the ISO/TR 20432 guide.
These two criteria would neeGWREHYHULILHGRQWKHSURGXFWs isochronous load-deformation
curves (Fig. 5.41).
Figure 5.41: Verification of deformation criteria and choice of partial factor Jcreep on
the geosynthetic isochronous load-deformation curves (according to the NF P 94 270 Standard).
Justifications
211
S Id '
tan
4
(5.29)
Ka is calculated with the design value Id (using the Design Approach 2, the value of JI equals
1 meaning that the design value IdHquals the characteristic value):
Jr:
unit weight of the embankment;
g:
permanent excess vertical load;
q:
variable excess load;
JJ:
partial factor on the soil unit weight;
JG: partial factor on the permanent actions;
JQ: partial factor on the variable actions.
Beneath the embankment slopes, the tensile force Td;thrust mobilized in the geosynthetic by the
embankment lateral active pressure decreases with distance from the slope crest; this force
may be calculated at any point based on the embankment slope and thickness above this given
point.
Remarks: The coefficient Jcreep required to determine the design long-term strength
Rt;d is obtained by adopting the same reference deformation value as that used for
vertical loads (Section 3.7.5.4).
It is important to note that this configuration may also induce lateral displacements
of the compressible soil capable of creating bending moments in the inclusions that
also require verification.
212
tan I geo
tan I' d
(5.33)
I: design value of the angle of friction of the soil atop the reinforcement layer (according
to the Design Approach 2, the value JI equals 1, and the design value Id equals the
characteristic value)
I: design value of the angle of friction of the soil-geosynthetic interface
JR,h: partial factor of the sliding resistance (equal to 1.1 with the Design Approach 2 as per
EC7)
JR,f: partial factor of the soil-reinforcement bond resistance.
It is proposed herein to adopt the value JR,f = 1.35, in compliance with the NF P 94 270
Standard (Design Approach 2).
Justifications
213
x,QWKHHPEDQNPHQWs longitudinal direction, length LaL, as measured from the last inclusion
needed for mobilization of anchorage forces by friction on both the upper and lower faces of
the geosynthetic, is such that (BS8006, 2010):
LaL t
I':
I':
h:
Td ;mesh
R ;h
(5.34)
J R; f
design value of the angle of friction of the soil above the geosynthetic reinforcement
design value of the angle of friction of the soil below the geosynthetic reinforcement
average height over length LaL.
x In the HPEDQNPHQWs transverse direction, length LaT, as measured from the last inclusion
(Fig. V.42) needed for mobilization of anchorage forces by friction on both the upper and
lower faces, is such that (BS8006, 2010):
LaT t
Td ;mesh Td ;thrust
R ;h
(5.35)
J R; f
irrelevant since settlement at the tip is not necessarily limited to 10%. In order to
introduce other Rb values than those calibrated on a 10% settlement, it is required
to interpret the results of an instrumented load test on an inclusion of the same type
as those actually installed.
The kp or kc value depends on the relative embedment depth, according to the
relationship given in the NF P 94 262 Standard.
4.1.2. Behavior along the shaft
4.1.2.1. Positive friction in the bearing layer and in the compressible soils below the neutral
plane
It must also be verified that the value of friction mobilized under the neutral plane does not
exceed the limit value qs calculated in accordance with the principles set forth in both
Eurocode 7 and the National Application Standard (NF P 94 262):
The limit value of lateral friction qs is evaluated by either the penetrometer or pressuremeter
method;
The qs values are listed in the NF P 94 262 Standard or else in the specifications intended
for the given technique; otherwise, experimental values established onsite are to be used.
Remark: It is to be observed that friction mobilized in the vicinity of the neutral
plane tends to be small, independently of the maximum friction at this elevation
(Fig. V.43). In some cases, it may be preferable for the load-bearing capacity
verification not to consider the total length below the neutral plane, but instead a
reduced length from the tip.
Justifications
215
Remarks:
The choice of the KtanG reference value must be made judiciously since this factor
is pivotal to the overall design and influences to a different extent the maximum
force in the inclusion and the resultant settlement.
This verification is mandatory when introducing finite element modeling; it
consists of verifying values of the ratio WVv at all points on the inclusion above
the neutral plane.
4.2. Load transfer platform behavior
This verification consists of establishing, for the applied external loading, the maximum
allowable stress value at the inclusion head that's compatible with both the load transfer
platform material characteristics and the structural geometry.
4.2.1. Method justification
4.2.1.1. Experimental results
x Load transfer mechanism on the inclusion head:
The mobile bottom tray tests in a centrifuge gave, for various geometric configurations, the
maximum values reached at the inclusion head (Okyay, 2010). The results obtained with a
granular load transfer platform placed beneath a loaded rigid plate (ground slab analogy)
provide a very worthwhile set of results, as they correspond to a material whose shear
characteristics were experimentally measured for different in situ densities of the material.
These results were obtained for various area replacement ratios D and load transfer platform
thicknesses HM (with HM/D = 1.4 to 2, with D being the inclusion diameter). Figure 5.44
presents an example of the variation of the load transfer on an inclusion head in the case of a
ground slab.
Figure 5.44: Load transfer ratio on an inclusion head vs. relative settlement y/D
during a centrifuge test with a mobile bottom tray.
216
Figure 5.45: Characteristic curves of shear tests on the material composing the load transfer platform
for moving bottom tray tests
Justifications
217
218
I ZKLFKLVUHSUHVHQWDWLYHRIWKHFULWLFDOVWDWHLHWKHVWDWHLQZKLFKWKHPDWHULDOVKHDUV
at constant volume) turns out to be better adapted to the final situation.
Figure 5.47: Comparison of calculated and measured values of limit stress on the inclusion head.
4.2.1.3. Conclusions
The limit stress value at the head of inclusions placed at the center of a mesh may be
determined on the basis of Equations 5.36 and 5.37, by adopting for the Nq coefficient value
the calculation result using the angle of friction at the critical state (i.e. shear at constant
volume). This value is less than the peak friction value for materials with a Relative Density
Index (ID) above 60%.
It is suggested herein to introduce the Nq coefficient in association with the shape factor value
sq previously proposed by Terzaghi, which equals 1 regardless of the foundation shape. This
choice has been adopted in order to compare the experimental results with theoretical values.
The Nq value associated with the Prandtl mechanism is representative of situations in which
the load transfer platform thickness exceeds the height Hmax for the full failure mechanism to
develop. The ratio Hmax/D increases with the angle of friction, as illustrated in Table 5.3.
For HM/D less than Hmax/D, the apparent load-bearing capacity coefficient Nq* would be more
favorable than the coefficient derived by the Prandtl mechanism, and it remains a safe
assumption to neglect this increase.
Justifications
219
I
Nq
Hmax/D
Lmax/D
L1/D
30
18.4
1.59
4.29
2.64
35
33.3
1.90
5.77
3.39
38
48.9
2.15
7.00
4.00
40
64.2
2.35
8.01
4.51
In a comparable manner, it may be noted that for a distance between inclusions of less than
2L1 (Fig. 5.48), the failure surfaces associated with two neighboring inclusions overlap. For
meshes smaller than this threshold, the factor Nq* is thus also higher and moreover neglecting
this increase would not compromise safety.
The width L1 is therefore the value to take into consideration when verifying the possible
interference between inclusions, as illustrated in Figure 5.49 showing the results for a finite
element model of an elementary unit cell modeling the mobile bottom tray system (Okyay,
2010). The domain of plastic points (shown in red) is actually very similar to Domain L1 of
the Prandtl mechanism.
Figure 5.49: Numerical simulation of a moving plate test below a slab on grade,
demonstrating the extent of the plastic zones.
4.2.2. Computation of limit transfer at the inclusion head over the general main section
4.2.2.1. Introduction
In this section, let's consider a network of rigid inclusions with diameter D = 2.rp installed on
a square mesh pattern of dimension s, covered by a load transfer platform of thickness H,
defined by LWVLQWULQVLFSDUDPHWHUVFIJ) and exposed to a uniformly-distributed load q0.
220
Remark: The load from the load transfer platform weight J.HM is separated here
from the loading q0 applied on the load transfer platform (eventually composed of
both permanent loads and operating loads).
When the load is applied on the system and depending on the geometry and nature of this
loading, two types of axisymmetric limit equilibrium can be generated between the applied
stress q0, the stress at the inclusion head qp+ and the stress applied on the soil between
inclusions qs+.
x 3UDQGWOs diagram:
The first limit equilibrium diagram within the load transfer platform may be modeled by the
Prandtl diagram, which associates a Rankine active limit state domain (I) above the inclusion
head, with a domain delimited by a logarithmic spiral arc (II) and another Rankine passive
limit state domain outside the inclusion head (III).
Justifications
221
x Shear cone:
The second limit equilibrium diagram within the load transfer platform can be modeled by a
vertical cone section opening onto the load transfer platform surface, originating from the
inclusion head and forming an angle I' with respect to the vertical plane equal to the angle of
friction of the load transfer platform material.
Figure 5.52: Failure diagram in the load transfer platform using a shear cone type model.
For a given external load q0, these two types of limit equilibrium serve to define the
maximum load that may be concentrated at the inclusion head qp+, i.e. the maximum system
efficiency at the level of the inclusion head/HWs note that an increase in load q0 allows an
increase in stress at the inclusion head qp+, and that the failure of the system will only occur
when it no longer becomes possible to increase the applied load q0.
Remark: The tests and experiments conducted in the centrifuge and on discrete
models have demonstrated that:
The Prandtl model failure diagram is applicable when considering the material's
critical angle (large deformations within a compacted material) (Section 4.2.1.2);
The shear cone failure diagram is associated with the peak angle of the material
(mobilized for small deformations).
9HULILFDWLRQRI*(2 limit states at the ULS
This verification is performed by implementing Design Approach 2 from EC7, with the
following combination of partial factor datasets:
$05
x For the Prandtl diagram:
According to Design Approach 2 from EC7 and based on Prandtl's diagram, the limit stress at
the inclusion head qp+ can be determined from: the stress applied on the supporting soil qs+;
WKHVHWRILQWULQVLFORDGWUDQVIHUSODWIRUPSDUDPHWHUVFM' and J; and the following formula:
qp
s q .N q u q s s C .N c u
c'
J
s J .N J .r p u
J c'
JJ
(5.38)
where:
sJ, sc, sq are the inclusion head shape coefficients
222
Nq
Nc
NJ
N 1u cot M ' J
2.N 1u tan M ' J
M'
M'
If H M H C
qp
HM
3
where :
R rp
tanM'
RC
c'
R J RC
1 RC
1 C u
u q0
1 u
rp
J c'
r p J J
rp
tanM' r p
(5.39)
M'
r H M u tan
J
I'
R rp
If H M ! H C
:
tanM'
RC
qp
H
c
3
1 R
R
c'
R
R J R
1 H M H C
u q0
1 u
u
rp
J c'
rp
rp J J rp
tanM' rp
Justifications
(5.40)
223
where:
R
R J J Rs J R ,d J s
Min b R ,d b
; f cd
2
S u rp
(5.42)
where, referencing the National Application Standard 94.262 and the "ground model"
approach:
Rb is the ultimate soil resistance at the tip of the inclusions;
Rs is the ultimate shaft friction under the neutral point;
Jb = Js = 1.1 at the ULS (set R2 defined in EC7-1);
JR,d = 1.25 (model coefficient);
fcd as defined in Section 3.1.1.
Therefore, in the (qs+; qp+) plane and independently of the load level, the domain of allowable
stresses at the ULS in the load transfer platform is determined by Equations 5.38, 5.41 and
5.42, which correspond respectively to curves 1a, 2 and 3 (shown in Fig. 5.54).
224
Therefore, in the (qs+; qp+) plane and independently of the load level, the domain of allowable
stresses at the ULS in the load transfer platform is determined by Equations 5.38, 5.41 and
5.42, which correspond respectively to curves 1a, 2 and 3 (shown in Fig. 5.54).
Under certain conditions, this domain may also be partially limited by the curve (1b)
associated with Equation 5.39 or 5.40 corresponding to a shear cone opening onto the load
transfer platform surface. This additional limit only exists as a design parameter for thin load
transfer platforms that are not topped by any rigid structural element such as a raft, a slab on
grade or a footing.
Figure 5.54: Allowable domain of stresses at the load transfer platform base.
Remark: For illustration purposes, the curves in Figure 5.54 above (1b) associated with the
Equation 5.39 or 5.40 correspond to the case of a thin embankment not covered by any rigid
structural element, with the additional constraints:
Load resulting exclusively from the weight of a load transfer platform with variable
thickness (shown in red);
The load transfer platform has a given thickness and an external load applied at the surface
(blue).
4.2.2.3. Load conservation principle
For a given load q (for which the partial factors of the set A1, as defined in EC7 have been
applied), stresses q0, qs+ and qp+ are in equilibrium and required to satisfy the load
conservation equation:
q 0 J G . J .H M
JJ
2
u s
q s u s 2 S.rp q p u S.rp
(5.43)
The mobilized stress parameters (qs+, qp+) must simultaneously satisfy the load conservation
equation (5.43) while falling within the allowable domain defined above in Figure 5.54.
For a given load q0, the allowable limit stress in the load transfer platform at the inclusion
head qp;d+ is thus determined by solving the system of equations associated with the curves
(1a), (1b), (2), (3) and (4). The principle is illustrated in Figure 5.55. For this load q0, the
allowable domain can be reduced to the intersection of segment (4) with the shaded surface
(delimited by curves derived from Equations (1a), (1b), (2) and (3)).
Justifications
225
Figure 5.55: Determination of qp;d (case where qpmax < sq.Nq.qsmax + sc.NcFJF)
The calculated design limit qp;d+ thus depends on the structural load q0, but remains
independent of the deformability of the various soil layers: this value solely depends on the
applied load q0, system geometry (load transfer platform thickness HM, mesh size s, inclusion
GLDPHWHU'DQGWKHLQWULQVLFSDUDPHWHUVFIJ) of the load transfer platform material.
In contrast, the value (qs+; qp+) actually mobilized at the load transfer platform base lies inside
the allowable domain, on segment (4) of the load conservation line, at a position that depends
on the deformability of the various soil layers. This value turns out to be heavily dependent on
the deformability of the soil layer located directly beneath the load transfer platform; the
greater WKLV OD\HUs compressibility, the closer the mobilized value moves towards the
allowable limit value.
4.2.2.4. System efficiency and load-bearing capacity
The system efficiency at the load transfer platform base is used to measure the proportion of
total load in the inclusion; this efficiency is expressed as follows:
q p u S.rp
q0 J.H M u s 2
(5.44)
The maximum efficiency that can be mobilized by the soil reinforcement at the level of the
load transfer platform base (eventually corresponding to the line in Equation (1), (3) or (1b))
decreases with the applied load, whereas the efficiency actually being mobilized during
loading tends to increase with the applied load.
A variation in external load q0 moves the point representing the equilibrium in the plane (qs+;
qp+) along a curve that, for high loads, tends towards the asymptotes of Equation (1), (1b) or
(3). In other words, the increase in applied external load q0 increases the efficiency towards its
maximum value yet can never cause internal failure of the load transfer platform by
intersecting with the equation (1) line (Fig. 5.56).
226
Figure 5.56: Evolution of the equilibrium point (qp ;qs ) vs. applied load
The maximum allowable loading on the load transfer platform + reinforced soil system
corresponds to the uniform stress qmax applicable to the system prior to its failure. This
maximum load (system load-bearing capacity) results from summing two terms:
qmax s 2 Vv ,d s 2 S.r 2 q p ;u S.r 2
(5.45)
The potentially mobilized stress at the level of the inclusion head qp;u can then be determined
by employing the following relation:
q p ;u
c'
J
Min s q .N q u V v ,d sC .N c u
s J .N J .r p u ; q p max
J c'
JJ
(V5.46)
In the specific example illustrated in Figure 5.57, the maximum load corresponding to a
uniform stress qmax on the mesh results in a stress Vv;d applied to the soil around the inclusion
and the stress qp,max on the inclusion head.
Figure 5.57: Maximum stress qmax on the load transfer platform + reinforced soil system.
(Note: The case illustrated corresponds to one in which qp;u = qp max.)
r0 u e
Justifications
S M
tan M ' T
4 2
(5.47)
227
Depending on the intrinsic load transfer platform parameters, the main dimensions of this
logarithmic spiral are as follows:
with:
height of the Rankine active limit state domain:
D
h1
u tan S 4 M ' 2
2
the extent of Rankine active limit state domain:
tan M ' .
r0
L
h2
S
2
Due
2 u cosS 4 M ' 2
(5.49)
2 u r0 u cos S 4 M ' 2
r0 u e
S M '
tan M ' . T max
4 2
tan M ' .
cos S 4 M ' 2
2
Due
cos S 4 M ' 2
. sin T max h1
S M '
tan M ' . T max
4 2
d 1 r0 u e
. cosT max
the angle at the top of the logarithmic spiral:
T max
(5.48)
1
tan 1
'
tan M
(5.50)
(5.51)
(5.52)
(5.53)
x Application example:
With the following hypotheses:
D = 0.30 m
s = 1.75 m
q0 = 50 kPa
C = 0 kPa
M = 38
the following values are obtained:
qp;d+ = 1,160 kPa
228
h1 = 0.31 m
h2 = 0.34 m
L = 2.10 m
d1 = 0.50 m
In order to allow the logarithmic spiral to fully develop within the load transfer platform, the
platform must have the following minimum dimensions:
HM > h1 + h2
d>L
Remarks:
For the case of thin embankments not covered by a structural element (such as a
footing, slab on grade, raft, etc.), then these criteria are used to determine a
minimum thickness Hmin WR EH UHVSHFWHG LQ RUGHU WR KRSH to optimize system
efficiency. Otherwise, the load potentially mobilized at the inclusion head (i.e.
efficiency) is reduced and failure occurs due to shear around a cone developing to
the upper embankment surface. The limit load is controlled by the cone weight to
which the double effect of cohesion and surface load must be added. Experiments
have led to the conclusion that the cone angle is equal to the peak friction angle of
the material.
For thin load transfer platforms (HM < h1 + h2) under a structural element (e.g. a
footing, slab on grade or raft), since in this case the failure mechanism can only
develop over a reduced height, then the potentially mobilized load at the inclusion
head is greater than the load obtained using the Prandtl diagram. Without a detailed
study, the use of the value obtained with the Prandtl scheme for load transfer
platform thicknesses of less than (h1 + h2) will lead to a safe design.
4.2.3. Calculation of the limit load transfer on the inclusions at the edge of the structure
4.2.3.1. Case of zero footing overhang
4.2.3.1.1. Presentation
The configuration adopted herein is that depicted in Figure 5.59. The edge of the inclusion
coincides with the edge of the footing. The load applied on the footing turns out to be nearly
entirely retransmitted onto the inclusion head. The vertical stress applied on the soil at the
level of the top of the inclusions is the stress JH generated from the surrounding ground.
Justifications
229
Figure 5.59: Extreme case of an inclusion at the footing edge with an overhanging load transfer
platform.
This value JH only depends on the geometry and not on the footing load; it controls the
failure mechanism capable of being developed from the top of the inclusion The limit value of
the qp+/JH ratio defines the coefficient Nq* associated with this case.
4.2.3.1.2. Nq* evaluation method
It is possible to determine the value of the coefficient Nq* simply by decomposing the
mechanism into two juxtaposed failure mechanisms: an active limit state equilibrium (above
the inclusion head), and a passive limit state equilibrium (elsewhere), as illustrated in
Figure 5.60.
Following the remark presented in Section 4.2.1.2, this analysis may be conducted as a plane
strain problem.
Figure 5.60: Decomposition into an active limit state equilibrium and a passive limit state equilibrium.
230
Since the equilibrium considered herein are those of a weightless medium, the intermediate
stress q in the plane separating the two domains is uniform. This stress q represents
respectively:
Volume 1 (left side of the figure): active earth pressure at the angle IGXHWRVWUHVVTp+ in the
weightless medium:
q = qp+ Kq1
(5.54)
Volume 2 (right side of the figure): passive earth pressure reaction at the angle IGXH WR
stress JH in the weightless medium:
q = JH Kq2
(5.55)
The coefficient Kq1 (respectively Kq2) is the surcharge pressure transmission coefficient in a
weightless medium relative to the active limit state (respectively, passive limit state). Its
analytical expression was established by Caquot and Krisel (1966) in the most general case
(Fig. 5.61). These expressions become simplified in the present case in recognizing that:
G0 = E = O = 0.
By applying Equations 5.55 and 5.56, it results that the ratio qp+/JH is equal to Kq2/Kq1; this
ratio is therefore similar to a load-bearing capacity factor Nq*.
Remarks:
This approach may be extended to the case of a material exhibiting cohesion
&DTXRWV&RUUHVSRQGLQJ6WDWHVWKHRUHP2QO\ the natural soils (and not the load
transfer platform) may require introducing an additional cohesion term.
Neglecting natural soil cohesion constitutes a safe assumption.
Justifications
231
232
4.2.3.1.3. Case of a zero footing overhang and an extended load transfer platform
The failure mechanism develops within a homogeneous medium. The Prandtl solution
therefore is directly applicable.
Figure 5.62: Configuration of the zero footing overhang and extended load transfer platform.
It is worthwhile however to compare the value Nq(Prandtl) with the value obtained using the
previous approach consisting of breaking down the problem into two juxtaposed states of
equilibrium.
The inclination G of stress q in the vertical plane separating the two domains is chosen equal
to the angle of friction IRIWKHORDGWUDQVIHUSODWIRUPLHIDLOXUHVLWXDWLRQ
Table 5.4 confirms that it is indeed acceptable to assimilate the term Nq* = Kq2/Kq1 with
PrDQGWOs load-bearing capacity factor.
Table 5.4: Comparison of Nq* values with Prandtl's load-bearing capacity factor Nq.
I1
Nq
(Prandtl)
30
33
35
38
40
18.40
26.09
33.30
48.93
64.20
Kq1(I1
G/I1
0.315
0.285
0.266
0.240
0.224
Kq2(I1,
G/I-1)
5.804
7.425
8.850
11.745
14.393
Nq*= Kq2/Kq1
Nq*/Nq
18.43
26.05
33.27
48.94
64.25
1.00
1.00
1.00
1.00
1.00
This observation shows that a decomposition of the bearing capacity problem into two
juxtaposed active - passive limit states of equilibrium constitutes a valid approximation.
It is proposed herein to extend this approximation to the case where the two domains are
made of different materials.
Justifications
233
4.2.3.1.4. Case of a zero footing overhang and a platform limited to the footing footprint
Figure 5.63: Configuration of the zero footing overhang and a platform limited to the footing footprint.
In this situation, Volume 1 is characterized by the angle of friction of the granular platform
(I1DQG9ROXPHE\WKHDQJOHRIIULFWLRQRIWKHVXUURXQGLQJVRil (I27KHLQFOLQDWLRQRI
stress q in the vertical plane separating the two volumes must be set equal to the smallest
angle of friction, which in theory would be the friction angle of the surrounding soil.
This approach was applied combining a range of values for I1 >- 40] for the platform
and another range of values I2 >- 30] for the surrounding soil.
Table 5.5 summarizes the Nq* values over the two defined intervals.
Figure 5.64 displays the same results graphically and then compares them to the Nq values
that would have been obtained with a load transfer platform extending beyond the limits of
the structure.
For a platform whose angle of friction is I1 WKHFRHIILFLHQW1q applicable to the main
central portion of the structure, equal to 33.3, is divided for the extreme case of an inclusion
placed at the edge of the structure by a factor of 2 or 3, depending on whether the angle of
friction of the surrounding soil is 25 or just 20.
234
Justifications
235
Nq(I1
18.4
26.1
33.3
48.9
64.2
Load
transfer
platform
I1
30
33
35
38
40
Kq1(I1 Kq2(IG
G/I2 1) I )
0.305
2.129
0.271
2.129
0.250
2.129
0.220
2.129
0.202
2.129
Soil I2 = 15
6.98
7.86
8.52
9.68
10.54
N q*
Kq1(I1 Kq2(I'G
G/I2 1) I )
0.304
2.872
0.270
2.872
0.249
2.872
0.220
2.872
0.201
2.872
Soil I2
9.45
10.64
11.53
13.05
14.29
N q*
Kq1(I1 Kq2(I'G
G/I2 1) I )
0.306
4.002
0.272
4.002
0.250
4.002
0.221
4.002
0.203
4.002
Soil I2
13.08
14.71
16.01
18.11
19.71
N q*
Kq1(I1 Kq2(I'G
G/I2 1) I )
0.315
5.804
0.278
5.804
0.256
5.804
0.225
5.804
0.207
5.804
Soil I2
Table 5.5: Values of the Nq* factor for a load transfer platform of limited extension (angle of friction I1LQDVRLODQJOHRIIULFWLRQI2.
18.43
20.88
22.67
25.80
28.04
N q*
As a consequence, these values indicate that under the same load transfer platform conditions
with I1 WKH OLPLW VWUHVV RQ DQ LQFOXVLRQ SODFHG DW WKH HGJH RI D VWUXFWXUH DQG DW P
below the surface of the surrounding soil does not exceed 160 kPa (soil I2 RU HYHQ
115 kPa (for a soil with I2
Figure 5.64: Comparison of Nq* values (case of a zero overhang platform) with Nq values
for a load transfer platform extending beyond the edge of the structure
Nq
D ( N q 1 ) 1
q0 J M H M
(5.56)
where:
HM:
load transfer platform thickness between the inclusion head and the footing;
I
angle of friction of the load transfer platform;
JM:
unit weight of the load transfer platform.
236
Justifications
237
Figure 5.66 presents, for illustration purposes, the limit stress values qp+ derived by this
approach for the two extreme situations of a load transfer platform largely extending beyond
the footing and limited to the edge of the footing. This study demonstrates that the limit stress
on the inclusion head varies rather linearly between the two values, associated respectively
with:
No overhang, i.e. load transfer platform limited to the edge of the footing: qp+(L = 0) = Nq*
JH;
An overhang of L, greater than Lmax, (load transfer platform extending well beyond the edge
of the footing) for which the limit value is that produced by the Prandtl mechanism qp+(P),
calculated with Equation 5.57 for a given load q0 and an area replacement ratio D.
Remark: The case studied herein is that of a load transfer platform with thickness
HM = 30 cm, an inclusion of diameter D = 30 cm, angle of friction I1 IRUWKH
platform and I2 IRU WKH VXUURXQGLQJ VRLO 1RWH WKDW KHUH WKH LQIOXHQFH RI
ratio HM/D has not been neglected).
The limit stress on the considered footing for a given overhang L can thus be estimated using
a linear interpolation between the two extreme values qp+(L = 0) and qp+(P) (Fig. 5.67):
q p L
238
q p L
0
Lmax
>q
p
P q p L
(5.57)
Figure 5.67: Principle used to determine the threshold stress on the inclusion head by interpolating
between the extreme values for an overhang greater than Lmax and a zero overhang.
Justifications
239
1
2
q p ( P ) q p ( L )
3
3
(5.58)
q p ,e
2
1
q p ( P ) q p ( L )
3
3
(5.59)
Figure 5.68
(5.60)
(a)o corner
q p ,a
7
5
q p ( P ) q p ( L )
12
12
(5.61)
(e)o exterior
q p ,e
5
1
q p ( P ) q p ( L )
6
6
Figure 5.69
(5.62)
4.2.4. Summary
The limit stress value at the inclusion head for an inclusion placed at the center of a mesh may
be determined based on Equations 5.38 and 5.43, by adopting for the coefficient Nq the value
associated with the Prandtl mechanism (Table 5.3), along with a shape coefficient sq equal to
1. The angle of friction is the one found at the critical state (i.e. shear at constant volume).
It is safe to neglect the increase in Nq when the load transfer platform is of limited thickness
relative to the inclusion diameter.
In the case of peripheral inclusions, the limit stress at the top of the inclusion depends on the
relative position of the edge of the outermost inclusion to the edge of the footing (overhang
L). Results obtained confirm the validity of the assumption that this stress varies linearly with
L over the interval [L = 0, Lmax] (Lmax: width of the Prandtl mechanism, see Fig. 5.58 and
Table 5.3) between these two extreme values:
For an overhang L greater than Lmax, i.e. the value qp+(P) associated with the Prandtl
mechanism as calculated by Equation V.56 in the configuration of a basic unit cell subjected
to a uniform load q0. The coefficient Nq only depends on the load transfer platform angle of
friction (Table 5.5);
No overhang, i.e. the value qp+(L = 0) = Nq* JH; this value depends on the vertical stress JH
applied outside the footing at the level of the inclusion head as well as on the coefficient Nq*,
calculated using the angles of friction of both the load transfer platform and surrounding soil
(Table 5.3).
240
The value qp+ (L) is obtained by linear interpolation between these two values
(Equation 5.57).
To determine the limit values for different inclusions at different locations under the footing,
different combinations of the weighted values of qp+(L) and qp+(P) are used depending on the
actual location of the inclusion with respect to the footing edge.
x Application example:
Load transfer platform thickness HM = 0.3 m, and angle of friction I1
Inclusion diameter D = 0.4 m
Area replacement ratio in the central mesh: D = S 0.22/1.22 = 8.7%
Surrounding soil: angle of friction I2 DQGXQLWZHLJKWJ = 18 kN/m3
Influence of the ratio HM/D neglected
Design load on the footing Qd = 2.2 MN (representing an equivalent uniform stress q0 ~ 300
kPa).
I1
Nq (P) = 48.9
(Table 5.3)
Lmax/B = 7.0
(Table.3)
Lmax = 7.0 u 0.4 = 2.8 m
The value qp+(P) applicable for an overhang of L greater than 2.8 m:
Nq = Nq(P) = 48.9
D = 0.087
q0 + JMHM ~ 300 kPa
(platform weight JMHM neglected)
qp+(P) = 2,839 kPa
(Equation 5.56)
qs+(P) = 2,839/48.9 = 58 kPa
The value qp+(L = 0) for a zero overhang
Nq*(I1I2
(Table 5.5)
+
qp (L = 0) =Nq* JH = 13.05 u 0.8 u 18 = 188 kPa
yielding:
qp+(L) = 188 + 0.3/2.8 (2,839 188) = 472 kPa
and the following limit values at the inclusion head for:
an interior inclusion
qp,i+ = 2,839 kPa
a corner inclusion
qp,a+ = 7/12 u 2,839 + 5/12 u 472 = 1,850 kPa
Justifications
241
Note:
The previous calculations were performed assuming that:
The stress qs+(P) obtained in the interior grid remains less than the calculated value of the
limit stress at the soil surface Vv,d;
The maximum load in inclusion Qp,max satisfies the strength criteria relative to both the soil
and native inclusion material.
Remark: In the case of a thin load transfer platform located under a rigid structure
(such as a footing, slab on grade or raft), at the edge of this structure and under
certain geometric conditions (distance from the inclusion compared with the
distance to the edge of the structure, platform thickness, etc.), two failure modes
may coexist: one described with the Prandtl mechanism, and the other for a plastic
flow starting from the inclusion head extending at the surface to the edge of the
structural element. The construction specifications adopted for purposes of this
document typically enable avoiding the generation of plastic flows up to the ground
surface; as such, no special verification beyond the one presented above is required
herein.
Figure 5.71: Development of failure mechanisms according to the extent of foundation overhang.
242
In light of the current state-of-the-art, it is proposed herein to adopt the same set of
recommendations. Close attention must be paid to adopting a failure criterion compatible with
this tensile strength.
References
BS8006 Code of Practice for Strengthened/reinforced soils and other fills. Section 8, Design of
embankments with reinforced soil foundations on poor ground, 2010.
Btons de sable - &DUDFWpULVWLTXHV HW SUDWLTXHV GXWLOLVDWLRQ UpDOLVp VRXV OpJLGe du projet national
Sablocrete. Presses des Ponts, 1994.
Berthelot P., Durand F., Glandy M., Frossard A Dallage et modules de dformation des couches de
sols ; applications aux renforcements de sols par inclusions et analyse du comportement du matelas
de rpartition . XIV ECSMGE 2007, Madrid, Spain.
Brianon L. 5DSSRUWILQDOGHOH[SpULPHQWDWLRQGH&KHOOHV . Rapport 2-08-1-05, juillet 2008.
Caquot A., Kerisel J. - Trait de mcanique des sols, Gautier-Villars, 1966.
Chevalier B. tudes exprimentale et numrique des transferts de charge dans les matriaux
granulaires. Application aux renforcements de sols par inclusions rigides. Thse de doctorat,
universit Grenoble I Joseph-Fourier, soutenue le 5 septembre 2008.
CUR Design guideline for piled embankments. CUR 226, 2010.
EBGEO Empfehlung fr den Entwurf und die Berechnung von Erdkrper mit Bewehrungen als
Geokunststoffen. In: Bewehrte Erdkrper auf punkt- oder linienfrmigen Traggliedern. Deutsche
Gesellschaft fr Geotechnik e.V., (German Geotechnical Society), Ernst & Sohn, (Kapitel 9), 2010.
Glandy M., Frossard A. -XVWLILFDWLRQ GXQH IRQGDWLRQ VXSHUILFLHOOH VXU XQ VRO UHQIRUFp
GLQFOXVLRQV . $QQDOHVGHO,%73, n 1, 2002, p. 45-53.
Hewlett W.J., Randolph M.A. Analysis of piled embankments. Ground Engineering, April 1988,
p. 12-18.
ISO/TR 20432 Lignes directrices pour la dtermination de la rsistance long terme des
gosynthtiques pour le renforcement des sols, 2007.
Jenck O. Le renforcement des sols compressibles par inclusions rigides verticales. Modlisation
physique et numrique. Thse de Doctorat, INSA Lyon, 29 novembre 2005.
Lawson C.R. Serviceability limits for low-height reinforced piled embankments. Proceedings
GeoEng 2000, Melbourne, Australia. Lancaster: Technomic Publishing Co.
NF EN 1997-1, Eurocode 7 Calcul gotechnique. Partie 1 : Rgles gnrales. (indice de classement
P 94-251-1) avec son annexe nationale (indice de classement P 94-251-2).
NF P 11-213, DTU 13.3 - Dallages - Conception, calcul et excution, 2007.
NF P 94-262 Justification des ouvrages gotechniques - 1RUPHV GDSSOLFDWLRQ QDWLRQDOH GH
O(XURFRGH- Fondations profondes.
NF P 94-270 Calculs gotechniques Ouvrages de soutnement, Remblais renforcs et massifs en sol
clou, 2009.
Nordic handbook Guidelines for reinforced soils and fills, Nordic geosynthetic group, rev. B,
october 2005.
Okyay U.S. tude exprimentale et numrique des transferts de charge dans un massif renforc par
inclusions rigides. Application des cas de chargements statiques et dynamiques. Thse de
doctorat, INSA Lyon, 24 novembre 2010.
Pecker A. Capacity design principles for shallow foundations in seismic areas. Keynote lecture, XIth
European conference on earthquake engineering, Paris, 1998.
Salenon J. Calcul la rupture et analyse limite. Presses des Ponts, 1983.
Schlosser F. Analogies et diffrences dans le comportement et le calcul des ouvrages de
soutnement en terre arme et par clouage des sols . $QQDOHVGHO,7%73 n 418, 1983.
SETRA Traitement des sols la chaux et/ou aux liants hydrauliques (GTS). Application la
ralisation des remblais et des couches de forme. Guide technique, 2000.
SETRA-LCPC Ralisation des remblais et des couches de forme (GTR). Guide technique,
1992 (rvis 2000).
Justifications
Summary
243
CHAPTER 6
Geotechnical investigations
1. GEOTECHNICAL ENGINEERING MISSIONS DEDICATED TO REINFORCING
COMPRESSIBLE SOILS BY MEANS OF RIGID INCLUSIONS
This document serves to define, with reference to the NF P94-500 Standard on geotechnical
missions, the set of preliminary geotechnical designs (G1) and detailed designs (G2), with the
objective of reinforcing compressible soils by means of rigid inclusions. It details the scope of
geotechnical investigations required during the various study phases.
Note: Throughout this chapter, the notations used to represent geotechnical
parameters will be defined in the table included in the Appendix to the
Recommendations section.
1.1. Geotechnical site feasibility study (G11)
This study is solely intended to broadly define the geological and geotechnical characteristics
of the project setting, typically without necessitating more in-depth geotechnical surveys. In
general, such a study only includes an analysis of pertinent documentation, a non-exhaustive
list of which is provided below:
Geological study;
Hydrogeological study;
Geotechnical databases: geotechnical consulting engineers, Infoterre, archives, publications,
etc.;
Site inspection, survey of soils and neighboring zones;
Natural Risk Prevention Plan (French acronym: PPR);
Documentary evaluations;
Environmental constraints.
The report generated from this initial study defines a draft preliminary geological and
geotechnical model that describes foundation soil compressibility. This model must expose
the various risks inherent in the predicted settlements of structures in terms of execution
difficulty, costs and scheduling; it must also propose a range of feasible technical solutions.
1.2. Preliminary geotechnical design (G12)
The purpose of this preliminary geotechnical design is to provide the basis for a preliminary
design of the structure by adopting the set of geotechnical hypotheses for consideration and
identifying the general construction guidelines. In relying on the outcome of a technicaleconomic study, this stage provides a choice of solutions that address the constraints imposed
by the Project Owner and Engineer, e.g. maximum settlements, differential settlements,
construction schedule. Upon completion of this G12 study, the Project Owner and Architect
select the approach that best satisfies their technical and economic constraints.
Geotechnical investigations
245
This preliminary geotechnical design must also comprise the geotechnical surveys needed to
specify the geotechnical data described in detail in Chapter 4; these investigations may
include:
Geophysical surveys;
Geological boreholes with sample extraction;
In-laboratory identification testing;
Laboratory tests to determine mechanical parameter values;
In situ testing campaigns;
Hydrogeological investigations.
The report produced on this preliminary design is to establish the definitive
geological/hydrogeological/geotechnical model; moreover, it provides a series of dimensional
sketches of the various solutions, with an emphasis on settlements. In some cases, this report
may reach the conclusion that the project structure does not require any improvements or
reinforcement, notably should the estimated site settlements remain compatible with the
structures stability and serviceability.
This G12 report is to set forth a schedule of additional surveys, depending on the potential
feasible solutions, to be conducted during the detailed geotechnical design phase (G2).
1.3. Detailed geotechnical design (G2) specific to a rigid inclusion-based reinforcement
project
From a general perspective, this detailed design offers an approximation of quantities /
schedules/execution costs for the projects geotechnical structures, and a list of the
consequences of residual geological risks incurred relative to each of the various feasible
techniques, as input to the Project Owners decision-making process.
Keep in mind that the contents of this design, as described in the present section, are specific
to reinforcement projects involving rigid inclusions.
During the Design phase, this study must serve as the basis, depending on constraints
imposed by the Project Owner and Engineer, for defining:
Type of inclusions;
Inclusion mesh pattern;
Inclusion length;
Thickness and implementation criteria relative to the transfer platform.
Within this framework, a campaign of additional geotechnical investigations is planned for
the purpose of determining the values of all parameters input into the design and
dimensioning calculations for the reinforcement solution by rigid inclusions; this campaign
features:
Cone penetration tests, piezocone tests to complete the lithographic sections and determine
the over-consolidated zones;
Shear characteristics: cuFMGHULYHGIURPWULD[LDOWHVWLQJ
Determination of cu using a vane test;
Compressibility characteristics: Vp, ROC, e0, Cc and Cs, Cv, kv under various stress levels,
ck, CDe;
Testing: installation of piezometers and subsequent monitoring, in order to identify the
amplitude of water table fluctuations over time.
246
Geotechnical investigations
247
248
Lithostratigraphic
sections of soft
layers
Geological
model
Objectives
Linear structures:
see Note 1 below.
x + one additional
sounding per
increment of
2.500 m2
Buildings:
x Fora surface area <
2.500 m2: three
soundings
Mesh pattern
Destructive soundings:
x pit
x continuous auger
x destructive boreholes with recording of
boring parameters
Geological soundings:
Cored boreholes (see Note 2) well suited to
soft soils, basis for soil stratigraphy and
identification
Geophysics:
electrical array or vertical electrical sounding
Seismic refraction
Type
Depth
Note 2: All geotechnical investigations must contain at least one reference cored borehole. Based on this reference, the surveying campaign may be complemented by
examining destructive soundings. Cone penetration tests will make it possible to identify the various soils by calibration on the cored borehole; in the case where the
penetrometer is equipped with a friction coupling (CPT), as well as a pore pressure measurement device (piezocone or CPTU), the series of abacuses shown in Figures VI.3
through VI.5 offer valuable assistance.
Note 1: The survey campaign may be organized by dividing the targeted alignment or zone into basic meshes containing the same soundings and tests. Figures VI.1 and VI.2
depict an example of cutting an alignment into meshes with length 4L, where the value of L should lie in a range between 50 and 250 m, depending on the site and soil
variability. For a linear structure, the boreholes must be located on the alignment axis as well as on two lines running some 30 or 50 meters on both sides of the axis. The
density of survey data points relates to soil variability and needs to be adapted to satisfy project consistency.
Project
layout plan
Projectinfor
mation
Figure 6.3: Soil identification abacus according to the cone penetration test (CPT).
(Douglas and Olsen, 1981).
Geotechnical investigations
249
Figure 6.4: Soil identification abacus according to both the CPT and CPTU tests
(Robertson et al., 1986).
Figure 6.5: Soil identification abacus according to the standardized CPT and CPTU parameters
(Douglas and Olsen, 1981).
250
Geotechnical investigations
251
Type
x Closed piezometers
(CPI)
x The piezocone is
able to provide an
estimation of
horizontal kh.
x During oedometric
testing, it is possible to
measure the vertical
permeability kv.
Note 1: The determination over time of water table raising and lowering is important for its contribution to soil over-consolidation.
Project information
Hydrogeology
Objectives
Mesh pattern
1
piezometer
per every
x Low, high and exceptional water table
2
3,000-m surface area
levels: EB, EH, EE
x Historical record of water table variations increment, in order to
more closely match the
(see Note 1)
historical study
x Prediction of water table variations
Permeability
x Water analyses, if necessary (corrosion,
aggressiveness / concrete binder)
Depth
Water table and
deeper
groundwater
aquifers, if
necessary.
252
Project constraints:
x Maximum allowable
settlement
x Maximum allowable
differential settlements
x Residual settlements
following acceptance of
the structure
Project information
x Service loads
x Embankment heights
x Construction timeline
x Service startup
schedule
Objectives
x Identifications of soft
layers
Classification: NFP11300; GTR 92
x Rate of excess pore
pressure dissipation
x Precise knowledge of
the soft thicknesses
x Failure parameters
x Compressibility
parameters
Fora surface area of
between 500 m2and 10,000
m2 :
one sounding per every
500 m2, with a minimum
of three soundings
minimum distance
between soundings
= 40 m.
Geotechnical
Mesh pattern
For a surface area <
500 m2: two soundings
Shear tests:
Triaxial tests UU, CD or CU+u,
with the exception of the box
shear apparatus
MDQGFNote 12), cu (Notes 13
and 14), and Ocu (Note 15).
Type
In situ tests (see Note 1)
Cone penetration and piezocones
vane tests
Pressure meters
Depth indication
x For vane tests (every
0.50 m):over the entire
compressible layer
thickness
x For pressuremeters,
CPT and cored
boreholes: 7 diameters
and 3-m minimum
below the inclusion
base, or at the stop for
the CPT.
Note 1: The cone penetration and pressuremeter tests will serve to justify the load-bearing
capacities of rigid inclusions. On the other hand, these tests cannot reliably estimate either the
consolidation settlement or creep settlement of soft soils, and even less reliably the settlement
durations. Use of a pressure meter may prove acceptable for relatively incompressible
unsaturated soils, whose settlement periods may be considered of short duration.
Note 2: The first identification testing campaign objective is to classify and then qualify the
soils (highly plastic clay, peat, etc.). The second objective consists of correlating these
identification parameters with both the shear and compressibility parameters, in order to
complete the project database (wn, with Cc/(1 + e0), MZLWK,p).
Note 3: Benchmark values for the physical parameters of compressible soils.
Table 6.4: Benchmark values of compressible soils.
Designation
Us
(Mg/m3)
Ud
(Mg/m3)
wn (%)
wL
Ip
Soft clays
2.6-2.7
0.9-1.8
1.2-2
30-100
20-80
15-50
Muck
2.4-2.7
0.7-1.5
1.5-3
60-150
80-180
Peat
1.4-2
0.1-0.5
3-10
200-1,000
MO (%)
2-10
10-100
Note 4: Oedometric testing is critical for estimating primary and secondary (creep)
settlements as well as their predicted durations.
Note 5: Preconsolidation pressure
The determination of preconsolidation pressure V'p is an essential, yet very delicate, step. An
indication of the degree of over-consolidation (ROC = Vp /Vv0) may be obtained from
soundings using the cone penetration and piezocone tests (Fig. 6.6).
For a normally consolidated clay, the theoretical variation of qtvs. depth can be plotted, with
adequate knowledge of the relation between cu and Vv0 or between cu and qt (see Note 13),
and then compared to the qt values recorded by the cone penetration or piezocone. This
comparison of the two values displays whether the soil is over-consolidated or instead nearer
the normally consolidated state. Figure 6.6 provides an example for a homogeneous clay,
which is distinctly over-consolidated at the top and then very slightly over-consolidated
throughout its depth.
Geotechnical investigations
253
Note 6: Compressibility
'e
on the unloading-reloading cycle
' lg V V'
'e
CC
Compression index:
on the initial loading
' lg V V'
Oedometric tangent modulus for a stress Vv, in the normally consolidated domain, vs. Cc (or
in the over-consolidated domain vs. Cs):
2. 3 V v '
1 e
Eoed V v'
Cc or C s
Oedometric secant modulus, Eoed, over a stress interval Vv1, Vv2,: quotient of the variation in
Vv divided by the volumetric variation over this same interval:
'V v '
1 e1
Eoed V v' 1 ,V v' 2
'e
Recompression index:
CS
Site
Clays with low or average
sensitivity
Marine Rio de Janeiro clay
Correlation
Reference
Cc = 0.009 (wL-10)
Cc = 0.013 (wL-18)
Ortigao (1975)
x Between the compressibility factor Cc/(1 + e0) and natural water content wn (Fig. 6.7):
254
Figure 6.7: Range of Cc/(1 + e0 ) values vs. wn for normally consolidated, non-organic clays
(Lambe and Whitman 1969, western U.S. soils).
Figure 6.8: Correlations for Bordeaux clays- Cc and Cs = f(wn) (A. Marache et al., 2009).
Note 8: Benchmark values for compression index Cc and compressibility factor Cc/(1 + e0).
Table 6.6: Benchmark values for the compressibility parameters of compressible soils.
Type of soil
e0
Cc
Cc/(1 + e0)
Soft clays
1.2-2
0.3-1
0.15-0.3
Muck
1.5-3
0.7-1.6
0.25-0.4
Peat
3-10
2-10
0.4-0.8
Geotechnical investigations
255
kv Eoed
Jw
This relation can also be written as a function of Cc within the normally consolidated domain
and, by extension, within the over-consolidated domain vs. Cs for a stress V DV FORVH DV
possible to the stress imposed on the soil by the structure to be built:
in the over-consolidated domain:VVp
k 1 e
cv
2.3 v
V'
Cs J w
in the normally consolidated domain:V!Vp
k 1 e
cv
2.3 v
V'
Cc J w
cv
256
'H
' lg t
'H
H0
' lg t
CDe
'e
' lg t
CD 1 eo
Type of soil
Soft clays
Muck
Peat
CDe/Cc
0.03-0.05
0.03-0.05
0.05-0.10
Note 12
Let us recall herein the principles of triaxial testing, as described in the NF P 94-070 and NF
P 94-074 Standards issued in October 1994.
For soils sheared in the normally consolidated domain F | 0; the angle of friction will be
derived at the critical state MNC = Mcrit (on the load face) and will remain indistinguishable
from the angle of friction at the peak (Fig. 6.10).
For soils sheared in the over-consolidated domain, if the deformation is limited and less than
that of the yield stress (beneath the yield surface), then the values will be set at the peak
(Fig. 6.10), resulting in: MSC = Mpeak DQGFSC Fpeak. In the over-consolidated domain, stress
paths all lead onto the ellipse. The path ends are therefore not aligned; however, the curve
will generally be linearized in order to calculate Mpeak DQG Fpeak. It can be observed in
Figure 6.10 that the determination of Mpeak DQG Fpeak requires considerable skill and is less
reliable than the calculation performed in the normally consolidated domain.
Figure 6.10: Example of the determination of M DQGFLQERWKWKHQRUPDOO\FRQVROLGDWHGDQG overconsolidated domains for a Romainville green plastic clay (Josseaume and Azizi, 1991).
Geotechnical investigations
257
Figure 6.11: Curves representing the deviator and excess pore pressure vs. axial strain during
undrained tests conducted on normally consolidated and over-consolidated saturated clays.
For a given clay, the angle of friction Mdrops with the plasticity index value. Several
correlations can be identified, among which let us cite Fahris (1970) corresponding to French
clays:
8
tan M ' 0.21
Ip 6
Note 13: Correlations relative to the undrained cohesion cu
x Based on the standardized strength measured using the cone penetration test, qc or qt and cu
(Eurocode EC7-2):
Cone penetration
cu = (qc v0)/Nk
Nk must be determined at the project site.
Generally, the value of Nk is set at: 15 < Nk< 17.
Piezocone
cu = (qt v0)/Nkt
Nkt must once again be determined at the project site.
Its value typically lies between 15 < Nkt< 20.
x Based on the plasticity index, Ip:
For normally consolidated clays: on any given site, Skempton proposed (Fig. 6.12):
cu = (0.11 + 0.0037 Ip) Vv0
258
Note 14: Classifications and benchmark values of cu, qc, EM and pl.
x According to Terzaghi (Table 6.8).
Table 6.8: Values of cu vs. clay consistency.
Clay consistency
Very soft
Soft
Firm
Stiff
cu (kPa))
< 12
12-25
25-50
50-100
Type of soil
EM (MPa)
Muck
Soft clay
0.2-1.5
0.5-3
pl (kPa)
(using a probe with very low
inertia)
20-150
50-300
Geotechnical investigations
259
Figure 6.13:
Determination of Ocu in the Lambe diagram.
Figure 6.14:
Determination of Ocu in the Mohr diagram.
For very recent soils, e.g. marine clay, for which it can generally be considered that the entire
layer is normally consolidated, short-term cohesion cu rises linearly with depth beginning
from zero at the surface (Fig. 6.15).
For most soils, we obtain: cu = 0.25 to 0.35 Vv0 for clays, and cu = 0.5 Vv0 for peat.
For over-consolidated soils at the surface (e.g. subsequent to water table level variations that
over-consolidate, by means of suction, the soil slice through which the water table is
fluctuating during drying-rewetting cycles), it can be concluded that cohesion exhibits a nearconstant value in the over-consolidated soil thickness, before increasing proportionally with z
(Fig. 6.15).
260
References
Eurocode 7-1 - Calcul gotechnique. Partie 1, Rgles gnrales, juin 2005.
Eurocode 7-2 - Calcul gotechnique. Partie 2, Reconnaissance des terrains et essais.
septembre 2007.
Leroueil S., Magnan, J.-P., Tavenas F. - Remblais sur argiles molles. Lavoisier, 1985.
Lunne T., Robertson P.K., Powell J.M. - Cone penetration testing in geotechnical practice.
Blackie academic and professional, 1997.
Marache A. et al. - Understanding subsurface geological and geotechnical complexity at
various scales in urban soils using a 3D model. Universit de Bordeaux, GHYMAC.
Georisk 2009.
NF P 11-213-1 - DTU Dallages 13.3 : Conception, calcul et excution, mars 2005.
NF P 11-300 (ou GTR 1992) - Guides techniques, fascicules 1 et 2. Ralisation des remblais
et couches de forme. SETRA, LCPC, septembre 1992.
NF P 94-500 -0LVVLRQVGLQJpQLHULHJpRWHFKQLTXH&ODVVLILFDWLRQHWVSpFLILFDWLRQV, dcembre
2006.
Recommandations sur la consistance des investigations gotechniques pour la construction
GHEkWLPHQWVGHOUSG . Le Moniteur (cahier dtach n 2 du 16-12-2005, n 5325).
SETRA-LCPC - tudes et ralisation des remblais sur sols compressibles. Guide technique,
novembre 2000.
SETRA-LCPC - Traitement des sols la chaux et/ou aux liants hydrauliques, application la
ralisation des remblais et des couches de forme. Guide technique, janvier 2000.
Geotechnical investigations
Summary
261
CHAPTER 7
Execution conditions
1. COORDINATION WITH GEOTECHNICAL ENGINEERING MISSIONS
To ensure high structural quality, the following steps are necessary and pertain to all elements
of the rigid inclusion system: embankments, inclusions, capping layers or platforms, rebar,
excavation pits, earthworks, foundations:
Definition by the Prime Contractor of the various target objectives: maximum settlements,
loading intensity, schedule restrictions, ground slab thickness, etc.;
Completion by a certified Engineer of the geotechnical project design (type G2);
Execution of the works by a Contractor assigned oversight of the execution plan and
geotechnical monitoring (type G3);
Completion by a qualified Structural Engineer of the geotechnical construction supervision
task (type G4), for the purpose of managing all external controls, especially technical
interfaces among the various works involved in the project;
Assignment, through an organization certified by the competent administrative authority, of
a technical control mission intended to provide risk prevention assistance.
Prior to the involvement of the Contractor, either the Project Owner or Engineer will have
completed all of the preliminary information requests regarding concessionary agreements.
Moreover, the soil reinforcements, like any other technique related to geotechnical works,
require supervising the works execution, in addition to monitoring and maintenance steps in
accordance with the prescriptions set forth in Eurocode 7, Section 4.
The purpose of this chapter is to define the execution steps and procedures relative to:
The working platform;
Rigid inclusions (inclusion layout, production, low cut-off);
Pile caps or head extensions;
Load transfer platforms (type, thickness, implementation);
Eventual geosynthetic reinforcement included in the platforms.
Lets point out that these various elements may in fact be installed by different contractors, a
situation that necessitates excellent interface management.
It should also be recalled that a series of execution designs are mandatory, in order to:
Validate appropriateness of the technique given the type of soil and working environment;
Verify the means and methods deployed in light of the targeted results, in terms of
performance, output rate and schedule;
Refine the design of the rigid inclusion solution selected during the project's geotechnical
study phase, by taking into account the full set of execution parameters, including geometric
and mechanical characteristics of the rigid inclusion, specific points implying adaptation of
the inclusion grid pattern, characteristics of the load transfer platform (this study process must
be conducted according to the principles outlined in Chapter 5);
Develop a layout showing the location of inclusions based on the set of technical data
validated by the Engineer;
Execution conditions
263
Adjust the works phasing plans according to the program of the other contractors;
Propose technical solutions that allow managing environmental constraints, geotechnical
hazards and execution-related risks (e.g. discovery during the execution phase of underground
or buried structures not detected by previous investigation surveys);
Present a method statement submittal for rigid inclusions, with their location, the different
stages of the process and a quality control plan.
2. CONSTRUCTION OF THE WORKING PLATFORM
2.1. Role of the working platform
A working platform is typically constructed before the rigid inclusion works. This platform
allows the transport of the rigid inclusion installation equipment during all weather
conditions.
The characteristics of such a platform must be adapted to actual execution conditions (type of
equipment, ease of circulation on the site, etc.). Moreover, platform execution may give rise
to an acceptance that guarantees the safety of onsite construction vehicles.
This working platform can be part of the load transfer platform but must be controlled by
specific tests described in this Chapter. In this case, the working platform must undergo an
acceptance procedure to verify the achievement of the specified performances.
Furthermore, upon completion of the rigid inclusions, this working platform is typically
affected by subsequent jobsite operations, which may undermine the integrity of previously
installed rigid inclusions (whether reinforced or not).
2.2. Works after rigid inclusion execution
Both during and after soil reinforcement measures, precautions need to be taken in order to
avoid damage to the inclusions (which in general have not been steel reinforced) as well as of
the capping layer.
The risks of damage to the upper part of a rigid inclusion may appear in the following cases:
Circulation of construction vehicles either directly above or in the vicinity of rigid inclusion
heads;
Insufficient platform bearing capacity with regards to the circulation of construction
vehicles;
Installation of utility lines (water, gas, electricity, etc.) either between or in alignment with
the rigid inclusion mesh;
Earthworks involved in excavating the footing pits;
Restoration of the working platform subsequent to soil reinforcement works in order to
proceed with localized material purging;
Restoration of the working platform by running binder treatment equipment (using lime,
cement);
Second compaction of the working platform and installation of an additional capping layer
extending to the underside of the structure.
264
265
Rigid inclusions installed by incorporating a binder with the soil (treated soil columns, jet
grouting, etc.).
Moreover, rigid inclusions may be distinguished by their component material: mortar,
concrete, grout, steel, or even wood.
From a general standpoint, all pile and micropile execution methods would be applicable for
installing rigid inclusions (see Appendix A: Standard NF P 94-262).
The inclusions that mix a binder with the soil constitute a special subcategory of inclusions
produced in situ, particularly as regards control procedures.
For a detailed description of each of these implementation techniques, the interested reader is
referred to:
Standard NF EN 1536 for bored piles without soil displacement;
Standard NF EN 12699 for piles with soil displacement;
Standard NF EN 14199 for micropiles;
Appendix A of Standard NF P 94-262 for pile and micropile-related techniques;
Standard NF EN 12 716 for techniques on soil-cement columns derived by jet grouting;
Standard NF EN 14 679 for the execution of treated soil columns.
3.2. Material-related requirements
The rigid inclusion component material exhibits a much higher deformation modulus than that
of any soil layers and a specific strength independent of the level of confinement capable of
being provided by the surrounding soil. In addition to the mechanical characteristics taken
into consideration in the design calculations, requirements related to the following must also
be verified:
Execution: resistance to leaching, workability of these materials;
Durability: corrosion, attacks from chemical and bacteriological agents.
3.3. Description of the execution methods
Among the aspects or criteria playing a role in the choice of execution method, let's cite the
following (non-exhaustive list):
Borehole stability;
Risk of false-refusal;
Stability of a column with fresh material;
Creation or not of excavations;
Vibrations and nuisances;
Possibility of reinforcing the inclusion;
Recording of execution parameters;
Rate of output;
Maximum lengths and diameters;
Strength of the embedment layer;
Aggressiveness of the soil and water;
Site accessibility of the inclusion execution equipment;
Local means;
Phasing constraints related to the method employed (scheduling plan).
266
267
268
Vibrations
Noise
Excavation
wood
yes
yes
no
metal (steel)
yes
yes
no
concrete
yes
yes
no
(2)
Drilled by
yes
yes
no
percussion
Vibro-driven
yes
no
shallow
Mortar: 5,000 - 10,000
Simple bored
no
no
yes
Driven,
Concrete C15: 9,000
no
no
yes
hammered Drilled cased or
Concrete C25: 11,000
sludge drilled
Fabricated and bored
Bored with a
no
no
yes
in situ
continuous auger
Bored with
no
no
shallow
displacement
Treated
Soil mixing
no
no
shallow
Variable: 250 - 9,000 (3)
soil
Jet grouting
no
no
variable
500 - 1,000 * compressive strength
columns
1
depends on the type of concrete and reinforcement; 2depends on the type of grout; 3depends on the type of
binder and soil.
Execution conditions
269
Depth of the rigid inclusions or their embedment within a load-bearing layer, in compliance
with the project design computations (the stop criterion must be subsequently defined at the
time of calibrating the workshop when construction starts up);
Maximum working stress of the rigid inclusions and compressive strength (fc28) of the
material used in the case of a binder-containing material (concrete, mortar, grout, soil-binder
mix, etc.);
The inclusion mesh underneath the ground slab or embankment, and the thickness of both
the working platform and capping layer forming the load transfer platform;
The layout of inclusions underneath soil blocks and footings, and the layouts adopted at the
soil block-inclusion interface in the case of moments or horizontal forces;
The procedure of cutting the rigid inclusions underneath the soil blocks;
Pile cap dimensions, if applicable;
Characteristics and performance of the reinforcement layer, if applicable.
3.6. Means and methods
The Contractor describes within a technical execution procedure both the human resources
and equipment scheduled for implementation in order to achieve the predefined objectives:
Worksite supervision: composition of the worksite supervisory team, references of the site
foreman;
The execution team: composition of the team, which in general features 3 members for
small jobs (1 site foreman, 1 drill operator, 1 pump operator);
The execution method, as ultimately described in the previous section, in specifying the
type and volume of excavation material to be removed;
The number and type of production workshops, their power rating (torque for an auger,
frequency and amplitude for a vibro-hammer, etc.), the maximum working depth, providing
The possibility to justify the output rates relative to job scheduling;
The resources selected for manufacturing and transporting the component material;
Transport onto the worksite when introducing ready-mix materials;
Onsite production in the case of installing an in situ mixing unit;
Transport onto the worksite: concrete pump or loader when using dry concrete.
The control plan appended to the execution procedure summarizes, for each implementation
phase, the acceptance criteria, as well as the type, frequency, resources and manager of the
corresponding control measures.
3.7. Inclusion execution drawings
When the inclusion design studies and execution drawings lie outside the Engineers mission,
these elements are typically established by the Contractor, in conjunction with the execution
studies produced by other site actors and then validated by the Engineer and control bodies;
they indicate:
The layout of the inclusion grid pattern with dimension benchmarks;
Distinctive signs of the various inclusions: reinforced or unreinforced, inclusions beneath
the structure and beneath the ground slab;
The cross-section view beneath the ground slab revealing: the working platform, the
inclusion surface, details of the load transfer platform;
A sketch of the cross-section view underneath the soil blocks, revealing: the lower block
surface, and the load transfer platform dimensions.
270
Execution conditions
271
272
Remark: This criterion may be extended to the case of medium to stiff soils (e.g. a
total expected settlement on natural ground of less than 5% of the embankment
height).
The use of reinforcements (geosynthetics, steel fabric mat) acting in tension may serve to
increase this spacing or reduce the pile cap diameter, provided justification of the differential
settlements.
3.10.4. Number, minimum distances and installation offset of inclusions
3.10.4.1. Inclusions underneath a ground slab or raft
When placed under distributed loads, inclusions are installed at their theoretical axis. The
tolerated axis offset is less than or equal to 20 cm in distance and 2% in inclination.
Execution conditions
273
At the border of the structure, it is necessary to lay out a row of inclusions around the edge of
the ground slab if it has become separated from the peripheral grade beam. The axis of this
row must remain at a distance from the ground slab edge of no more than one-fourth of the
spacing.
3.10.4.2. Inclusions underneath footings
3.10.4.2.1. General principles
As a general rule, it is imposed that inclusions are laid out on at least two parallel axes, i.e.:
Along the x-axis once Mx/Q > B/6;
Along the y-axis once My/Q > L/6.
The minimum distance from the bare part of the inclusion (or pile cap, if present) to the edge
of the footing equals 15 cm, when taking into account the execution tolerance.
It is to be recalled that the design presented in Chapter 5, Section 4.2.3, entitled Calculation
of the transfer limit on inclusions heads at the edge of the structure, accounts for the distance
between the bare part of the inclusion and the edge of the footing.
In the case of soils with especially low shear strength, in general it would be prudent to set an
upper overhang at the minimum value of 15 cm, in order to ensure better platform
confinement around the inclusions heads and thereby enhance inclusion efficiency.
The execution tolerance in all directions must not exceed 10 cm.
The inclination tolerance equals 2% in all cases.
3.10.4.2.2. Inclusions under isolated footings
The number of inclusions under a given footing depends on both the allowable stress and
settlement.
Remark: Within a given structure, it is entirely realistic to find some footings with
reinforced underlying soil while others resting on natural soil have been slightly
loaded. For such a situation to be authorized, settlements must remain
274
Execution conditions
275
upon completion of the design studies for a number of situations (longer inclusions with small
diameters and placed in soft soils) relative to the parasitic bending capable of being caused,
even under a vertical load.
4. PREPARATION OF THE INCLUSION HEAD
4.1. General principles
4.1.1. Cut-off of cement, mortar or grout inclusion
The cut-off operation refers to adjusting the level of the inclusion head. The heads of cement,
mortar or grout columns have to be set to an appropriate elevation while the material is fresh.
The inclusion head can stop on the surface of the working platform, but the elevation can also
be located below the working platform (i.e. low cut-off).
The leveling step may be performed in one of several ways, depending on an array of
parameters (e.g. depths, tolerances required, type of soils, reinforcement layer elevation,
inclusion diameter):
Either by halting the casting operation;
Or by manual or mechanical excavation (using valves) of inclusion material before setting;
Or by mechanical excavation of the soil and inclusion material before setting until reaching
the specified elevation (ditching bucket);
Or by drilling in the fresh mortar until the specified elevation.
Figure 7.4: Manual adjustment of the inclusion head elevation in fresh concrete.
If possible, a cut-off of the inclusion head at the same elevation as the platform surface is
preferred.
276
Remark: Stopping the casting operation or excavating below the water table is
strictly prohibited.
4.1.2. Cutting of inclusions
The cutting operation designates preparation of the inclusion head installed on the hardened
material so as to reach the intact material at the required level.
This cutting step must be carried out over at least 10 cm, in order to verify the quality of the
inclusion head material.
The earthworks contractor must exercise great vigilance when conducting works adjacent to
the inclusions, both before and after this cutting step.
Inclusion cutting typically entails:
A hand-held jackhammer;
Performing sawing;
The assistance of concrete spreaders;
Use of a chemical cutting system: after a wait time of at least 7 days to allow for setting,
this process makes it possible to excavate all at once and then remove the pieces of extra
inclusions without fearing a degradation of the inclusions already installed.
Use of hydraulic rock breakers is strictly forbidden.
It is not necessary to cut off the inclusion when the leveling technique employed yields highquality inclusion heads.
In the case of a low level that lies beneath the footing, the inclusion is systematically cut over
at least 10 cm as a means of verifying the quality of the material composing the head.
Figure 7.5: Chemical cut-off for the head inclusions below the raft level.
Execution conditions
277
278
Execution conditions
279
5.1. Reconstitution of the working platform in the event of a low inclusion elevation
5.1.1. Untreated granular pad incorporated into the platform
The working platform must be closed above the inclusion heads by a sandy or sandy gravel
and easily-compacted material, with a surplus volume. The shear strength characteristics of
this material must be equivalent to those of the granular platform. The platform is then once
again compacted to form the base of the load transfer platform.
5.1.2. Platform treated with hydraulic binders (cement or road binder)
Except when special measures are implemented, a working platform treated with hydraulic
binders cannot be easily reworked once the inclusions have been installed; in this case, it is
preferable to refrain from incorporation into the platform and instead cut-off the inclusion at
this platform's upper level.
5.1.3. Lime-treated platform
A working platform treated with lime proves easier to rework once the inclusions have been
executed (delayed material strength increase) and, hence, easier to integrate into the load
transfer platform should its characteristics and homogeneity meet requirements. In this case, it
is essential to ensure and verify the consistency of execution tolerances for the various
operations: inclusion leveling elevation and thickness of the reprocessing step.
5.2. Location of a platform under large-sized foundations (ground slab or raft)
These construction specifications are inspired from the specifications issued for implementing
the capping layer underneath the ground slab, as intended in the DTU Technical Guideline
13.3.
It must be ensured that the platform has been well compacted and remains free from pollution
(cuttings, equipment traffic on the platform) over its entire height, especially at the base
where it is exposed to the heaviest loads above the inclusion heads.
5.3. Implementation of a platform underneath a small-sized foundation
Placement of a platform beneath a small-sized foundation is only possible provided the
excavation pit walls are stable or retained and moreover at the time of its installation, the
water table is below the earthworks base.
For very deep excavation pits, special measures need to be designed to guarantee pit stability
and human safety.
This platform may be built before or after installation of the inclusions.
In all cases, it would be necessary for the implementation methods to guarantee platform
compaction quality over its entire thickness.
The choice of a treated soil platform might, under certain conditions, lead to overcoming the
set of compaction difficulties.
280
The effective management of interfacing among the various project actors requires that all
execution constraints be fully understood by the Engineer as of the design stage.
Under certain conditions, it could be recommended to conduct a preliminary full-scale test.
5.3.1. Platform installation after inclusion execution
The inclusions may be installed from a high elevation platform with a low cut-off at the
theoretical platform base. This installation consists of excavating down to the inclusion head,
then installing and compacting the load transfer layer, before casting the footing (Fig. 7.9).
Execution conditions
281
It sometimes proves necessary to excavate below the inclusion leveling elevation (e.g. as
required by the design or when the low soil bearing capacity limits the densification of the
load transfer layer (see Fig. 7.12)). In this case, it is necessary for the inclusions to be steel
reinforced over a sufficient height, extending a minimum of 4 times the maximum inclusion
diameter or exposed height.
Figure 7.12: Need for steel reinforcement of the upper part of the inclusion to allow for earthworks
below the inclusion heads.
282
283
Tests
Static loading Plate
Dyna-plate
Dynamic
Load-bearing
loading
capacity meter
Standards
NF P 94-117.1
NF P 94-117.2
Static loading
NF P 94-117.3
Deflection
under a 13ton axle
Westergaard
Benkelman
Beam or
deflectograph
NF P 98-200
Values
Reference
EV2 03D
GTR Guide
Kw 03DPHWHU
d 0 mm
(granular subgrade
layer)
d PP
(treated subgrade
layer)
DTU 13.3
Catalogue of
pavement
structures
284
Remark: LA: Los Angeles test, Standard NF EN 1097 2; MD: Micro-Deval test,
Standard NF EN 1097 1
Compacted to at least 95% of the OMP (Optimum density of Modified Proctor compaction
test).
In terms of compaction, a minimum of 95% OMP is recommended in all cases.
5.4.2. Treated materials
The strength of treated materials must be compatible with the loading transmitted in the
platform.
An assessment of the treatment ability must be conducted according to Standard NF P 94 100.
Remark: The cohesion durability assumed in the design calculations must be
guaranteed throughout the structural life cycle.
5.5. Minimum thickness
The minimum thickness is set equal to 40 cm under a raft or ground slab, and at 30 cm under
a footing. This thickness must be raised should it fail to meet the compactness criteria sought
at the layer bottom.
5.6. Platform overhang in the presence of footings
A minimum platform overhang relative to the space occupied by the footing is necessary in
order to ensure platform compaction quality over its entire thickness underneath the footing.
This overhang is greater than or equal to half the platform thickness, with a 30-cm minimum.
5.7. Eventual incorporation of a reinforcement layer
5.7.1. General remarks
The load transfer platform may be reinforced by installing one or more geosynthetic layers or
a steel fabric mat.
The reinforcement layers and their positioning must be chosen on the basis of the type of
product, the component materials (i.e. type of polymers for geosynthetics), their mechanical
or design characteristics, their evolution over time and the physicochemical aggressiveness of
the contact soil or water.
The set of specifications relative to the reinforcement layers are the same as those listed in
Standard NF P 94 270: Retaining structures, reinforced embankments and nailed soil
blocks, as complemented by the following considerations:
The specifications focusing on geosynthetics and the interaction with embankments must be
indicated in accordance with Standard NF EN 13251: Geotextiles and geotextile-related
products-Characteristics required for use in earthworks, foundations and retaining structures;
Execution conditions
285
Within the scope of the EC Marking campaign, and in compliance with Standard NF EN
13251, the adopted characteristics must be expressed by the material producer in the form of
an average value and the corresponding tolerance value(s) at the 95% confidence level.
Durability-related information must be provided in accordance with the directives set forth in
Appendix B (normative) of Standard NF EN 13251.
5.7.1.1. Types of geosynthetics
The geosynthetics applied in these contexts may be:
Reinforcement geotextiles, either woven or knitted, ideally placed at the base of the load
transfer platform or embankment;
Or reinforcement geogrids, preferably placed in the thickness of the load transfer platform
granular layer.
A reinforcement geogrid might also be installed on the load transfer platform base, at the
interface with the supporting soil, yet it would be necessary to combine the eventual geogrid
with a separation geotextile in order to avoid contaminating the granular layer by the finegrained soil.
5.7.1.2. Component material
The component polymer must be compatible in durability terms with the embankment
material. The polymers used to produce embankment reinforcements are mainly polyester (or
polyethylene terephthalate, PET), polypropylene (PP), along with other polymers like
polyaramide (pAR) and polyvinyl alcohol (PVA).
These polymeric components are chosen depending on the chemical characteristics of the
materials (pH, etc.) in contact with the reinforcement, including water when applicable.
The minimum requirements have been defined in Standard NF EN 13251.
5.7.1.3. Required characteristics
The design characteristics of geosynthetics for the purpose of reinforcement are as follows:
Its long-term tensile strength, as determined based on short-term tensile strength (as per
Standard NF EN ISO 10319, see Fig. 7.14) and on the isochronous creep curves (as per
Standard NF EN 13431, Fig. V7.15);
Its stiffness, which most often is expressed based on tensile strength at a given level of
deformation (e.g. at 2%, 3%, 5% or 10%);
Its interaction coefficients at the interfaces with materials in contact with the geosynthetic
reinforcement layer.
286
Figure 7.14: Curves derived from a tensile test conducted according to Standard NF EN ISO 10319.
287
The platforms receiving the reinforcement layers must be compacted, leveled and cleared of
all angular elements or other component capable of damaging the layers (by either punching
or tearing) during the steps of their spreading, positioning and pre-tensioning.
The direct contact between the geosynthetics and inclusion heads must be avoided in order to
minimize the risks of perforation and tearing along the edges and in the corners. A level of
protection can be provided by a few-centimeter layer of fine-grained material (whether sand
or gravel) or else a punching-resistant geosynthetic material.
The reinforcement layers are spread on the platform, then cut along the length or ultimately
deployed by means of prefabricated panels. The compliance of the layer's mechanical
anisotropy with the direction of the forces the layer will be required to absorb in the structure,
is to be verified in accordance with the indications included in the execution drawings.
Figure 7.16: Layout of geosynthetic layers displaying a mechanical anisotropy under an embankment.
The geosynthetic layers must be placed so as to prevent any folds while facilitating an initial
pretensioning during embankment installation.
The circulation of construction vehicles directly on top of the geosynthetics needs to be
prohibited. An initial protective layer containing at least 20 cm of materials is necessary to
allow for the circulation of construction site vehicles or equipment.
5.7.2.2. Longitudinal overlap and end protection of the reinforcement layer
To the extent possible, reinforcement overlap in the direction of the forces to be absorbed
must be avoided.
Otherwise, the layouts that ensure reinforcement continuity between adjacent layers are
shown on the following drawings according to the justifications provided in the design
calculation.
In the case of reinforcement by two perpendicular layers on the same level, the longitudinal
overlap between two adjacent layers must be at least 30 cm.
288
Figure 7.17: Transverse and longitudinal overlap between geosynthetic layers under an embankment.
Execution conditions
289
x Anchorage with a trench cut and cover (Fig. 7.19): A trench is cut beyond the inclusion
zone to allow anchoring the reinforcement layer with a cover;
Figure 7.19: Anchorage of the geosynthetic layer by means of a trench cut and cover method.
290
References
FNTP, FFB, EGF-BTP, SBTF, USG, Syntec Ingnierie - Guide pour la ralisation des
terrassements des plates-IRUPHV GH EkWLPHQWV HW GDLUHV LQGXVWULHOOHV GDQV OH FDV GH VROV
VHQVLEOHVjOHDX, dcembre 2009.
NF EN 1097 2 -Essais pour dterminer les caractristiques mcaniques et physiques de
granulats - Partie 2 : Mthodes pour la dtermination de la rsistance la fragmentation,
juin 2010.
NF EN 1097 1 - Essais pour dterminer les caractristiques mcaniques et physiques des
granulats -3DUWLHGpWHUPLQDWLRQGHODUpVLVWDQFHjOXVXUHPLFUR-Deval), aot 2011.
NF EN 1536 - Excution des travaux gotechniques spciaux. Pieux fors, octobre 2010.
NF EN ISO 10319 - Gosynthtiques. Essai de traction des bandes larges, aot 2008.
NF EN 12699 - Excution de travaux gotechniques spciaux. Pieux avec refoulement de sol,
mars 2001.
NF EN 12716 - Excution des travaux gotechniques spciaux. Colonnes, panneaux et
structures de sol-ciment raliss par jet, octobre 2001.
NF EN 13251 - Gotextiles et produits apparents.&DUDFWpULVWLTXHVUHTXLVHVSRXUOXWLOLVDWLRQ
dans les travaux de terrassement, fondations et structures de soutnement, septembre 2001.
NF EN 13431 - Gotextiles et produits apparents. Dtermination du comportement au fluage
en traction et de la rupture au fluage en traction, novembre 2000.
NF EN 14199 - Excution des travaux gotechniques spciaux. Micropieux, septembre 2005.
NF EN 14475 - Excution de travaux gotechniques spciaux. Remblais renforcs, janvier
2007.
Execution conditions
291
292
Summary
CHAPTER 8
293
Figure 8.1: Results of loading tests conducted on inclusions of various lengths (Chelles experiment).
294
Figure 8.2: Distribution of forces in the inclusion by taking into account the negative skin friction and
neutral plane position.
2.2.1.2. Case where the inclusions only serve to reduce settlement (Domain 2)
Generally speaking, these cases pertain to ground slabs, rafts or embankments, or else to
footings whose stability at the ultimate limit state (ULS) is provided without requiring an
inclusion.
The test load equals the design load Qmax increased by the negative skin friction FN.
A test conducted using this maximum load is called a quality test.
295
2.2.1.3. Case where inclusions provide structural stability at the ULS (Domain 1)
These cases generally pertain to footings whose stability at the ULS cannot be ensured
without introducing an inclusion.
The test load is equal to: Dq u Q + 2 DFN FN, where Dq = 1.5 and DFN = 1.5.
(with Q being the load at the inclusion head).
A test performed with this maximum load is called a load-bearing capacity test.
2.2.1.4. Value of FN
FN is evaluated via the following relation: qs x inclusion perimeter x neutral plane depth.
qs must be the highest value per layer from among:
The maximum positive skin friction value expected for this type of inclusion depending on
the type and composition of the soil being crossed;
The maximum negative skin friction value output by the computation model;
The value obtained by the test defined in Section 2.2.2.
Remark: To interpret the value derived for FN, another solution consists of
neutralizing the shaft friction above the neutral point during the test.
2.2.2. Estimation of friction in soft soil
A compression test on a short inclusion conducted over the height of the soft soil layer
serves to estimate the skin friction value within this layer.
Remark: The inclusion length is typically equal to the depth of the neutral point.
According to this configuration, the contribution of the tip effect is small, which
justifies the compression test for its ease of implementation.
2.2.3. Test frequency
Loading tests are mandatory.
It is acceptable to opt not to conduct loading tests for small-sized projects (with a total
cumulative length of less than 2,000 lm for instance).
In this case, the geotechnical type verifications (GEO) are carried out with an additional
reduction factor of 1.5. This factor is applied to the characteristic values of tip resistance and
shaft friction below the neutral plane.
This point addresses the verifications described in the following sections of Chapter 5:
1.1.3: Alternative verification of bearing capacity SLS;
1.2: Ultimate limit state ULS;
4.1: Inclusion behavior.
Remark: A similar reduction factor of 1.5 related to material strength is already
included in the k3 coefficient values given in Table 5.2 when no static loading test
is performed.
296
The frequency of these tests is set according to the prescriptions listed in Table 8.1, with at
least 2 tests per independent structure and/or per designated homogeneous geotechnical zone.
Table 8.1: Testing frequency.
Number of inclusions
1 per group of 75
1 per additional group of 150
297
If settlements are found to be significantly less, the eventual impact on structural design
parameters should be assessed.
3. INFORMATION TESTS
Information tests are conducted before the general site production adjacent to representative
boreholes of each zone considered to be homogeneous from a geomechanical perspective.
These tests are to be completed and compared with results obtained from tests carried out
when producing the inclusions.
These information tests are intended to:
Verify that the anticipated depths as well as the geometric characteristics are feasible using
the designated equipment;
Adjust inclusion length whenever possible to identify the anchorage layer;
Control for material consistency (concrete or mortar), whether delivered or produced onsite.
The inclusions developed during these tests become an integral part of the project soil
improvement steps. With the project Engineer's approval, other information tests may be
undertaken in order to establish correlations between the various project zones, should their
importance be considered necessary.
These tests are also intended to determine the execution and concreting parameters for
inclusions, on an as needed basis. Information tests are to be customized for each individual
mode of inclusion execution, which will be discussed in detail below.
Remark: It is highly recommended to perform these tests in the presence of both
the Engineers representative and the soil engineering consultant.
If such is not the case and should an anomaly appear when conducting these
information tests (e.g. noncompliant soil strength values, inclusion either too short
or too long relative to the boreholes, considerable loss of concrete), the Project
Owner, Project Engineer and geotechnical specialist (soil engineering firm) must
be immediately notified in order to issue a ruling on the adaptation or
optimization of the geotechnical structure proposed by the developer.
Moreover, these tests must be written up in a report by the project developer, with
distribution to all project actors as expeditiously as possible.
3.1. Inclusion-driving method
This technique pertains to either metal or prefabricated concrete inclusions directly driven or
to inclusions made of concrete cast inside a driven closed steel tube.
During inclusion installation, a pile-driving test sheet is established, consisting of:
Inclusion identification number and type;
Date and time of the beginning and end of the driving operation;
Type of drop hammer used (single-acting, double-acting, diesel, etc.);
Weight of the striking mass and the mass being struck (e.g. rigid inclusion, follower pile,
pile helmet), as well as the driving energy;
298
The number of hammer strikes necessary in order for the inclusion to penetrate a given
length, typically 10 cm. This reading is taken over the entire penetration length in the soil by
indicating the height(s) the hammer drops or the driving energy;
Elastic refusal at the end of the driving operation;
Refusal during the final three rounds of 10 strikes of the drop hammer;
Any incidents during the driving operation (obstacles interfering with penetration,
displacements of neighboring inclusion heads, interruptions to the driving process, etc.);
The concrete volumes ultimately introduced (for the inclusions cast in place).
From this test sheet, the driving plot can be derived as follows (Fig. 8.3).
In the case of a driven tube, keep in mind that the tube's base is composed of either a lost plate
or a valve.
In the case of concrete inclusions implemented inside a casing, both the driving curve and
concreting curve need to be provided for each test. Any eventual loss of concrete must be
reported. The tube lifting (extraction) speed would also need to be indicated.
3.2. Boring with or without soil displacement
The boreholes are monitored and described in borehole data sheets.
Whenever the drilling technique so allows, a detailed soil cross-section must be established by
specifying, to the extent possible, the color, type texture, odor and humidity of the soil in
order to verify whether the soils being crossed are in fact those described by the surveying
campaign.
Alternatively, a continuous parameter recording is produced so as to offer the opportunity to
generate a summary geotechnical cross-section, in addition to ensuring that the required
inclusion embedment is achieved. These parameters shall be displayed in real time during
execution of the inclusion.
Moreover, should drilling be of the simple type (see section on the modes of inclusion
execution), then it would be necessary to ensure excavation wall stability prior to the
concreting step.
299
During concreting, the injection pressure (should the technique so allow), the volume of
concrete set into place and the pumping rate need to be specified. Any overconsumption or
under-consumption and eventual loss of concrete must also be notified.
During execution of the borehole, an execution fact sheet is produced (Fig. 8.4), detailing:
Identification number and type of pile;
Date and time of the beginning and end of the drilling operation;
Characteristics of the various boring tools employed (drill bit, valve, hammer grab, auger,
etc.), along with the eventual associated casing;
Drilling parameter recording curves, should the technique implemented yield such curves,
which if available are to be completed with any observation entry aiding the step of
identifying the crossed layers (i.e. indices collected during execution);
Description of characteristic samples for each soil layer crossed, as the technique allows;
Any boring-related incidents (rockslides, cavities, accidental loss of drilling fluid, water
inflow/seepage, etc.);
Concreting curve (Fig. 8.5) should the technique so allow, or else concreting parameter
recordings.
300
301
Date and time of the beginning and end of the vibro-driving operation;
Curves presenting recordings of vibro-driving parameters, as permitted by the technique.
These curves are complemented by any observation aiding the identification of layers crossed
(indices collected during execution),
Characteristics of the various tools implemented and the casings;
Any incidents (sinkholes, water inflow/seepage, premature refusal, etc.);
The volume of concrete placed.
4. SHAFT CONTROL TESTS
4.1. Structural integrity
Verification of the structural integrity of a concrete inclusion consists of conducting a test
involving reflection or impedance.
These tests are mandatory.
The only exception is dedicated to the Domain 2 inclusions on operations of more minor
importance (e.g. threshold on the order of 2000 linear meters).
These tests are conducted according to the NF P 94-160-2 Standard (reflection-based method)
and the NF P 94-160-4 Standard (impedance method).
7KHIUHTXHQF\SUDFWLFHGIRUWKLVWHVWLQJFDPSDLJQHTXDOVIRULQFOXVLRQVIURP'RPDLQ
ZLWK D PLQLPXP RI LQFOXVLRQV ZKHUHDV WKLV IUHTXHQF\ FKDQJHV WR IRU LQFOXVLRQV
IURP'RPDLQZLWKDPLQLPXPRI
Remark: These tests are not considered as reinforcement control from the
standpoint of the National Application Standard of Eurocode 7 (NFP 94-262:
Deep foundations).
4.2. Diameter (isolated inclusion / inclusion group)
Inclusion diameter verification, when the inclusion has been cast in place, must be
systematically performed upon the initiation of each project at a rate of 1 inspection per group
of 500 inclusions or 5,000 linear meters placed.
The inclusion is executed in the immediate vicinity of the project footprint. Once the concrete
has been set, the inclusion is released over its maximum possible height and then ultimately
extracted over this same height. Its diameter or circumference is then measured and any
variation in diameter or circumference requires a specific analysis.
Remark: A variation in diameter or circumference may be due to:
The execution method, causing excessive soil disturbance;
A pumping pressure either too strong or too weak;
A soil either too stiff or too soft;
Hydrogeological conditions.
Level by level, these variations in diameter or circumference must be taken into
account in the design, for increases in excess of 30% or for any kind of decrease
(Fig. 8.6).
302
303
304
The elevation readings after the cutting-off, along with inclusion implementation, must be
systematically controlled.
5.2. Under the ground slab, raft and embankment
In the case of a low leveling elevation, the elevation readings following the ultimate cuttingoff step and inclusion implementation must also be systematically controlled.
In these cases, it must be verified that the leveling surface does not extend below the base of
the working platform.
6. QUALITY AND THICKNESS CONTROL OF THE LOAD TRANSFER
PLATFORM
6.1. Thickness
Each of the contractors and subcontractors involved on a project site must provide a set of asbuilt drawings with elevation readings for the project structures, in accordance with the firm's
execution monitoring requirement. The reconciliation of these level recordings with the
design parameters is performed by the Project Engineer as part of the assigned task of
providing geotechnical supervision services for project execution.
Total thickness must be verified at least by means of simple comparison between the leveling
of the working platform performed after reprofiling (see Chapter 7) and the leveling
introduced following execution of the transfer platform. To control the load transfer platform
thickness above the inclusion heads, the leveling of inclusion heads after an eventual cuttingoff step must also be taken into account.
The leveling measurement frequencies are defined in the works contract documents. As a
default value, the following recommendations have been issued:
For extended structures (e.g. raft, ground slab, embankment):
- a minimum of 3 points,
- a leveling point, at a rate of:
- 1 every 400 m, until reaching a surface area of 2,000 m2,
- 1 every 1,000 m beyond the 2,000 m threshold;
Under a footing:
- 1 leveling point in alignment with each footing.
Remark: In the case of inclusions executed by soil displacement within a
saturated clayey soil with a dense mesh pattern, it would be necessary to control
the eventual lifting of the working platform before building the load transfer
platform.
6.2. Quality
Load transfer platform quality must be controlled in alignment with the inclusion as well as in
between inclusions. The load transfer platform material must be the target of identification
and compaction controls.
305
306
307
Figure 8.7: Depiction of instrumented structures within the scope of the ASIRI national project.
308
The type of sensors introduced must be adapted to the settlement being measured, which
tends to be quite small. The sensors used to monitor embankment settlements on soft soils
(settlement balls, settlement plates, etc.) may prove unsuitable due to both their level of
accuracy and their size.
For a preliminary test block, the number of rigid inclusions must be high enough to ensure
that the measurement is not adversely affected by edge effects:
In a rectangular mesh configuration for example, sixteen inclusions yield a measurement of
the central mesh under satisfactory conditions (Fig. 8.9a);
To measure the amount of load transferred onto an inclusion, a 9-inclusion test block would
be sufficient (Fig. 8.9b).
Figure 8.9: Sensor implementation zone determined as a function of the intended result:
a) mesh measurement; b) load transfer on an inclusion.
To ensure that the instrumentation remains efficient and accepted by the various project
actors, it must be designed so as to impede construction phasing as little as possible. Along
these lines, the settlement sensors embedded into the soil are preferred over settlement meter
gauges crossing the structure and capable of either slowing construction progress or being
damaged (Fig. 8.10).
309
Sensor size and measurement range both need to be adapted to the target structure.
The acquisition mode (automatic vs. manual) and frequency (greater over the construction
phase than during service) must take into account the execution phasing.
To protect the load transfer mechanisms from being disturbed by sensor installation, the setup step must be integrated within the scope of normal works phasing.
Remark: The fact of installing the sensors afterwards (e.g. inside a ditch or
excavation) may locally modify material layer properties while interfering with
measurements.
Special attention must be paid to implementing the instrumentation and conducting
earthworks in the vicinity of the sensors:
In case of a binder-based treatment on the load transfer (mixing) platform;
During compaction of the load transfer platform;
During excavation of the pit or trenches.
The instrumentation must at least enable measuring the structural settlement. Other
measurements may prove pertinent as well, i.e.:
Measurement of load transfer onto rigid inclusions;
Pore pressure measurement in the soft soil;
Rigid inclusion settlement measurement;
Measurement of reinforcement layer deformation on the load transfer platform:
geosynthetics, latticework;
Ground slab deformation measurement;
Measurement of both lateral displacements and rigid inclusion angles of inclination at the
foot of the slope in the presence of an embankment or structural perimeter.
9.1. Settlement measurement
The settlement may be measured along the vertical and/or horizontal profiles.
9.1.1. Measurement along a vertical profile
In the case of an embankment on soft soil, it may be worthwhile to measure the vertical
profile of the settlement beginning at the substratum layer and extending to the top of the
embankment. Such a measurement may entail placing a borehole magnetic extensometer,
310
though its use requires embankment implementation precautions that must be taken into
consideration by the earthworks contractor.
The settlement of soft soil can also be measured using a rod extensometer. This device offers
the advantage of being buried and not interfering with implementation of the structure.
For any instrument placed into vertical boreholes, every assurance must be made that the
sealing grout exhibits a sufficiently low stiffness so as not to alter the measurements.
Reference should be made to the recommendations issued by Peignaud and Chaput (1983).
9.1.2. Measurement along a horizontal profile
For measuring horizontal profiles, the sensors introduced should be equipped with an
appropriate precision and measurement mode; moreover, they should be of a size that does
not modify the behavior of the surrounding medium.
It may be beneficial to measure horizontal profiles at the level of inclusion heads above the
load transfer platform.
Horizontal profiles can be measured either by sending probes into a casing installed ahead of
time (hydrostatic profilometer, horizontal inclinometer) or by measuring the vertical
displacement of isolated sensors laid out in a row (settlement meter gauges, pressure sensor
for the liquid level measurement - called a transmitter, whose operating principle will be
explained in Section 9.1.3 - and settlement meter cells).
The primary advantage of probe-based measurements is that only the casing is lost. On the
other hand, the probe dimensions necessitate a measurement increment; the introduction of
casings requires cutting trenches, to be filled by sand in the transfer platform, which then is
capable of locally modifying the settlement.
In the case of specific structures (e.g. reservoirs), the set-up may be directly inserted into the
raft of the given structure.
This type of measurement also imposes preparing two points to serve as references on both
sides of the profile, whose monitoring must also be extended to the leveling. Lastly, the
precision of both the measurement and its position within the tube is considered
unsatisfactory whenever the expected settlements are small; in addition, the differential
settlement needs to be clearly identified between rigid inclusions and soft soil.
Remark: In light of these difficulties, the settlement measurement protocol laid
out herein may prove complicate to perform.
9.1.3. Point-specific measurements
Settlement meter gauges are inexpensive; their use however causes structural implementation
difficulties along with the need for external intervention in order to perform the
measurements.
The ball-type settlement meter cells are too cumbersome to produce a precise settlement
measurement.
311
Experience has shown that the use of transmitters for liquid level measurements (Fig. 8.11)
is the best adapted in terms of precision, space requirements, installation and acquisition:
These transmitters are connected in series by means of both a hydraulic line to a reservoir,
filled with antifreeze and fastened onto a support outside the structural footprint, and an
electrical line running to the recording box;
The reservoir ensures that the system always remains saturated and at constant load;
Each transmitter measures the pressure variation between its position and the reservoir
level;
A reference transmitter is fastened beneath the reservoir and serves to calculate the
settlement of the other transmitters placed underneath the structure. Each transmitter is
compensated in atmospheric pressure by a capillary running along the electrical cable.
This technique requires taking great precaution both in localizing the benchmark and in
installing the sensors.
Figure 8.11 presents a basic configuration for measuring horizontal profiles at the inclusion
head level and above the load transfer platform using transmitters. A two-reference solution
is preferred should the project site provide a feasible layout.
9.2. Load transfer control
The load transfer operation may be measured through the use of earth pressure cells (sensors)
installed either on the rigid inclusion head, within the rigid inclusion, on the soft soil, or
above and inside the load transfer platform (Fig. 8.12).
Nonetheless, installation on the rigid inclusion head is preferable to that on soft soil, as the
pressure measurement proves to be more reliable when the sensor is placed on a rigid surface;
in this case, it is more effective to use circular sensors of the same diameter as the rigid
inclusion.
The measurement range of sensors positioned on rigid inclusions must be selected under the
hypothesis that the entire load of a mesh lies on top of the inclusion.
312
Figure 8.12: Position of earth pressure cells used to indicate load transfer.
313
Fiber-optic sensors inserted into geosynthetic layers are capable of measuring the point
deformations or else deformations all along the optical fiber. The use of optical fibers has
yielded excellent results in terms of precision, durability and ease of implementation,
provided a few precautions have been strictly followed during installation (ASIRI Report
2.07.1.01, 2007) (Fig. 8.13).
314
References
ASIRI Rapport davancement de Chelles , 2.07.1.03, 2007.
G 38060 Textiles. Articles usages industriels. Recommandations pour lemploi des
JpRWH[WLOHVHWSURGXLWVDSSDUHQWpV0LVHHQXYUH6SpFLILFDWLRQV&RQWU{OHGHVJpRWH[WLOHV
et produits apparents, juin 1994.
NF P 94-500 - 0LVVLRQVGLQJpQLHULHJpRWHFKQLTXH&ODVVLILFDWLRQHWVSpFLILFDWLRQV.
NF EN 1997-1 (P 94-251-1) - Eurocode 7 Calcul gotechnique. Partie 1 : Rgles
gnrales .
NF EN 1997-1/NA (P94-251-1/NA) - Eurocode 7 Calcul gotechnique. Partie 1 : Rgles
gnrales . Annexe nationale la NF EN 1997-1, 2005.
NF EN 1997-2 (P94-252) - Eurocode 7 Calcul gotechnique. Partie 2 : Reconnaissance des
terrains et essais .
NF P 94-150-1 - Sols : reconnaissance et essais. Essai statique de pieu sous effort axial.
Partie 1 : En compression .
NF P 94-150-2 - Sols : reconnaissance et essais. Essai statique de pieu sous un effort axial.
Partie 2 : En traction .
NF P 94-160-2 - 6ROV UHFRQQDLVVDQFH HW HVVDLV $XVFXOWDWLRQ GXQ pOpPHQW GH IRndation.
Partie 2 : Mthode par rflexion .
NF P 94-160-4 - 6ROV UHFRQQDLVVDQFH HW HVVDLV $XVFXOWDWLRQ GXQ pOpPHQW GH IRQGDWLRQ.
Partie 4 : Mthode par impdanc e.
NF P94-262 (P94-262) - Calcul gotechnique. Fondations profondes.
NF P 11-300 - Excution des terrassements : classification des matriaux utilisables dans la
FRQVWUXFWLRQGHVUHPEODLVHWGHVFRXFKHVGHIRUPHGLQIUDVWUXFWXUHVURXWLqUHV.
NF P 94 117-1 - Sols : reconnaissance et essais. Dformabilit des plates-formes. Partie 1 :
Module de dformation statique la plaque .
NF P 94 117-2 - Sols : reconnaissance et essais. Portance des plates-formes. Partie 2 :
Module sous chargement dynamique .
NF P 94 117-3 - Sols : reconnaissance et essais. Portance des plates-formes. Partie 3 :
Coefficient de raction de Westergaard sous chargement statique dune plaque .
NF P 94-078 - Sols : reconnaissance et essais. Indice CBR aprs immersion ; Indice CBR
immdiat ; Indice portant immdiat. Mesure sur chantillon compact dans le moule
CBR.
NF P 94-093 - Sols : reconnaissance et essais. Dtermination des rfrences de compactage
GXQPatriau. Essai Proctor normal ; Essai Proctor modifi.
NF P 11-213-1 (DTU 13-3) - Dallages. Conception, calcul et excution. Partie 1 : Cahier
des clauses techniques des dallages usage industriel ou assimils . Homologue mars
2005.
NF P11-213-1/A1 (DTU 13.3) - Dallages. Conception, calcul et excution. Partie 1 : Cahier
des clauses techniques des dallages usage industriel ou assimils . Amendement
homologu mai 2007.
NF P 11-213-2 (DTU 13-3) - Dallages. Conception, calcul et excution. Partie 2 : Cahier
des clauses techniques des dallages usage autre quindustriel ou assimils .
Homologue mars 2005.
NF P 11-213-2/A1 (DTU 13-3) - Dallages. Conception, calcul et excution. Partie 2 :
Cahier des clauses techniques des dallages usage autre quindustriel ou assimils .
Amendement homologu mai 2007.
NF P 11-213-3 (DTU 13-3) - Dallages. Conception, calcul et excution. Partie 3 : Cahier
des clauses techniques des dallages de maisons individuelles . Homologue mars 2005.
315
316
Summary