Vous êtes sur la page 1sur 40

ISP7-PRESSIO 2015. Frikha, Varaksin & Gambin (Eds.

) © 2015

THE USE OF PRESSUREMETER TESTS FOR MODELING RESIDUAL SOILS


GEOMECHANICS AND FOUNDATION BEHAVIOUR

L’UTILISATION DES ESSAIS PRESSIOMETRIQUES POUR LA MODELISATION DE LA


GEOMECANIQUE DES SOLS RESIDUELS ET LE COMPORTEMENT DES FONDATIONS
1 1 2
António VIANA DA FONSECA , António TOPA GOMES , Roberto Quental COUTINHO
1 University of Porto, Faculty of Engineering, Civil Engineering Dep., Porto, Portugal
2 Federal University of Pernambuco, Recife, Brazil

ABSTRACT – The assessment of mechanical characteristics of residual soils, especially when


these are derived from the interpretation of in situ testing techniques, is highly conditioned by
important factors, such as: microstructure, cohesive-frictional characteristics, stiffness non-
linearity, small and large strain anisotropy, weathering and destructuration, condition of
saturation, consolidation/permeability characteristics and rate dependency. These specificities
will be discussed in the light of parametrical correlation proposals develop for these soils, both
saprolitic as lateritic. Conventional empirical techniques and numerical modelling, using a
“distorted” hyperbolic elastic model coupled with Mohr-Coulomb failure, will be presented as a
close formulation to derive constitutive parameters. Additionally, the analysis of the axially and
laterally load-tested piles in a particular residual soil, by recourse of PMT semi-empirical curves
will be discussed in the light of their singularities.

RÉSUMÉ – L'évaluation des caractéristiques mécaniques des sols résiduels, en particulier


lorsque celles-ci sont obtenues à partir de l'interprétation des essais in situ, est fortement
conditionnée par des facteurs importants, tels que: la microstructure, les caractéristiques
cohésive-frictionnel, la non-linéarité de la rigidité, les petite et grande anisotropie de contrainte,
les intempéries et déstructuration, les condition de saturation, la consolidation / caractéristiques
de perméabilité et de la dépendance de taux. Ces spécificités seront discutés à la lumière des
propositions de corrélation paramétriques développent pour ces sols, tant saprolitique comme
latéritique. Techniques empiriques classiques et modélisation numérique, en utilisant un
modèle élastique pseudo-hyperbolique couplé avec le critère de rupture Mohr-Coulomb, seront
présentés comme une formulation proche de dériver des paramètres constitutifs. En outre,
l'analyse de la réponse de pieux testée en chargement axial et latéral dans un sol résiduel, en
ayant recours des courbes semi-empiriques du PMT sera examinée compte tenu de leurs
singularités.

1. Introduction

The assessment of mechanical characteristics of natural soils implies new techniques or,
better, new interpretation methods (Viana da Fonseca and Coutinho, 2008). Non saturated
conditions prevail in most of soils and will have to be dealt carefully. Interpretation methods for
the knowledge of bonded geomaterials, such as residual soils, are as much challenge as the
complexity of their behavior. Cross-correlation of different and multiple measurements from
different tests is one preferential solution: more measurements in one test the better!
Residual soils, particularly from igneous rocks, are geologically formed by upper layer of
heterogeneous soil masses of variable thickness, overlaying more or less weathered rocks.
Variable weathering processes, temperature, drainage, and topography, have reduced the
rocks in-place to form overburden residual soils that range from clay topsoil to sandy silts and
silty sands that grade with depth back into saprolite and partially-weathered rocks. As
weathering proceeds, the stress release as a result of the removal of the over-lying material
accelerates the rate of exfoliation (stress release jointing) and the wetting and drying processes

33
ISP7-PRESSIO 2015. Frikha, Varaksin & Gambin (Eds.) © 2015

in the underlying fresh rock. These processes increase the surface area of the rock on which
chemical weathering proceeds, which leads to deep weathering profiles. Attempts have been
made (ex: Irfan, 1996; Viana da Fonseca, 1996) to correlate relationships between the degrees
of weathering and engineering properties of weathered rocks with reasonable success.
However, for practical design and construction purposes these correlations are not sufficiently
stable and accurate.

In the weathering profile (the sequence of horizons that is formed) one can possibly found
materials varying from sound rock, weathered rock, soil which retain rock characteristics (called
young residual soil or saprolitic soil) to the horizon where no remaining rock characteristics can
be seen (structure and mineralogy – knows as mature residual soil or lateritic soil). In this top
layer the presence of transported soil may also occur (ex. colluvium), which may be difficult to
distinguish from the true residual soils. A general view of many soils which are called unusual
soils (see Schnaid et al. 2004a; Coutinho et al. 2004a) where is included bonded soils –
granitic saprolitic soil and laterite soils, unsaturated – collapsible soils. Bonding and structure
are important components of shear strength, in general residual soils present a cohesive-
frictional nature (characterized by c’ and f’). Anisotropy derived from relic structures of the
parent rock can be also a characteristic of a residual soil. In those conditions, the structure
formed during the weathering evolution / process can become very sensitive to external loads,
requiring adequate sample technique in order to preserve that relic structure. Usually, the
resulting void ratio and density of the soil are not directly related to stress history, unlike the
sedimentary soils (Viana da Fonseca, 2003). The presence of some kind of bonding, even
weak, usually implies the existence of a peak shear strength envelope, showing a cohesion
intercept and a yield stress which marks a discontinuity in stress strain behaviour. Examples of
observed yield stresses in residual soils are found, in Viana da Fonseca (1988, 1996, 1998,
2003), in Coutinho et al. (1998, 2000) – from gneiss, (Viana da Fonseca et al. 2006) from
granite, or, in (Machado and Vilar, 2003) - sandstone and magmatic rocks.
Structure in natural soil has two “faces”: the “fabric” that represents the spatial arrangement
of soil particles and inter-particle contacts, and “bonding” between particles, which can be
progressively destroyed during plastic straining, given place for the term “destructuration”
(Viana da Fonseca and Coutinho, 2008). Most – if not all! - geomaterials are structured, but
naturally bonded soils are dominated by this effect on their mechanical response (Leroueil and
Vaughan, 1990). Here cohesive component due to cementation can dominate soil shear
strength, at engineering applications involving low stress levels (Schnaid, 2005) or in specific
stress-paths where this component is relevant (cuts or slopes, for instance).
Residual masses present strong heterogeneity, changing gradually their characteristics
laterally and vertically (with depth), especially regarding their mechanical properties. As a sign
of this, it is common to have to adopt shallow foundations - footings and mats - and driven piles
or drilled and bored piles, in very limited areas, depending upon the consistency of the
overburden soils and the depth to parent rock (Viana da Fonseca and Coutinho, 2008). An
accurate mapping of the spatial variability of the mechanical properties, necessary for
geotechnical design, is very challenging being this improved recently by the use of geophysical
methods (Viana da Fonseca et al. 2006). This will be explored later in this text. Several in situ
tests, such as conepenetrometer (SPT, CPT), dilatometer (DMT) and pressuremeter (PMT and
SBPT), and geophysical survey, surface and borehole seismic tests, electrical resistivity and
GPR, have been used to give different insight to these particular soils. Emphasis will be made
in the use of prebored (PMT) and selfboring (SBPT) tests to focus some of their advantages in
modelling residual soils geomechanics and foundation behavior. Most of the work at this level
has been made in laboratory tests carried out on natural specimens retrieved from the field, but
difficulties in testing natural soils are related to disturbance to the structure that can occur
during the sampling process (Viana da Fonseca et al, 2011). To distinguish features of
behaviour emerging from bonded structure from those related to changes in state, constitutive
laws conceived for the unbonded material are modified accordingly to introduce the bond
component (Leroueil and Vaughan, 1990; Pinyol et al, 2007; Jiang et al, 2007;Yu et al, 2007).

34
ISP7-PRESSIO 2015. Frikha, Varaksin & Gambin (Eds.) © 2015

Still, the level of disturbance has to be carefully evaluated, as proposed by (Ferreira et al, 2011)
to have confidence in these laboratory tests.
Granite residual soils are very common in the northern part of Portugal, and assume
paramount economic importance as they represent the most superficial cover of the most
populated cities in the region. In these soils, experience with pressuremeter testing in general,
and self-boring pressuremeter (SBP) testing in particular, is limited. One of the topics that has
been addressed to capitalize the good adaptability of pre-bored (PMT, Ménard type)
pressuremeter to derive ultimate load capacities of piles, different solutions available and
commonly used in foundation practice were applied in the several research programs in the
University of Porto since the eighties and nineties (Viana da Fonseca,1988,1996), towards the
first decade of the 20th millennium, with special emphasis to the period from 1998 to
2004,where the city of Porto has assisted to a “revolutionary” renew of the transportation
infrastructures, by upgrading its existing railway network to an integrated metropolitan transport
system with 70 km of track and 66 stations. Seven kilometers of this track and 10 stations were
constructed in the very first phase under the historical and densely populated city, an UNESCO
world heritage site. The underground tunnels were driven by two Earth Pressure Balance (EPB)
TBMs, with an internal diameter of 7.8 m and 8.0 m to accommodate two tracks with trains.
Line C develops in tunnel for 2,350 m from Campanhã to Trindade and has four underground
stations, a maximum cover of 32 m and a minimum of 3m before reaching Trindade station.
Line S is 3,950m long and runs from Salgueiros to São Bento with 7 stations and a maximum
overburden of 21 m details in (Viana da Fonseca and Topa Gomes, 2010). As described by
(Babendererde et al, 2004), tunnel driving started in August 2000, driving from Campanhã to
Trindade, and was originally planned that the EPB TBM would run with a partially full,
unpressurized working chamber in the better quality granite in order to take advantage of the
higher rates of advance in this mode as compared with operating with a fully pressurized
working chamber. The highly variable nature of the rock mass made it extremely difficult to
differentiate between the better quality rock masses in which the working chamber could be
operated safely with no pressure and the weathered material in which a positive support
pressure was required on the face (Babendererde et al, 2004). There is a very “fluid” difference
or similarity in the concept of Intermediate Geomaterials (IGM) and residual soils, a natural
product of extreme weathering of rock masses. As stated by (Cruz, 2010), beyond the first
three weathering degrees of ISRM classification (W1 to W3), chemical weathering is extended
to the whole massif, and so the mechanical evolution is mainly governed by increasing material
porosity, the weakening of mineral grains and the reducing bonding between grains, with the
rock massif becoming more and more friable and chemically weakened. Weathering degrees
W4 and W5 represent transition behavior where micro and macro fabrics have similar
influence, towards a residual soil-mass where the relict macro-fabric is no longer present. This
process is followed by a mechanical degradation that leads to substantial reduction of strength
and stiffness. Table I illustrates orders of magnitude typically associated to rock and soil-
masses of some reference parameters.

Table I - Geotechnical parameters (Babendererde et al., 2004)


UCS (qu) c΄ Εd
MPa (MPa) (GPa)
Rock 2 - 300 > 0,1 >400
Soil <2 < 0,1 < 300
UCS (qu)- uniaxial compression strength; c’ – effective
cohesion intercept; Ed – Young Modulus

When macrofabric is no longer represented, then a general cohesive-frictional behavior of


soil takes place, with the overall mechanical behavior governed by a wide range of factors such
as micro-structure, stiffness non-linearity, small and large strain anisotropy, weathering and de-
structuration, consolidation characteristics and flow rate dependencies (Schnaid, 2005). The

35
ISP7-PRESSIO 2015. Frikha, Varaksin & Gambin (Eds.) © 2015

author emphasized that IGM or residual soils satisfy at least one of the following criteria: (i)
classical constitutive models do not offer a close approximation of its true nature; (ii) it is
difficult to sample or to be reproduced in laboratory; (iii) very little systematic experience has
been gathered or reported; (iv) values of geomechanical parameters are outside the range that
would be expected for more common sands and clays; (v) the soil state is variable due to
complex geological conditions.

This geological and geotechnical “environment” is very favorable to the use of


pressuremeters for in situ testing, in view of the necessity to conciliate a technique that can
enable the execution of load-deformation tests in such a variable rigidity of the involved
horizons that need to be characterize and the deduction of such an enriched parametrical
information that comes out from a stress-strain data output.

2. PMT tests for prediction of ultimate load capacity on three classes of soils
constructed in saprolitic (residual) soils from Porto granite

Description of the exercise of a class A Pile Prediction Behaviour in Residual Soils (PPB-RS)

In the Fall of 2003, the Faculty of Engineering of the University of Porto (FEUP) and the High-
Tech Institute of the Technical University of Lisbon (ISTUTL) invited the international
geotechnical community to participate in a Class A prediction event on pile capacity and pile
response to an applied loading sequence for Bored, CFA and Driven Piles in Residual Soils of
Porto. An extensive site investigation had been carried out, in which several in-situ testing
techniques were used. Undisturbed samples were recovered and an extensive laboratory-
testing programme was carried out (Figure 1).

PILES
E - Bored (D=0.60m)
T - CFA (D=0.60m)
C - Driven (0.35x0.35m) 1 2 3
2.00 2.00
SITE CHARACTERIZATION
DMT2
S - Borehole with Sampling (+SPT) CPT1
CPT7 DMT7
SPT; CPT; DMT; PMT DMT1
P - Piezometer C2 T2
S1+SPT
1.75

Cross-Hole

A E7 E8
2.00
4.00

T1
CPT2 P S2

B E5 E6 DMT3 PMT3
CPT9
2.00

DMT9
S4+SPT PMT2
4.00

CPT5
C1
S3+SPT
E4
C E3 DMT8
CPT8 DMT6

A
4.00

TRENCH FOR SEISMIC PROFILES

E9
REFERENCE SLAB

CPT3
D E1 E2
S5+SPT
2.20

DMT4 PMT1 E0
CPT6

CPT4

CROSS SECTION A-A


variable (min. 0.5m; max. 0.9m)
DMT5 ground level

0.00 Reference Elevation


new platform trench
(opened at 29/09/03) (opened at 02/07/03)

Figure 1. Plan of the location of the piles and in situ tests and excavation after piles after
extraction

36
ISP7-PRESSIO 2015. Frikha, Varaksin & Gambin (Eds.) © 2015

Two each of three different kinds of piles were installed: 600 mm diameter bored piles (E0
and E9) using temporary casing, 600 mm diameter CFA piles (T1 and T2), and 350 mm
square, driven, precast concrete piles (C1 and C2). 3 piles were tested dynamically (E0, T2 and
C2), and 3 were loaded axially up to failure (piles E9, T1 and C1). Piles E0 and E9 were
constructed by first using a rotary drilling rig to install a temporary casing that was cleaned out
using a 500 mm cleaning bucket (details in Viana da Fonseca et al. 2009). The external
diameter of the cutting teeth at the bottom of the temporary casing was 620 mm. Concrete, with
slump 180 mm, was placed by tremie, casing withdrawn on completion of the concreting. Piles
T1 and T2 were constructed using a rotary drilling rig and a 600 mm continuous flight auger.
Concrete grout was ejected with a pressure of 6000 kPa at the beginning of line and flow rate
was steady at 700 l/min. Concrete slump was 190 mm and “overconsumption” was less than
6%. C-piles were driven on with a 40 kN drop hammer. Figure 2 presents some pictures of the
excavation for the analysis of the facie of these three types of construction in piles’ practice and
how this can condition the estimation of these foundations capacity, in terms of shaft, base and
overall parcels.

Dynamically Driven Precast

Continuous Flight Auger

Bored pile with


temporary casing

Figure 2. Some of the differences in the three types of piles after extraction

37
ISP7-PRESSIO 2015. Frikha, Varaksin & Gambin (Eds.) © 2015

The Predictions

For the Pile Prediction Event, the participants were provided with information on pile geometry,
soil profile, equipment and high strain dynamic test results. They were challenged to predict the
performance of the piles under static load by submitting a table of load vs. settlement at the pile
head, and also: (i) parameters and models used; (ii) calculation methodology; (iii) pile base
resistance and shaft resistance, separately if applied; (iv) ultimate compressive resistance and
indication of criteria used to determine this; (v) allowable bearing capacity, and factor of safety,
if applied, used to determine this; (v) explanation of the methods used to make the predictions.

In December 2003, a total of 33 persons from 17 countries submitted predictions, and static
loading tests on piles E9, T1 and C1 were then performed in January 2004. A summary of the
predictions and the test results is given in Viana da Fonseca and Santos (2008). A compilation
of the predictions for the ultimate compressive resistance is represented in Figures 3 a, b, c,
respectively, for piles E9, T1 and C1,. The pile base resistance (Rb) and shaft resistance (Rs)
are indicated separately when applied; otherwise the total resistance (Rt) is shown. The
predictors applied different calculation methods, such as analytical or empirical methods,
results of dynamic load tests or a combination of both. It is also important to note that different
criteria or calculation approaches were used to define the ultimate compressive resistance.

The predictions presented in the figures are very scattered demonstrating that the accurate
estimation of pile axial capacity is still a very difficult task in residual soils. For the non-
displacement piles (bored piles and CFA piles), most predictions have overestimated the
ultimate capacity, probably because the settlements induced in the static tests have not
mobilised the failure mechanism assumed in theoretical formulations for the evaluation of base
resistance. The predictions for the displacement piles (precast driven piles) were conservative,
because there was an underestimation of the strength gains due to pile installation effects. In
terms of the suitability of the different approaches, herein focused on the results of PMT
methods (Figure 4), a complete analysis can be seen in Viana da Fonseca & Santos (2008).

Predictions based on analytical methods give a very large scatter, since there is a great risk
with such methods in using the input data with no judgement. It is generally preferable to use
directly the results of in situ tests for prediction of ultimate resistance of piles. Synthesis of that
data for the ISC’2 event site is presented in Viana da Fonseca and Santos (2008). SPT based
predictions performed fairly well for the non-displacement piles, being in fairly good agreement
with the load test results for the bored pile (E9), while generally over-predicting the ultimate
capacity of the CFA pile (T1). As for CPT based predictions, these show good agreement with
the load test results for the bored pile (E9), with some scatter, but only reasonable agreement
for the CFA pile (T1), for which there were fewer predictions. For the driven pile C1, the
predicted ultimate capacity values, of which there were a large number, were very low when
compared with the load test results, with a few exceptions.

38
ISP7-PRESSIO 2015. Frikha, Varaksin & Gambin (Eds.) © 2015

(a)

(b)

(c)

Figure 3. Ultimate resistance: predictions for bored, E9 (a), CFA, T1 (b), and driven, C1 (c),
piles.

39
ISP7-PRESSIO 2015. Frikha, Varaksin & Gambin (Eds.) © 2015

(a)

(b)

(c)
Figure 4. Predictions from PMT results of ultimate resistance of bored (a), CFA (b) and (c)
precast driven piles (Viana da Fonseca and Santos, 2008).

It was concluded that analytical methods generally are very scattered and potentially unsafe,
and are therefore to be avoided. More specifically methods based on SPT are very scattered,
and may be too unsafe for CFA piles, while too conservative for driven piles. CPT based
methods show less scatter than SPT based methods, being generally reasonable for non-
displacement piles, but grossly conservative for displacement piles. PMT methods, although
not so numerous in the prediction exercise which is a sign of international practice, are clearly
the most accurate, especially for bored piles and CFA piles.

40
ISP7-PRESSIO 2015. Frikha, Varaksin & Gambin (Eds.) © 2015

PMT method for predicting axial load-settlement curves of piles (Frank, 2008)
The axial load-displacement behaviour of single piles can be predicted by elastic continuum
theory through the solutions developed by Poulos and Davis (1980), by finite elements (Poulos,
1989), and by approximate closed-form analytical solutions (Randolph & Wroth, 1978, 1979;
Fleming et al., 1985). Continuum theory defines soil stiffness by two elastic parameters: an
equivalent elastic soil modulus (Es) and Poisson's ratio (s). Four generalised cases are
considered: (1) a homogeneous case where Es is constant with depth; (2) a Gibson-type
condition where Es is linearly-increasing with depth; (3) friction or floating-type piles; and (4)
end-bearing type piles resting on a stiffer stratum.
Elastic continuum methods also provide an evaluation of axial load transfer distribution. The
fraction of load transferred to the pile base (Pb) is given by Mayne and Schneider (1999), based
on the work of Fleming et al. (1985), and additionally, a realistic non-linear load-settlement (Q-
s) model of a pile can be determined by inserting a modified hyperbolic model, as introduced
above, in the following expression:
Q  I
s(   )  (1)
 
d  E 0  1  f  (Q / Qult ) g

where, Q = applied load and Qult = axial capacity for compression loading, in a pile with d as
diameter, installed in an homogeneous soil which the stiffness degrades from the elastic
modulus (E0) in accordance to the stress-level ( Q / Qult ), as proposed by Fahey and Carter
(1993), ruled by two fitting constants (f and g), which are properties of the soil.
This simple model may be appropriate for many practical cases of foundations which are
between the pure floating pile, and piles well toed in to stiff bases with a shaft through softer
materials, either homogeneous or with stiffness linearly increasing in depth. This model was
applied in the granitic residual soil which was thoroughly studied for ISC’2, applying it to the
bored piles with temporary casing, CFA, and driven piles. Figure 5 shows the results of
applying the model solely considering a shaft embedded in homogeneous soil to such diverse
conditions, resulting in fairly reasonable reproduction of the observed load/settlement curves,
using very different parameters (Viana da Fonseca et al., 2012).
Other cases, however, the nonhomogeneity of the ground may require modelling of
successive layers in depth, and recourse to other less simple models may be necessary. This
may be particularly the case in residual soil profiles (Viana da Fonseca et al., 2012).
In this method shaft friction mobilisation “τ-z” curves and end load – end displacement curve
(q-spcurves) are used for the prediction of the load-settlement curve of single piles under axial
loading. The pressuremeter approach uses a load transfer method developed by Gambin
(1963) and Frank and Zhao (1982), where the pile is cut in a series of equal length elements
(Figure 6). According to Frank and Zhao (1982) the settlement sp of the pile base is given by:

B
sp    qp (2)
k 
 
p

where,in this notation, B is the diameter of the pile and qp is the pile base pressure, while the
skin friction qsi mobilised during the settlement of each pile shaft element i is then obtained as a
function of zsi which is the local displacement of the ith shaft element against the adjacent soil
layer (Figure 6).
kt or kq are functions of the pressuremeter modulus EM and the diameter B of the pile, with
models proposed by Frank and Zhao and shown in Figure 7:
kt = 2.0 EM/B and kq= 11.0 EM/B for fine soils
kt = 0.8 EM/B and kq= 4.8 EM/B for granular soils

Figure 8, taken from Bustamante & Gianeselli (1993). presents a comparison of measured
and calculated load-settlement relationship for a pile in Koekelare.

41
ISP7-PRESSIO 2015. Frikha, Varaksin & Gambin (Eds.) © 2015

CFA pile
f=1; g=0.22

Bored pile
f=1; g=0.12

Driven pile
f=1; g=0.3

Figure 5. Axial load-displacement behaviour of single piles calculated by elastic continuum


theory with non-linear stiffness integration (After Mayne & Schneider, 1999)

42
ISP7-PRESSIO 2015. Frikha, Varaksin & Gambin (Eds.) © 2015

Figure 6. Load transfer method to estimate pile settlement (After Frank & Zhao, 1982).

Figure 7. PMT τ-s and q-sp curves after Frank and Zhao (1982)
Load Qo (kN)

0 500 1000 1500


0
(mm)

-10
So

-20
settlement

measured

-30 calculated

-40

Figure 8. Comparison of measured and calculated load-settlement relationship for the


Koekelare pile (After Bustamante & Gianeselli, 1993).

43
ISP7-PRESSIO 2015. Frikha, Varaksin & Gambin (Eds.) © 2015

Lateral pile loading

Piles are often subject to lateral loading, and should be designed considering the lateral
interaction with the ground. The methods using the subgrade reaction modulus (or p-y reaction
curves, where p = reaction pressure, and y = horizontal displacement) are well known for the
design of piles under lateral loads. These methods, which consider the pile to be a beam on
linear or non-linear elastic spring, are supported by the Winkler theory:
d4y
EI  k B  z  0 (3)
d z4
The value of k is readily obtained from the settlement equation w = f(EM) given by Ménard
(1963) for an infinitely long strip footing, of width B, since k.B = p/w. In this condition, the
ground reaction is similar to that observed under a PMT borehole expansion (Gambin & Frank,
2009). When EM values are averaged, for B larger than 0.6 m, below the critical depth:

 18 
k B  Es  Es    (4)
 
 4  2.65  B / B0   B0 / B  3  
where, α is the Ménard rheological factor ( 1/4 < α < 2/3) and Bo is a reference diameter
equal to 0.6 m. According to Gambin and Frank (2009), this was checked as early as 1962
using PMT data on various laterally loaded prototype piles. The present design rules (Frank,
1999), using generalised p-y curves, include the decrease of k as y increases. In Figure 9, four
different loading conditions are shown to be dependent on the creep (or yield) pressure pc
(which can be estimated as: pc = pL/2).
Gambin and Frank (2009) present a small modification to Equation (3) where a soil applies
a horizontal thrust to a pile (Figure 10). The equation becomes:

d4y
EI  k B   y ( z )  g ( z )  0 (5)
d z4
where, y(z) is replaced by [y(z) – g(z)], g(z) being the free horizontal displacement of the
soil without pile.

Figure 9. Soil reaction against lateral displacement for actions at head level (After Frank, 1999),
for: a) permanent actions at pile head; b) soil lateral thrust; c) short time actions at pile head;
d) unexpected instant actions at pile head.

44
ISP7-PRESSIO 2015. Frikha, Varaksin & Gambin (Eds.) © 2015

Figure 10. Example of a pile subjected to earth pressure (After Frank, 1999)

PMT data interpretation and analysis of the behavior of laterally loaded piles in ISC’2
experimental site

A description of the data interpretation and analysis of the behaviour of laterally loaded piles in
ISC’2 experimental site using PMT and DMT based methods has been given by Tuna (2006)
and Tuna et al. (2008).

Bored piles E0 and E9 had 6 m net embedment depth, and reaction piles E1 to E8, all 22 m
long, were embedded in bedrock, being their construction process described in Tuna et al.
(2008), as well as CFA piles (T1 and T2), about 6 m depth and 60 cm diameter, and driven
piles (350 mm square). Piles were laterally loaded by pushing each pair apart (E0-E1, E7-C2
and E8-T2) using a hydraulic jack. Measurement of displacements (vertical and horizontal) and
rotations of pile heads were made using LVDTs. Detailed inclinometer measurements were
made in piles E0, E1 and T2, to define the development of these displacements and rotations
alongt he full length of the piles. To determine the bending moment in pile E0, strains in the
concrete were measured using retrievable extensometers (Figure 11). Pile head lateral
displacements, as measured using the LVDT and also from inclinometer integration, are
presented in Figure 12.

Tuna et al. used the p-y curve method based on PMT (MPM) parameters suggested by
Ménard et al. (1969) through equations (6 and (7), where B is the pile diameter, B0 is the
diameter of the reference pile (taken equal to 0.6 m) and α is Ménard’s rheological factor,
varying between 0.25 and 1.
18  EM
For B>B0: K  (6)
4  2.65  B / B0   B0 / B  3

18  EM
For B<B0: K  (7)
4  2.65  3

45
ISP7-PRESSIO 2015. Frikha, Varaksin & Gambin (Eds.) © 2015

The reaction curve is modified for depth values, z, less than a critical depth zc, by multiplying
the pressure by a coefficient equal to 0.5(1+z/zc). For cohesive soils, zc is about 2B and for
granular soils about 4B

b)

c)
Figure 11. (a) Assemblage of the test with reference beam; (b) structure; (c) instrumentation:
inclinometer, retrievable extensometer and LVDT mounted horizontally.

600

500

400 Estaca E0 - LVDT - 1º ensaio


E0 - LVDT - 1º test
Load (kN)

Estaca E0 - inclinó metro - 1º ensaio


E0 - inclinometer - 1º test
Estaca E0 - LVDT - 2º ensaio
300 E0 - LVDT - 2º test
Estaca E0 - inclinó metro - 2º ensaio
E0 - inclinometer - 2º test
Estaca E1- LVDT - 1º ensaio
E1 - LVDT - 1º test
Estaca E1- inclino metro - 1º ensaio
200 E1 - inclinometer - 1º test
Estaca E1- LVDT - 2º ensaio
E1 - LVDT - 2º test
Estaca
C2 C2 - LVDT
- LVDT
100 Estaca T2 - LVDT
T2 - LVDT
T2 - inclinometer
Estaca T2 - inclinó metro

0
0 50 100 150 200 250 300
displacement (mm)

Figure 12. Load versus observed pile head lateral displacement

46
ISP7-PRESSIO 2015. Frikha, Varaksin & Gambin (Eds.) © 2015

The development of limit pressure (pL), creep pressure (pF) and pressuremeter modulus
(EM), with vertical effective stress (depth), is defined by the following equation (8), (9) and (10);
a value of 0.33 was adopted for α:
pL (kPa)  6.4s 'vo (kPa)  1017.2 (8)
pF (kPa)  2.4s 'vo (kPa)  365.3 (9)
E M (kPa)  92.5s 'vo (kPa)  12949 (10)
Additionally, a study was made in order to adjust the soil strength to the points A and B (see
Figure 13) by power laws. The main purpose was to decrease the high values of pu, obtained
for depths less than 0.5 m. For example, for the bored pile E0, pu was approximated by a power
law as shown in Figure 14.

Displacements of pile heads were calculated by the application of the original method
(model A) proposed by Ménard et al. (1969) and by applying the power law approximation
(model B). Figure 15 shows the calculated and observed displacements. For the driven pile, the
displacements by both methods are generally underestimated. For the bored and CFA piles,
model A overestimates the calculated displacements for small loading levels, but
underestimates then for higher loading levels. Model B overestimate the displacements for the
bored piles, while it predicts fairly well for CFA piles.
p [FL-1 ]

pu=pL B
B C z>zc
pL B

K/2
A B'
pf B
pu=pL B/2
z=0

K/4
K

K/2

y [L]
O

Figure 13. P-y curve by Ménard et al. (1969)

1200

1000

800
pu (kN/m)

600

0,4044
400 y = 530.55x
2
R = 0.99
200

0
0 1 2 3 4 5 6 7
z (m)

Figure 14. Pile E0 – linear relationship of pu with depth, with power law and trend line

47
ISP7-PRESSIO 2015. Frikha, Varaksin & Gambin (Eds.) © 2015

140
E1 - Model A E1 - Model B
E0 - Model A E0 - Model B
120
T2 - Model A T2 - Model B

calculated displacement (mm)


C2 - Model A C2 - Model B
100

80

60

40

20

0
0 20 40 60 80 100 120 140
measured displacement (mm)

Figure 15. Measured and calculated displacements using Ménard et al. (1969) method

From this study, it can concluded that PMT method, originally proposed by Ménard et al.
(1969), generally underestimates horizontal displacements in residual soils, mostly at high
loading levels. This can be corrected by decreasing soil strength for shallow depths. Both the
Ménard proposal and the one designated as “modified”, overestimated displacements for very
small loading levels, as a consequence of the excessively low value considered for the reaction
modulus (K). Even so, it can be seen that this method leads to a reasonable simulation of the
behaviour of bored (E0, E1) and CFA (T2) piles. For driven piles (C2) this method is unreliable,
because the power law considered for the variation of pu with depth, is inadequate to model the
behaviour of this type of pile (Viana da Fonseca et al., 2012).

3. Modeling complex geotechnical structures in cohesive-frictional residual soils

3.1. Strength from in situ tests

Empirical correlations are very useful for the purpose of common design of foundations and
other simple geotechnical structures and have been derived from an accumulated database
(Gomes Correia et al. 2004, Viana da Fonseca et al., 2004, 2006). SPT and CPT can be used
to predict the peak angle of shear resistance in granular soils when normalised for stress-level
with pa= 100 kPa:
N 60 qc
N1 60  q c1  (11)
sv / pa  0,5
s v / p a 0,5
Hatanaka and Uchida (1996) obtained the equation (12) for the evaluation of the angle of
shearing for peak, also corroborated by Mayne (2001) for a residual silty sand in Atlanta and
Georgia:
f p  15.4N1 60  0.5  20º (12)

Robertson and Campanela (1983) and Kulhawy and Mayne (1990) recommended,
respectively, for unaged, uncemented quartz sands, the following correlation with CPT:
f p  arctan0.1  0.38  logq c / s v0  (13)

f p  17.6º 11.0  log q c1  (14)

48
ISP7-PRESSIO 2015. Frikha, Varaksin & Gambin (Eds.) © 2015

In residual soils from Porto, values of (N1)60, taken from the SPT tests, allowed to derive the
angle of shearing resistance from Décourt’s (1989) or Hatanaka and Uchida (1996). Viana da
Fonseca et al. (1998) presented a proposal, adapting for an average state-of-practice with 60%
efficiency in Portugal, expressed by:
f p  15.4  N1 60  20  (15)
ranging from 35º to 41º, with an average of 38º. The relation between qc from CPT and s’v0
(expressed in equation 13) is presented in Figure 16. CPT results reveal a moderate increase
of qc in depth. Robertson and Campanella’s proposal tends to lead to higher values of f’,
especially at lower depths, than those obtained from triaxial tests, since the effective cohesive
component is not considered.
15.0
CPT1 f'=48 f'=46 f'=44 f'=42 f'=40
CPT2
12.5
CPT3
CPT4
10.0 f'=38
CPT6
qc (MPa)

7.5
f'=36

5.0 f'=34

f'=32

2.5 f'=30

0.0

0 25 50 75 100 125 150 175 200


s'v0 (kPa)

Figure 16. qc vs s’v0, and the angle of shearing resistance, f’ – adaptation of Robertson and
Campanella (1983) to CEFEUP results (Viana da Fonseca et al, 1998).

This reflects the simultaneous sensitivity of qc towards frictional and cohesive components.
In the present case, the CPT results are rather constant in depth, crossing a wide range of
friction angles (35-42º) with higher incidence at 37º, which is much lower than the one obtained
in the laboratory tests. Being attributed to the behaviour of a cohesive-frictional soil, where the
lower confinement levels are dominated by the cohesive component, while the higher are
mostly governed by friction (Viana da Fonseca and Coutinho, 2008). Theoretically, peak and
post peak resistance can be obtained by pressuremeter tests (PMT/SBPT). Due to the
influence of disturbance during installation, peak strength is unreliable from PMT. For drained
conditions the angle of shearing resistance can be estimated by the approach proposed by
Hughes et al. (1977), assuming the material as a linear Mohr-Coulomb plastic and zero
cohesion intercept, with a constant dilation angle (y, for deformations up to around 10%. The
pressure-expansion data, when plotted as log p (cavity pressure) versus log e (cavity strain),
tends to be linear, with a slope, S. This value of S may be used to determine the values of f′
and y by the following expressions, considering where f′cv as the “constant volume” friction
angle, by the following equations:
S
sin f  (16)
1  S  1sin fcv

sin y  S  S  1sin fcv (17)


Residual soils are, however, characterized by a cohesive-frictional strength, for what any of
these procedures cannot be used by themselves for strength evaluation purposes (Viana da

49
ISP7-PRESSIO 2015. Frikha, Varaksin & Gambin (Eds.) © 2015

Fonseca and Coutinho, 2008). The usual way of interpreting pressuremeter tests by plotting
data in ways that a single parameter can be extracted from a specific part of the experimental
pressure-expansion curve is not feasible under these conditions (such as the undrained
strength in clay and friction angle in sand). An obvious alternative is to reproduce the whole
pressuremeter curve, varying the constitutive parameters in order to obtain a best fit of the
experimental data as advocated by many authors (Jefferies, 1988; Fahey & Carter, 1993;
Schnaid et al. 2000). Since different sets of parameters can produce an equally good fit to
measured data, engineering judgement is always required and independent data from
laboratory tests is generally desirable. The fitting process is part of the definition (checking) of
initially estimated parameters.
A cavity expansion model that incorporates the effects of structure and structure
degradation into cylindrical cavity expansion theory was introduced by Mántaras & Schnaid
(2002). The Euler Method is applied to solve simultaneously two differential equations that lead
to the continuous variations of strains, stresses and volume changes produced by cavity
expansion. Despite the mathematical complexity, an explicit expression for the pressure-
expansion relation is derived without any restriction imposed on the magnitude of deformations.
The propose approach consists of a unified solution for interpretation of pressuremeter tests. A
case study presented by Schnaid & Mántaras (2004) illustrated the applicability of the proposed
approach.
Schnaid & Coutinho (2005) provides an overall picture of the state-of-practice for
pressuremeter testing in Brazil. Pressuremeters are complementary to exploratory techniques
such as SPT and CPT tests and are primarily used in complex geological formations,
particularly in research projects. The extension of the cavity expansion theory to accommodate
de framework of unsaturated soli behavior in the interpretation of pressuremeter tests is being
developed, although the authors recognize the need for further research efforts in assessing
soil parameters in unusual conditions. The extension of the cavity expansion theory to
accommodate the framework of unsaturated soil behaviour in the interpretation of
pressuremeter tests is discussed by Schnaid & Coutinho (2005). Fontaine et al. (2005) present
experimental results and the analytical interpretation of several Menard pressuremeter tests on
an unsaturated soil matrix. The authors deal specifically with the curve fitting methodology to
derive the parameters of the PMT test.
Other approaches have been proposed, starting from a set of results derived from
laboratory tests and cross-hole seismic tests, and applying a “distorted” hyperbolic model,
incorporated into numerical programs, such as CAMFE (Fahey & Carter, 1993), aiming to find
the set of parameters that may give the best fit to the observed results. A large number of
variables may be involved, making the finding of a unique set of parameters to match each
curve quite complex. This will be referred below.

Cohesive-frictional shear strength derived from multi-size loading tests

A simple approach, although limited to shallow foundations, is the recourse to multi-size loading
tests (PLT, for instances). Punching type failures in these residual soils, give very little definition
of an inflexion zone of the pressure-settlement curve, but log-log plotting can be a better way to
detect it (Viana da Fonseca and Siva Cardoso, 1999).
This has been used preliminarily to interpret the results of loading tests (Viana da Fonseca,
1996) and the values for the failure load were obtained: qf =700kPa (B=0.30m); qf =821kPa
(B=0.60m); qf =950kPa (B=1.20m). Failure load is associated to a “punching” mechanism,
which can be distinct from the ultimate equilibrium defined by limit state analysis (Terzaghi-
Meyerhof-Vésic). If we use the values of those failure loads (Viana da Fonseca et al. 1997b), in
the bearing capacity formulation, taking account of the water level position, three equations are
obtained. These can be optimised to get the two strength parameters range. The derived
values were: c' = 6.3–6.9kPa and f' = 36.5-37.0, revealing a fair agreement with the results

50
ISP7-PRESSIO 2015. Frikha, Varaksin & Gambin (Eds.) © 2015

obtained in extensive laboratory testing of undisturbed samples (Viana da Fonseca, 1998) and
in situ testing (Viana da Fonseca et al. 1998).

3.2. Self-boring pressuremeter tests in underground stations in Metro do Porto


Description of the equipment and procedure

The equipment used was a Cambridge Selfboring Pressuremeter from the National Laboratory
of Civil Engineering (LNEC), in Lisbon, with six extensometer arms and maximum load capacity
of 10 MPa. Considering soil’s composition, the hard soil/soft rock configuration was adopted for
these tests. Using this configuration, the equipment has a rubber membrane, thicker than the
usual adiprene one, and reinforced at the ends. The membrane’s outer diameter was 87 mm
and the rock-roller bit was 73 mm diameter. Both membrane clamping rings and beveled tip
were 87.7 mm diameter, meaning that there is a gap of 0.7 mm between the membrane and
the cavity wall, which corresponds to a radial extension of around 0.8 %. By selfboring srilling
technique, t the tools operate by a static force and an inside rock roller bit, ensuring minimum
disturbance of the surrounding soil. Figure 17 shows aspects of the tests.

Drilling tube

Rubber
membrane
Rock-roller bit

a) b)

c) d) e)

Figure 17. Aspects of SBPT tests at the experimental site: a) main parts of the equipment;
b) rock roller bit; c) pressuremeter placed on the drilling device; d) pre-drilling before the
selboring process; e) overall view of the tests

51
ISP7-PRESSIO 2015. Frikha, Varaksin & Gambin (Eds.) © 2015

81.15C

39.44

TBM tunnel

C’
Limits of excavation
Depth - 22m
Fill Material Residual Soil (NSPT<20) Residual Soil (NSPT>20)

P - PRESSUREMETER CH - CROSS HOLE S - BOREHOLE BLOCK SAMPLES Granite W4 Granite W3

Figure 18. Metro station site plan and in situ tests in the vicinity of SBP tests, transverse (C-C’)
geotechnical section.

The tested residual soil and the project for what it was characterized

The SBP tests were performed in the Granite Residual Soil from Porto and this was part of an
extensive ground characterization from laboratory and in situ tests results conducted in a
specific site (see Figure 18) for modelling the behaviour of a deep cut and cover station
included in the new project of the Metro of Porto (Topa Gomes et al., 2008a, Topa Gomes,
2009, Viana da Fonseca e Topa Gomes, 2010). In this project some deep shafts were
executed by the Sequential Excavation Method in the vertical direction, which can be
considered a very innovative solution (Topa Gomes 2008a). The excavation is performed
excavating rings with heights less than 2 m and applying immediately shotcrete to support the
sequential phased excavation. Thicknesses for the support are reduced, varying from 0.30 m to
0.60 m details of the construction process on several steps for the excavation of these shafts
can be seen in Topa Gomes (2009) and Viana da Fonseca and Topa Gomes (2010). The
solution proved to be very effective and the economic gain of the solution, together with the fast
rhythm of the construction, is high when compared to a classical solution, formed by diaphragm
walls or piles. The maximum settlements around the excavation can reach significant values
that can be unacceptable when architectural and structural integrity has to be preserved, but
when the construction is in the middle of a square and adjacent buildings were more than 30 m
away from the limits of the excavation, the movements can induce small damage and the
excavation is to be completed with success.

In Salgueiros station the Sequential Excavation Method in the vertical direction was adopted
intersecting two ellipses, as illustrated in Figure 19.a presenting a picture of the station when
the excavation reached the final depth, while in Figure 19.b plan view and a longitudinal section
of the station is represented, with reference to the main dimensions.

52
ISP7-PRESSIO 2015. Frikha, Varaksin & Gambin (Eds.) © 2015

79.95
WP 15 WP 16 WP 2 WP 3

WP 14
WP 1
WP 4

0.60
R1
.65
WP 5

1.60

38.27

25.40
Primary lining Final lining
WP 13 WP 6
Well Points

WP 7
WP 12

WP 11 WP 10 WP 9 WP 8

0.30

6.60

2.00
0.45

3.60

28.09
0.60 0.60

12.02

2.45
Figure 19. a) Salgueiros station in the final phase of the excavation; b) Geometry of Salgueiros
station with reference to the main dimensions (Topa Gomes et al. 2008b).

A detailed description of this station may be found in Topa Gomes (2009). It must be
referred that the excavation was performed lowering previously the water table, which created a
decisive non saturated condition, increasing significantly both the stiffness and strength of the
residual soils involved by the excavation. The profile was mostly constituted by a saprolitic
material with less than 10% clay, around 20% silt and almost 70% sand. A small percentage of
gravel also appears (Topa Gomes et al. 2008b).

This material preserves the natural fabric of the original rock, having a natural cementation
that confers a cohesion that in certain cases can reach values up to 50 kPa (Viana da Fonseca,
1999, Topa Gomes, 2009). Figure 20 presents the grain size distribution of the soil where the
SBP tests were performed. The soil are almost non-plastic (Ip are very low), giving a
considerable amount of data that was explored for a detailed characterisation of the material.
Figure 18 has shown a plan view of the site, with the locations of these tests. Figure 21
presents SPT results from Borehole SE5, the closest to pressuremeter borehole P1 and
borehole SC1, the closest to pressuremeter borehole P2. In the same figure, values of G0
obtained from cross-hole tests performed close to SBP tests are also presented. It should be
noted that P2 is located around 15 m away from the closest in situ test, and thus, bearing in
mind the sudden geotechnical variations of granite residual soil, the above-presented results
may not be so directly related to the SBP tests in P2 as they are to those in P1.

53
ISP7-PRESSIO 2015. Frikha, Varaksin & Gambin (Eds.) © 2015

76.1
9.51

38.1
19.1
4.76
2.00

25.4

50.8
ASTM Opening (mm)
Sieve Sieve number

200

140

80

60

40

20

10

4
100 0

90 10

80 20

70 30

% RETAINED
60 40
% PASSED

50 50

40 60

30 70

20 80

10 90

0 100
0.001 0.01 0.1 1 10 100

SILT SAND GRAVEL


CLAY
FINE MEDIUM COARSE FINE MEDIUM COARSE FINE MEDIUM COARSE

Figure 20. Granulometric curve and pictures taken from the material washed on the selboring
excavation and recoiled one of the boreholes of the residual soil where the pressuremeter tests
were performed.
G 0 (MPa)
0 500 1000 1500
0
CH1+CH2 CH3+CH4
SPT-SE5 SPT - SC1

5
Depth (m)

10

(2051)
15
0 25 50 75
NSPT
Figure 21. SPT N-values and G0 values from cross-hole seismic tests, obtained from tests
close to SBP locations (Topa Gomes et al. 2008b).

Laboratory Tests

In the vicinity of the SBP tests, some good-quality block samples were collected (Figure 22.a)
in order to perform laboratory tests. These laboratory tests included direct shear tests on
remoulded samples, in order to obtain an approximate value for f′cv, the friction angle at
constant volume, and triaxial tests (Figure 22.b).

54
ISP7-PRESSIO 2015. Frikha, Varaksin & Gambin (Eds.) © 2015

Figure 22. a) Recoiling undisturbed block samples; b) triaxial test with high-precision
instrumentation.

The direct shear tests on remolded samples indicated values of f′cv of 36º under vertical
effective stresses of 100 kPa and 200 kPa, after several forward and backward cycles of
displacement. The conventional interpretation for a direct shear test is likely to slightly
overestimate (by the order of 5%) the direct shear angle of friction (Jewell, 1989). Bearing this
in mind, and also that reference values for f′cv in the material here studied tend to be slightly
lower (Viana da Fonseca, 2003), a value of f′cv of equal to 34º was assumed for the material.
In the five triaxial tests, local strain measurements over the central sections of the samples
were used, allowing accurate measurement of the deformation parameters. Table II presents
the main information on these tests. Figure 23 presents a plot of the results of the triaxial tests
at failure and the derived resistance parameters. It must be said that the obtained values may
underestimate the in situ resistance as most of the samples failed along highly-weathered relic
joints in the material. As the dimension of the ground involved in the test increases, these
singularities tend not to be continuous and, hence, it is the matrix resistance that determines
the failure stresses. In such conditions, values of friction angle up to 45º could be expected.

Table II. Triaxial tests performed 250

s'v0 s'h 0 s'vf s'hf


Test 200
(kPa) (kPa) (kPa) (kPa)
SalgB1-C1 50.0 30.0 146.4 30.1
150
t (kPa)

SalgB1-C2 100.0 60.0 382.7 59.9


SalgB1-C3 200.0 120.0 523.0 119.9 f′ = 39º; c′ = 9 kPa
100
SalgB1-CL1 50.0 30.0 50.5 5.6
SalgB1-CL2 100.0 60.0 99.7 5.7
50

0
0 100 200 300 400
s ′ (kPa)

Figure 23. Triaxial tests results and


derivation of resistance parameters

SBP tests results


As already presented, two boreholes (P1 and P2 in Figure 18) were selected to perform the
SBP tests. Table III presents the depths and vertical stresses at which the several tests were

55
ISP7-PRESSIO 2015. Frikha, Varaksin & Gambin (Eds.) © 2015

performed. Five SBP were performed in P1 and another six tests were performed in P2 (Topa
Gomes et al. 2008b). Figure 24 presents a detail of the tri-cone rock-roller bit used for the
drilling operation with the SBP. This type of bit ensures that good drilling progress can be
made in the types of soils present on this site, but it can lead to “over-drilling”, as discussed
later, which affects the quality of the resulting pressure-expansion data obtained from the test.

Table III. Location and details of the SBP tests.

Test Depth sv0 s'v0 u0


(m) (kPa) (kPa) (kPa)
P1-T1 3.85 69.3 69.3
P1-T2 4.90 88.2 88.2
P1-T3 7.50 135.0 135.0
P1-T4 8.60 154.8 148.9 5.89
P1-T5 9.70 174.6 157.9 16.68
P2-T1 4.60 82.8 82.8
P2-T2 6.60 118.8 118.8
P2-T3 7.70 138.6 138.6
P2-T4 11.10 199.8 169.4 30.41
P2-T5 12.20 219.6 178.4 41.20
P2-T6 16.60 298.8 214.4 84.37
Figure 24. Detail of the tri-cone
rock-roller bit in the SBP.

The results of one typical SBP test from borehole P1 and one from borehole P2 are
presented in Figures 25.a and 25.b, respectively, as plots of pressure in the instrument (‘cavity
pressure’), versus the expansion of the pressuremeter expressed as “cavity strain”, e:

e = r/ro (18)

where r is the increase in radius of the pressuremeter, and ro is the initial (uninflated)
value.

Observing the curves of the various tests (as typified by the plots in Figure 25), it is clear
that there are distinctly different types of materials in the two boreholes: while the material in P1
is a residual material with a weathering degree W5 (confirmed by the boreholes in the vicinity),
the material in P2 is a much more resistant and stiffer material, indicating a weathering degree
W4 and, hence, a rock-like material. For example looking at the pressure-expansion curves for
tests P1-T3 and P2-T3 (Figure 25.a and b, respectively), performed at similar depths, the first
reaches a maximum pressure of around 1100 kPa while the second reaches a value of almost
3000 kPa at similar cavity strains.

It is also clear from these test curves that the self-boring process in these tests was far from
being perfect; with the test curves showing evidence of significant over-drilling (i.e. the borehole
formed by the self-boring process was somewhat larger in diameter than the body of the
instrument). As a result, the first part of the expansion phase shows no resistance, as the
membrane expands to make contact with the over-sized hole (Topa Gomes et al. 2008b). This
has severe implications for the determination of K0 from the tests, as discussed later.

56
ISP7-PRESSIO 2015. Frikha, Varaksin & Gambin (Eds.) © 2015

1200 3000

1000 2500

Cavity pressure (kPa)


Cavity pressure (kPa)

800 2000

600 1500

400 1000

200 500

0 0
0 2 4 6 8 10 12 14 0 2 4 6 8 10 12 14
Cavity strain e (%) Cavity strain e (%)

Figure 25: Curve pressure versus extension for tests: a) P1-T3 and b) P2-T3.

Determination of the soil parameters from the SBP curves

In order to interpret the tests, a first approach used was that proposed by Hughes et al. (1977),
which assumes the material has a linear Mohr-Coulomb plastic behaviour and zero cohesion
intercept, with a constant dilation angle (y , for deformations up to around 10%. The pressure-
expansion data, when plotted as log p (cavity pressure) versus log e (cavity strain), tends to be
linear, with a slope, S. This value of S may be used to determine the values of f′ and y by the
expressions (16) and (17). Fahey & Randolph (1984) carried out SBP tests in sand in which an
oversized-hole was created deliberately in one set of tests. They showed that the test could
still be interpreted to give reasonable values of f′ and y using the Hughes et al. (1977)
approach, provided the origin for strain was adjusted appropriately. The procedure is described
in detail in Topa Gomes (2009). As an example, with the test curve shown in Figure 25.a, the
cavity strain was corrected by about 1.5% to take account of the over-drilling, while a correction
of about 3% was required for the curve shown in Figure 25.b.

Using this approach, values of f′ and y were derived assuming a f′cv of 34º, as previously
justified. Table IV presents the values obtained for the tests.

Table IV. Values of f′ and y obtained using the Hughes et al. (1977) procedure.
Test S f′ (º) y (º)
P1-T1 0.63 52.7 25.1
P1-T2 0.51 44.5 13.4
P1-T3 0.56 48.1 18.5
P1-T4 0.54 46.6 16.3
P1-T5 0.60 50.8 22.4
P2-T1 0.44 39.5 6.9
P2-T2 0.51 44.4 13.3
P2-T3 0.51 44.8 13.9
P2-T4 0.41 38.1 5.0
P2-T5 0.44 39.7 7.1
P2-T6 0.52 45.5 14.8

The values obtained appear to be very high, indicating that cohesion may have some
importance in this material (Viana da Fonseca, 2003). This aspect is due both to the fact that
the natural cementation of the material gives a certain cohesion, and to the non-saturated
conditions of some of the tests, as they were performed above the water table. Another

57
ISP7-PRESSIO 2015. Frikha, Varaksin & Gambin (Eds.) © 2015

relevant aspect is related to the fact that tests on borehole 2 seem to indicate a lower value of
the friction angle, which is to some extent contradictory, with the much higher resistance of the
soil in this borehole. This clearly indicates the importance of the cohesion on the SBP curves
for borehole P2 as is it almost a rock like material, as previously discussed.
The values obtained very much corroborate the expression proposed by Bolton (1986)
relating the peak friction angle with the dilation angle:
fmax  fcv  0.8y (19)
A very important parameter to determine from the SBP tests is the value of K0 as it is almost
the only test that allows its direct determination, at least in theory. However, because of the
over-drilling issue mentioned previously, the usual assumption made in deriving K0 from SBP
“lift-off” pressure – that the insertion process has maintained the in situ sh on the body of the
instrument – is no longer valid. However, interpretation of tests such as these to give in situ sh
(and hence K0) is problematic, and approaches used in interpreting standard Ménard tests
have to be used instead of the “lift-off” approach.

As presented by Topa Gomes et al. (2008b), two procedures were adopted to estimate the
in situ horizontal stress (sh0): the first was the traditional procedure, that assumed that the “lift-
off” pressure is equal to sh0; the second, used as consequence of the over excavation,
interprets the test as if it were a the Ménard pressuremeter test, and assumes that sho
corresponds to the intersection of the initial linear elastic part of the curve with the initial
expansion part, as illustrated in Figure 26.
Table V presents the K0 values for the tests obtained using the procedures above
described, where the superscript refers to the relevant method. Although some discrepancy
between the derived values exists, K0 tends to be between 0.55 and 0.70. These values are
slightly higher than normally assumed for design in the Porto granite residual soils, although
corroborate that low values are incident in highly weathered horizons and tend to increase for
less weathered classes (Viana da Fonseca and Sousa, 2001). Bearing in mind that, due to the
low degree of weathering (W4) of this particular material, values of K0 reveal that trend to
become higher, and this may explain this discrepancy. However, it should also be pointed out
that there are no reliable alternative measurements of K0 in the W4 Porto granite soil. Some
values were excluded either because they were not possible to determine or because the
procedure resulted in completely unrealistic values.

Table V. K0 values from the SBP tests.


650

550
C sh01 K01 sh02 K02
Cavity pressure (kPa)

Test
450 (kPa) (kPa)
P1-T1 41.0 0.59 50.0 0.72
350 P1-T2 37.5 0.43 150.0
P1-T3 82.2 0.61 90.0 0.67
250 P1-T4 90.0 0.56 110.0 0.70
P1-T5 105.0 0.56 107.0 0.57
150 B Deduced s ho P2-T1 12.0 13.0
A P2-T2 78.0 0.66 80.0 0.67
P2-T3 72.0 0.52 74.0 0.53
50
P2-T4 120.0 0.53
0 1 2 3 4 5
P2-T5 102.0 0.34
Cavity strain e (%)
P2-T6 160.2 0.35 182.0 0.46
Average 0.53 0.58

Figure 26. Interpretation of the pressure-


expansion curve to get sh0: first approach

58
ISP7-PRESSIO 2015. Frikha, Varaksin & Gambin (Eds.) © 2015

For tests above the water table, the values of K0 were also determined on the assumption
that the water pressure was hydrostatic (negative) above the water table, all the way to the
ground surface. Table VI presents the values of K0 obtained on this basis; these values are, of
course, higher than the previous values, since the latter were all less than 1.0. Particularly near
to surface, values are exceptionally high and this may result from an over-estimation of the
suction, or rather from an over-estimation of the contribution to effective stress of suction in
non-saturated soil close to the surface. Near the surface, evaporation may change the water
content equilibrium and hence suction levels tend to be lower.

Table VI. K0 values assuming hydrostatic (negative) pore pressures above the water table.

Test K01 K02

P1-T1 1.18 1.31


P1-T2 0.77 2.05
P1-T3 0.65 0.70
P2-T1 0.55 0.56
P2-T2 0.77 0.79
P2-T3 0.54 0.56
Average 0.66 0.78

An alternative procedure for determining sh0 from Ménard pressuremeter test results in stiff
clay was proposed by Marsland and Randolph (1977). This is based on the assumption that if
the drilling process results in an unloading of the wall of the borehole to the extent that inward
plastic movement occurs, the initial non-linear part of the subsequent pressure-expansion curve
(e.g. A-B in Figure 26) corresponds to an initial plastic reloading phase, and this is followed by
a linear phase (B-C), which corresponds to the elastic reloading. For a Tresca-type material
(i.e. elastic perfectly plastic material, with shear strength equal to su), this latter phase should
start at the point (point B) where the pressure on the borehole wall is equal to sh0–su and
should end when the pressure is sho+su (point C), so that the cavity pressure corresponds to
sho somewhere within the linear elastic part of the pressure-expansion curve. Similar reasoning
could be applied to these tests in a Mohr-Coulomb material, which would give a value of sh0 of
350 kPa for the test shown in Figure 26. This would imply a Ko value of well over 2, which is
unrealistically high for this material, on the basis of local experience. Thus, it appears that the
Marsland-Randolph procedure cannot be used to interpret these tests.

Description of the model and model parameters

The SBP tests were modelled numerically (Topa Gomes et al. 2008b) using a model based on
a “distorted” hyperbolic stress-strain relationship, incorporated into the finite element program
CAMFE by Fahey & Carter (1993). The model has as input parameters: c′, cohesion of the
material; f′, friction angle; y, dilation angle;  the Poisson’s ratio; C and n, parameters that
relate the initial tangent stiffness (G0) to the effective stress level by Equation 20; f and g,
parameters that control the non linearity of the stress-strain behaviour up to peak strength by
Equation 21; and , a parameter that modifies the value of G0 depending on the mean shear
stress at the point of load reversal (Yu and Richart, 1984). The main relationships are:
n
G0  p' 
 C   (20)
pa  pa 
g
G   
 1 f   (21)
G0   max 

where pa is a reference pressure (the atmospheric pressure equal to 100 kPa).

59
ISP7-PRESSIO 2015. Frikha, Varaksin & Gambin (Eds.) © 2015

Therefore, the model has 8 parameters, and in many cases, a good fit to the pressuremeter
curve can be obtained with different combinations of these parameters – i.e. the fitting does not
yield a unique set of parameters. However, some constraint is provided to many of the
parameters by using results from other tests – laboratory tests on adjacent undisturbed
samples to provide values for strength parameters, and in situ seismic tests to provide values
for G0.

Bearing in mind the low values of c′, and the low contribution of c′ to overall resistance in
SBP tests taken to high cavity pressures, c′ was assumed to be zero. This is illustrated in
Figure 27, which shows three SBP curves generated by the CAMFE model. Initially taking f′ =
45° and y = 13°, the effect of changing c′ from 0 to 20 kPa is shown to be only slight. The
figure also shows that the same effect on the curve can be obtained by increasing f′ to 45.6°
and y to 13.8°. Thus, neglecting the low values of c′ in the fitting process might result in the
best-fit values of f′ being higher by no more than 1° or 2°.

2000
c' = 20 kPa; f ' = 45°, y = 13°

1500
Cavity pressure (kPa)

f ' = 45.6°, y = 13.8°

1000 f ' = 45°, y = 13°

500

s ' h0 = 100 kPa


0
0 5 10 15
Cavity strain e (%)

Figure 27. CAMFE output, showing effect of c′ on the pressure-expansion curve.

The dilation angle y was assumed to be related to the friction angle by Bolton’s relationship
- Equation (19) - assuming f′cv of 34º. Poisson’s ratio  was assumed to be 0.2. From Equation
(21), it can be seen that the secant stiffness at peak strength (when f = fmax) is equal to 1–f,
and therefore f can be derived from good-quality triaxial tests in which internal instrumentation
is used to determine Go. In this case, 5 such tests were performed, giving f values between
0.84 and 0.96, with an average value of 0.88. Given the small range of values, the mean value
(0.88) was assumed for all cases. The value of parameter  was considered to be 0.3, as the
best fit lead, in all the tests, to values close to this.

On this basis, there remain four independent parameters to vary: f′, C, n and g (though the
values of C and n were also constrained to some extent, as the value adopted in each case,
when inserted into Equation (20), had to result in a sensible value of G0 when compared to the
G0 values from nearby cross-hole seismic tests). These parameters were adjusted to obtain
the best possible fit to the two “unload-reload” shear modulus values (Gur) obtained from the
test, and the overall shape of the curve between the first unload-reload loop and the maximum
expansion. Because of the over-drilling mentioned earlier, the fit to the initial section of the
curve was always imperfect (Topa Gomes et al. 2008b). Due to some creep effects observed at
the start of the unload-reload cycles in the tests, the value of Gur was defined between the
vertices of the cycle. Table VI presents the parameters obtained after the back analysis, for the
tests in borehole P1.

60
ISP7-PRESSIO 2015. Frikha, Varaksin & Gambin (Eds.) © 2015

As observed in Table VII, the differences between the parameters obtained are very
significant. Regarding the parameters controlling stiffness, though these show a wide range,
they are in accordance with the results of the cross-hole tests. Resistant (strength) parameters
appear to be very different, and clearly distinct from the values obtained in the lab on good
quality block samples. Most of values are out of the normal values for this type of soils. This
phenomenon was even more pronounced in borehole 2, where the deduced values of the
resistant parameters were completely outside (higher than) the range considered “normal” for
this type of soil in Porto. This seemed to indicate that one of the fixed parameters was not well
adopted. As a consequence, a second round of calculations was performed changing the value
of K0 in order to achieve the best fit between the real test and the numerical model.
Table VII. Parameters obtained from back-analysis for SBP tests in borehole 1, taking c′ = 0.
c′ G0*
Test f'(º) y(º) g 
(kPa) C n (MPa)
P1-T1 0 38 5 1050 0.2 0.3 0.3 94.6
P1-T2 0 41 8.8 1220 0.2 0.3 0.3 111.8
P1-T3 0 32 0 1125 0.2 0.3 0.3 107.6
P1-T4 0 45 25 1400 0.6 0.2 0.3 144.7
P1-T5 0 46 15 2200 0.2 0.3 0.3 226.7
* values derived from C and n

Table VIII. Parameters obtained from back-analysis for the SBP tests in both boreholes,
allowing c′ to be non-zero, and varying K0
c′ f' y G0*
Test g 
(kPa) (º) (º) C n K0 MPa
P1-T1 10 39 5 1050 0.2 0.3 0.3 0.57 91.3
P1-T2 10 39 10 1220 0.2 0.3 0.3 0.74 111.8
P1-T3 0 32 0 1125 0.2 0.3 0.3 0.60 107.6
P1-T4 50 43 15 1275 0.6 0.2 0.3 0.73 131.8
P1-T5 50 43 10 2000 0.2 0.3 0.3 0.70 206.0
P2-T1 1000 45 15 3200 0.50 0.3 0.3 0.80 270.4
P2-T2 1000 45 15 2750 0.30 0.3 0.3 1.40 267.8
P2-T3 1000 45 15 4800 0.20 0.3 0.3 1.20 474.3
P2-T4 1000 45 15 3900 0.15 0.3 0.3 1.10 398.9
P2-T5 1000 45 15 2600 0.35 0.3 0.3 1.30 345.8
P2-T6 1000 45 15 5500 0.25 0.3 0.3 1.33 691.2
* values derived from C and n

Referring to the resistant parameters, the results obtained in triaxial tests were used as the
most probable starting point for fitting the results from both boreholes, but assuming that they
could vary within a range also defined from the results of the triaxial tests. However, for tests on
borehole 2, which show a much higher resistance, these values were clearly inadequate and
much higher resistant parameters had to be used. Thus, values of c′ = 1 MPa and f′ = 45° were
adopted for the W4 weathered granite. Then, the parameters C and n were adjusted to obtain
the best fit to the Gur values and the overall test curve. Table VIII presents the parameters
obtained via this process, including the new values of K0.

Figure 28 presents a picture of the fit obtained on test P1-T4 and Figure 29 a similar picture
for test P2-T4. The results in Table VII show the clearly different characteristics of the materials
in the two boreholes. Material on P2 is much more resistant and stiffer, appearing to be in the
transition between a soil-like material to a rock-like material (Topa Gomes, 2009).

61
ISP7-PRESSIO 2015. Frikha, Varaksin & Gambin (Eds.) © 2015

1400 3000
P1-T4 test
1200 Numerical model 2500

Cavity pressure (kPa)


Cavity pressure (kPa)

1000
2000
800
1500
600
1000
400
P2-T3 test
200 500
Numerical model

0 0
0 2 4 6 8 10 0 2 4 6 8 10
Cavity strain e (%) Cavity strain e (%)

Figure 28. Curve pressure versus extension Figure 29. Curve pressure versus extension for
for P1-T4 test superposed with the numerical P2-T3 test superposed with the numerical back-
back-analysis. analysis.

Almost all of the back-analysis showed quite a good agreement between the modelling and
the test (if the initial section is ignored, as in Figure 28), with the exception of P2-T1, where
results were poor, at least compared with the other tests.

The change in the K0 values allowed a better fit of the tests and the variation on this
parameter is quite reduced. For tests on borehole 1, K0 has an average value of 0.67 ranging
from 0.60 to 0.74. This value does not differ very much from the value obtained from the lift-off
of the membrane, even bearing in mind the amount of over-drilling involved. However, referring
to borehole 2, the values appear to be much higher, with an average value of 1.19. The
variation is also higher, ranging from 0.80 to 1.40. If test P2-T1 is excluded, due to poor overall
fit, the average K0 value obtained is 1.27, with the range being 1.10 to 1.40. The comparison
between the two boreholes suggests a decrease in the K0 value as weathering increases, as
pointed out by Viana da Fonseca & Sousa (2001). In addition, the conclusion of a K0 value
greater than 1.0 for the less-weathered granite soils in Porto is novel, and contrary to previous
assumptions. Obviously, these preliminary results should be confirmed with more carefully
tests. Referring to the stiffness parameters, the values obtained from the fitting process show
quite good agreement with the measured field values (Figure 21), although the nearest field
values are from a location more than 15 m from borehole 2.

Another interesting aspect to observe is that the deduced Gur/G0 ratio ranges from 0.40 to
0.59 with a mean value of 0.50, without significant differences between the results from the two
boreholes. Fahey et al. (2003) suggested that Gur can often be used as a “working stiffness” for
design of large spread footings on stiff sandy soils. This leads to a very tentative suggestion
that an appropriate “working stiffness” value for these soils could be taken to be about 0.4 to
0.5 times the Go value. However, this suggestion should be treated with caution, and only
applied to “well-designed foundations” (Burland, 1989), where strains are small, and not
necessarily to other situations (retaining walls, tunnels, etc), where local strains might be
greater (Topa Gomes et al. 2008b).

62
ISP7-PRESSIO 2015. Frikha, Varaksin & Gambin (Eds.) © 2015

4. Collapsibility: identification and classification of collapsible residual soils

The most qualitative criterion for determining the susceptibility to collapse are based on
relationships between the porosity, void ratio, water content, or in place dry density (Viana da
Fonseca and Coutinho, 2008). Examples of these criteria applied to in situ tests (SPT-T) were
proposed by Décourt and Quaresma Filho (1994). Quantitative criterion have been developed
from results of simple and double oedometric tests in order to classify soils’ susceptibility by a
collapse potential (CP), which is the deformation due wetting obtained through simple
oedometric tests according to the expression:
Δe
CP  100 (22)
1  eo
where: e and eo is void ratio variation caused by wetting and eo the initial void index.

Although SPT is a test broadly utilized world-wide, to the authors’ knowledge, no


identification criterion based on the results from SPT test has been published. However, in
Brazil, porous collapsible soils in southern and central-west regions have been characterized
for showing low values in N (in general N < 5). A typical example is showed in Figure 29a, by
Marques et al. (2004) concerning the porous clay layer in Brasília, Brazil, where the penetration
index down to the 7 m in depth fit in this limit. Suctions in these regions are low (< 50 kPa) and
it may account for the small variations in the results when the test is carried out in the dry or
humid season. In Petrolândia City, in Pernambuco, Northeastern region of Brazil, suctions may
reach values such as 10 MPa in the dry season, influencing significantly the test results.
Results from SPT in the dry season in the collapsible layer (9 < N < 20) are about 100% greater
than those obtained in the humid season (4 < N < 13), shown in Figure 29b, reflect clearly this
aspect (Coutinho et al. 2004a and Souza Neto et al. 2005). Kratz de Oliveira et al. (1999)
present a proposal of collapsible soil identification from results of double pressuremeter tests,
consisting of a comparison between a test in the natural condition with another in a borehole
previously flooded, similar to double oedometer test considered by Jennings and Knight (1957).
The pressuremeter collapse is defined by:

2 2 2 2
r f  ri r wet  ro nat
C press  2
 o 2
(23)
ri ro nat

being ri and rf both are the radii of the cavity for the ground under condition of natural water
content and for the wetting soil for the level of the creep pressure PF of the wetting PMT tests;
and ronat and rowet are the initial radii of the cavity for soil conditions of natural water content and
soaked, respectively. Figure 30.a presents double PMT tests obtained in a collapsible soil in
Northeast of Brazil. The results show a great loss of soil stiffness (Dourado & Coutinho, 2007).
Another example of reduction in pressuremeter stiffness response, after soaking the area, is
shown by Schnaid et al. (2004b).

Dourado & Coutinho (2007) utilize this procedure and compares the results of the Cpress (Eq.
2) with results obtained by Souza Neto et al. (2005) for the same local through simple
oedometer tests for a soaking pressure of 200kPa using values of collapse potential (CP) (Eq.
1) and the classification of the soil collapsibility proposed by Jennings & Knigth (1975). These
results are presented in Figure 30.b. The values of Cpress vary between 6.7 and 32.4%, with
highest values in the stretch between 1.0 and 1.5m. Similarity of behavior is observed, when
compared with the oedometer tests (CP), characterizing a stretch (between 1.0 and 2.0m) more
susceptible to the collapse. Considering the coherence observed in the results showed in the
Figure 29.b and the results obtained by Kratz de Oliveira et al. (1999), Dourado & Coutinho
(2007) presented a proposal for classification of the soil collapsibility from PMT tests showed in
Figure 31.

63
ISP7-PRESSIO 2015. Frikha, Varaksin & Gambin (Eds.) © 2015

800

600

Volume (cm )
3
400
Depth = 2 m

200
PMT - Natural W.C.
(a) PMT - Wetting
0
0 500 1000 1500 2000 2500
Pressure (kPa)
(a)

Depth Cpress Classification of


CP
soil collapsibility
(m) (%) (%) (Jennings & Knight, 1975)
0.75 6.7 4.8 Moderately problematic
1.00 11.2 5.2 Problematic
1.50 32.4 6.2 Problematic
2.00 8.7 3.3 Moderately problematic
2.60-2.90 9.5 1.5 Moderately problematic

(b)
Figure 30. a) Pressuremeter tests in a collapsible
soil - Petrolandia, PE; b) Cpress, CP (p/
(b) ssoak.=200kPa), classification of collapsibility
(Dourado & Coutinho, 2007).

Figure 29. a) SPT results from a collapsible soil in Brasilia (Marques et al. 2004);
b) SPT borehole and water content profile in Petrolandia, PE (Coutinho et al. 2004b
and Souza Neto et al. 2005).

Ferreira and Lacerda (1995) and Houston et al. (1995) presented two similar – although
developed independently - for performing in situ collapse tests. The former authors reported
several tests, with the so-called “Expansocollapsometer” (ECT), performed inside an open
auger borehole using a rigid circular plate of small diameter (100mm). Souza Neto et al. (2005)
utilized an improved version of this equipment. The in situ tests, basically, consist in the loading
of a plate inside an open auger borehole up to a determined vertical stress. After stabilization of
settlements, the soil is soaked, the process being monitored.

Settlements due to soaking (collapse) are measured until their stabilization. Collapse strain
is determined by dividing the collapse settlement for the wetting front depth obtained at the end
of the test. Ferreira and Lacerda (1995) show that the results from field tests can be correlated
to single oedometric tests, thus allowing the criteria of identification and classification (e.g.
Jennings and Knight, 1975), based on oedometric tests, to be extended to field tests.
Considering the results of UFPE experimental site (Petrolândia-PE), Dourado & Countinho
(2007) present and discuss the most soil collapse criterion (laboratory and in situ tests).

Figure 32 presents results obtained by Souza Neto et al. (2005) at the same area in
Northeast of Brazil, performing plate load test (Figure 32.a) and the ECT test (Figure 32.b).
After wetting the collapse settlement measured in the plate load test was 45.0mm. For the ECT
test the collapse settlements increase with the soaking vertical stress, as expected, and in the
same vertical stress of 100 kPa the settlement predicted (53mm) by the collapse strain
determined was of the order of 18% higher than the collapse settlement measured (plate load

64
ISP7-PRESSIO 2015. Frikha, Varaksin & Gambin (Eds.) © 2015

test). This result can be considered to be situated in the error range generally obtained by other
design methods. Complementary studies involving numerical modeling associated to more data
from in situ tests will allow better evaluation of the in situ collapse behavior and, consequently,
a more rational interpretation of the in situ collapse tests - ECT.

0 1 2 0 1 2
0 0 0 0
Damages Pressuremeter
Collapse
(Kratz de Oliveira
0,5 0,5 svi = 200kPa 0,5 0,5 et al. 1999)
I I

1 1 1 1
........
MODERATELY PROBLEMATIC

PROBLEMATIC
WITHOUT PROBLEMS

..

..

MODERATELY PROBLEMATIC

PROBLEMATIC
..
WITHOUT PROBLEM
.

.
Depth (m)

1,5 1,5 1,5 1,5

II
2 2 2 II 2

2,5 2,5 2,5 2,5


Classification:
Classificação:
SPT> 50 Jennings ee Knight
Jennings Knight (1975)
(1975)
(a) SPT> 50 (b) (a)
(a) 3 3 3 3
0 2 4 6 8 10 0 10 20 30 40

ec (%) Cpress (%)

Range of values Classification proposed

For Cpress(%) < 5 => Soil without problems

5  Cpress(%)  10 => Soil with moderate problems

For Cpress(%) > 10 => Problematic soil


(b)

Figure 31. a) Classification of the soil


collapsibility according Jennings & Knigth (1975)
and variation of the Cpress versus depth – (b)
Petrolandia - PE; b) proposal of classification of
the soil collapsibility from PMT results (Dourado
& Coutinho, 2007).
Figure 32. a) In situ collapse test with plate load tests (PLT);
b) Typical result of the in situ collapse test with
“expansocollapsometer” (ECT) in Petrolandia, PE (Souza Neto
et al. 2005)

Dourado & Coutinho (2007) calculated the collapse settlement using the PMT results
considering the traditional methodology of Briaud (1992) for footing settlement on sand. The
settlements were calculated in both natural water content and wetting conditions and their
difference in these two conditions was considered as collapse settlement (rc) of the soil. It was
observed that for any stress that the PMT underestimatet the rc measured in the plate load test.
The use of a factor of increase (Fm) of 2.5 in rc predicted (Fm . rc pred = rc med) by the PMT,
would lead to one better approach of rc measured for the in situ collapse tests. The authors
emphasized that it must be seen as a suggestion, having to be applied only to the soil studied.

65
ISP7-PRESSIO 2015. Frikha, Varaksin & Gambin (Eds.) © 2015

5. Conclusions

Residual soils are a product of rock mass weathering and complex diagenesis that generated
materials dominated by strong inhomogeneity, fabric and structure (macro and micro). The
aspects for classification of these soils may be enumerated as: morphological and
mineralogical characteristics, physical identification testing and engineering properties.
Structure, specially the effect of interparticle bonding, has a significant influence in the
interpretation of test data, especially from in situ tests. These singularities were addressed in
these lines, focusing on the advantages of using pressuremeter tests, giving relevance to the
consequences in very objective parameters, useful to geotechnical design, such as: (i) at rest
stress state,; (ii) stiffness - from the very small strain levels (so important for dynamic analysis,
as for competent definition of stress-strain constitutive laws) to serviceability levels (useful for
common projects, such as foundations); (iii) non-linearity of stiffness (cyclic and monotonic) or
degradation laws (iv) strength, where the triode “cohesive intercept (‘effective cohesion’) –
dilation – and friction (‘angle of shearing resistance’)” rules the behaviour in failure…; (v)
special features, such as permeability and collapsibility and few lines on the very important
subject of non-saturated conditions.

The special features that demand for unusual or non-classical approaches in the
characterization of residual soils have been recognized to be the key point for consequent
analysis and subsequent remarks. This text is limited to authors’ experience and to what they
have managed to recoil. Residual soils are variable in their origins (parent rocks, climate
conditions and else) and too much complex in their physical and stress state, to be dealt in
such a limited space. Nevertheless, some specific characteristics were identified for such non-
textbook approach, which is translated in singular non-linear stress-strain behaviours, and
unsaturated conditions – which implications with the mechanical behaviour of these soils are
strongly interdependent with the former interparticle structure.

Some specificities were addressed giving rise to these considerations:

(i) The behavior of deep foundations is influenced by many factors, which are difficult to
characterize and there is still, at present, considerable uncertainty regarding pile design criteria,
particularly in residual soils. Thus, it is important to conduct load tests on instrumented piles to
enable the observation and modeling of its behavior. For that purpose, an experimental test site
in residual soil from granite, which included an extensive geotechnical site characterization,
was developed in the campus of the Faculty of Engineering of the University of Porto (FEUP).
Several piles were executed by means of different construction techniques, which were then
tested under vertical and horizontal loading. In this paper PMT tests based predictions for the
ultimate load capacity of the three classes of piles constructed in saprolitic (residual) soils from
Porto granite were evaluated, giving a very good insight to their potentialities (Viana da
Fonseca and Santos, 2008).

The predictions that have used the PMT in test data for the bored piles, agree very well with
the results from Static Pile Load Test (SLPT), independently of the method or the PMT
profile(s) that have been used. The exception is the method proposed in Bustamante &
Gianeselli (1993) expressed in the French code (MELT, 1993) that gives values out of the
range referred. Since three diverse applications of the same method, developed by
Bustamante and co-authors in LCPC, give such dispersion, a distinct selection of the input data
may explain it, as logical explanation. While, the predictions that have used PMT test results for
the ultimate resistance for CFA piles, clearly overestimated the SLPT observed values. Only
the method proposed by Bustamante & Gianeselli (1982) gives values in the range, the ones
that have used PMT results tests data for the driven pile, only 3 of the predictions are out of the
defined range, but two were very near to the lower border and the rest fitted very well.

66
ISP7-PRESSIO 2015. Frikha, Varaksin & Gambin (Eds.) © 2015

Generally, the PMT based predictions do not agree very well with the QSPLT(s/b=10%) but are
better than the ones that used different in situ tests. This is not strange to the fact that the pile
head load-relative settlement curves obtained in the static load tests converge with the pattern
obtained in studies in centrifuge models of displacement and non-displacement piles in
granular soils. For large strains, the behaviour of the piles in this residual soils seems to be
close to the one of granular soils deposits.

(ii) Analyses of the laterally load-tested piles in Porto granite residual soils recurring to PMT
semi-empirical p-y curves, originally proposed by Ménard, generally, underestimated horizontal
displacements in residual soils, mostly in high loading levels, which can be corrected by
decreasing soil strength in very small depths, that is,in shallow horizons. A proposal outlined In
the text, designated as “modified”, overestimated displacements for very small loading levels,
consequence of the excessive low value considered for the reaction modulus (K). It can be
admitted that this method leads to a reasonable simulation of the behavior of bored and CFA
piles. For driven piles this method is still unreliable, by the fact that an exponential law is
considered for the variation of pu with depth, which is, as referred, inadequate to the behavior
of this type of piles. As a general conclusion it can be point out that these methods based on
PMT results can capture reasonably the behavior of the test piles, especially when some
improvements are introduced to take into account the particularities of these residual soils.

(iii) Results of SBP tests performed on a granite residual soil in Porto, and their
interpretation, clearly evidenced the over-drilling in the insertion process, which made difficult to
interpret the tests, especially with regard to the K0 value. However, starting from a set of
results derived from laboratory tests and cross-hole seismic tests, a “distorted” hyperbolic
model, incorporated into the program CAMFE (Fahey & Carter, 1993), was used to model the
tests, aiming to find the set of parameters that gave the best fit to the observed results. A large
number of variables was involved, which meant that finding a unique set of parameters to
match each curve was quite complex. However, a pattern was achieved and an overall
tendency of the results was found, which allowed values for the parameters to be deriving with
a certain degree of confidence.

(iv) A good overall fit to the SBP curves could only be obtained in some cases by adopting
much higher values of K0 than have previously been assumed for moderately-weathered to
highly-weathered Porto granite soils. It is clear from these tests that, because of the over-
drilling issue, the method of determining Ko by considering the overall fit to the pressure-
expansion curve may be a better approach than by focussing on the initial part of the
expansion. This is similar to the conclusion reached by Jefferies & Shuttle (1995) for
“imperfect” tests in clay soils where insertion disturbance makes the method based on “lift-off”
pressure unreliable.

(v) Identification, classification and evaluation of soil collapsibility by Ménard pressuremeter


is a promising approach, being necessary to enlarge the volumetric range of the equipment to
turn it a method for the explicit estimation of a quantitative index, like the collapse potential.

6. Acknowledgments

This research was developed under the activities of FCT (Portuguese Foundation for Science
and Technology) research unit CONSTRUCT and the Group of Geotechnical Engineering of
Slopes and Landscapes (GEGEP) of the Federal University of Pernambuco, in Recife, Brazil,
and this paper was developed under the project “Study of the Geotechnical Soil Behaviour of
Different Geological Units of the Metropolitan Region of Recife/Pernambuco for Risk Analyses
of Natural Disasters and Conception of Protection Solutions”, financed by Brazilian

67
ISP7-PRESSIO 2015. Frikha, Varaksin & Gambin (Eds.) © 2015

NationaLCouncil of Science and Technology (CNPq) under the Programme Science without
Borders (Grant PVE, Process N. 313999/2014-7).

7. References

Babenderde, S., Hoek, E., Marinos, P. & Cardoso, A.S. 2004. Characterization of Granite and the
Underground Construction in Metro do Porto, Portugal. In: Viana da Fonseca, A. & Mayne, P.W. (eds.)
Geotechnical & Geophysical Site Characterization. Rotterdam, Millpress. 1, 39-47.
Bolton, M.D. 1986. The strength and dilatancy of sands. Géotechnique, 36(1), 65–78.
Briaud, J.-L. 1992. The Pressuremeter. Trans Tech Publications. A.A. Balkema, Rotterdam, 322p
Burland, J.B. 1989. Ninth Laurits Bjerrum memorial lecture: Small is beautiful - the stiffness of soils at small strains.
Canadian Geotechnical Journal, 26(4), 499–516.
Bustamante, M. & Gianeselli, L. 1982. Pile bearing capacity predictions by means of static penetrometer CPT. In:
Proceedings of the 2nd European Symposium on Penetration Testing, ESOPT II, 24-27 May 1982, Amsterdam.
Rotterdam, A. A. Balkema. 2, 493 - 500.
Bustamante, M., Gianeselli, L. 1993. Design of auger displacement piles from in situ tests, 2nd Intern. Geotech.
Seminar: Deep Foundations on Bored and Auger Piles, Balkema.
Coutinho, R. Q., Souza Neto, J. B., Barros, M. L. S., Lima, E. S. & Carvalho, H. A. 1998. Geotechnical
characterization of a young residual soil/gneissic rock of a slope in Pernambuco, Brazil. 2nd International
Symposium on The Geotechnics of Hard Soils and Soft Rocks, Vol.1, Naples, Italy, 115-126
Coutinho, R.Q., Souza Neto, J.B., Costa, F.Q. 2000. Design Strength Parameters of a Slope on Unsaturated
Gneissic Residual Soil. Advances in Unsaturated Geotechnics, ASCE – Geotechnical Special Publication Nº 99,
pp.247-261.
Coutinho, R. Q.; Souza Neto, J. B.; Dourado, K.C.A 2004a. General report: Characterization of non-textbook
geomaterials. In: Viana da Fonseca, A. & Mayne, P.W. (eds.) Geotechnical & Geophysical Site Characterization.
Rotterdam, Millpress. 2,1233-1257.
Coutinho. R.Q., Dourado, K.C.A. & Souza Neto, J.B. 2004b. Evaluation of the collapsibility of a sand by Ménard
pressuremeter. In: Viana da Fonseca, A. & Mayne, P.W. (eds.) Geotechnical & Geophysical Site
Characterization. Rotterdam, Millpress. 2, 1267-1273.
Cruz, N. 2010. Modelling geomechanics of residual soils by DMT tests. PhD thesis, University of Porto, Portugal.
Décourt, L. 1989. The standard penetration test. State of the Art Report. Proc. XII ICSMFE, Rio de Janeiro.
Balkema, Rotterdam. Vol. 4: 2405-2416.
Décourt. L. & Quaresma Filho, A.R. 1994. Practical appli-cations of the standard penetration complemented by
torque measurements, SPT-T; present stage and future trends. Proc. XIII ICSMFE, New Delhi, India, Vol.1,
pp.143-146.
Dourado, K.C.A. & Coutinho, R.Q. 2007. Identification, classification and evaluation of soil collapsibility by Ménard
pressuremeter. XIII Panamerican Conference on Soil Mechanic and Foundation Engineering. Isla de Margarita,
Venezuela, pp. 724-730.
Fahey, M, and Carter, J.P. 1993. A finite element study of the pressuremeter test in sand using a non-linear elastic
plastic model. Canadian Geotechnical Journal, 30(2), 348–362.
Fahey, M. and Randolph, M.F. 1984. Effect of disturbance on parameters derived from self-boring pressuremeter
tests in sand. Géotechnique 34(1), 81–97.
Fahey, M., Lehane, B. and Stewart, D.P. 2003. Soil stiffness for shallow foundation design in the Perth CBD.
Australian Geomechanics, 38(3), 61–89.
Ferreira, C., Viana da Fonseca, A. & Nash, D. 2011. Shear wave velocities for sample quality assessment on a
residual soil. Soils and Foundations. Special Issue on Deformation Characteristics of Geomaterials, IS-Seoul
2011: 51 (4): 683-692
Ferreira, S. R. M. & Lacerda, W. A. 1995. Volume Change Measurements in Collapsible Soil by Laboratory and
Field Tests. Unsaturated Soils, Alonso & Delage eds., v. 2, pp.847-854.
Fleming, W.G.K., Weltman, A.J., Randolph, M.F., and Elson, W.K. 1985. Piling Engineering, Surrey University
Press, Wiley & Sons, New York, 380 p.
Fontaine, E., Cunha, R.P. and David, C. 2005. A simplified analytical manner to obtain soil parameters from
Menard pressuremeter tests on unsaturated soils. Fifth Int. Symp. on Pressuremeters, Paris, ISP-5.
Frank R. 2008. Design of foundations in France with the use of Ménard pessuremeter tests. Invited lecture,
Comptes rendus Jornada El ensayo presiométrico en el proyecto geotécnico, 24 June 2008, Madrid, CEDEX-
UPC, pp. 1- 16.
Frank, R. & Zhao, S.R. 1982. Estimation par les paramètres pressiométriques de l'enfoncement sous charge axiale
de pieux forés dans des sols fins. Bulletin Liaison Laboratoire Central des Ponts et Chaussées. 119, 17-24. (in
French).
Frank, R. 1999. Design of Shallow and Deep Foundations. Paris, Presses de l’Ecole Nationale des Ponts et
Chaussées et Techniques de l’Ingénieur, ENPC, (in French).
Gambin, M. and Frank, R. 2009. Direct Design Rules for Piles Using Ménard Pressuremeter, Proc. IFCEE
‘09, ASCE.

68
ISP7-PRESSIO 2015. Frikha, Varaksin & Gambin (Eds.) © 2015

Gambin, M. 1963. The Ménard Pressuremeter and the Design of Foundations. Actes Journées des Fondations,
Laboratoire Central des Ponts et Chaussées, Paris. (in French)
Gomes Correia, A., Viana da Fonseca, A. & Gambin, M. 2004. Routine and advanced analysis of mechanical in-situ
tests. Keynote lecture. In: Viana da Fonseca, A. & Mayne, P.W. (eds.) Geotechnical & Geophysical Site
Characterization. Rotterdam, Millpress. 1. 75-95.
Hatanaka, M., & Uchida, A. 1996. Empirical Correlation Between Penetration Resistance and Internal Friction Angle
of Sandy Soils. Soils and Foundations, Vol. 36, No. 4, pp. 1-9.
Houston, S.L., Mahmoud, H.H. & Houston, W.N. 1995. Down-hole collapse test system. Journal of Geotechnical
Engineering, ASCE, Vol.121, No4. pp. 341-349.
Hughes, J.M.O, Wroth, C.P and Windle, D.W. 1977. Pressuremeter tests in sands. Géotechnique, 27(4), 455–477.
International Journal for Numerical Methods in Engineering, 69(6), 1158-1193.
Jefferies, M. G. 1988. Determination of horizontal geostatic stress in clay with self-bored pressuremeter. Can.
Geotech. J., 25, pp. 559-573.
Jefferies. M.G. and Shuttle, D.A. 1995. Disturbance does not prevent obtaining reliable parameters from SBP tests
in clay. Proc. Int. Symp. on the Pressuremeter and Its New Avenues, Sherbrook, Canada, 177–183.
Jennings, J.E. & Knight, K.1975. A Guide to construction on or with materials exhibiting additional settlement due to
a collapse of grain structure. 4th Confer. for Africa on Soil Mech. and Found. Engineer. Durban, pp. 99-105.
Jewell, R.A. 1989. Direct shear tests on sand. Géotechnique, 39(2), 309–322.
Jiang, M. J., Yu, H. S., and Leroueil, S. 2007. A simple and efficient approach to capturing
Kratz de Oliveira, L.A.; Schnaid, F. & Gehling, W.Y.Y. 1999. Use of Pressuremeter Tests in the Prediction of
Collapse Potential Soils. Soils and Rocks Journal, São Paulo - Brazil, v. 22, Nº 3, pp.143-165.
Kulhawy, F.H., & Mayne, P.W. 1990. Manual on Estimating Soil Properties for Foundation Design. Report EL-6800,
Electric Power Research Institute, Palo Alta, CA, August, 306 pp.
Leroueil, S & Vaughan, P.R. 1990. The general and congruent effects of structure in natural clays and weak rocks.
Géotechnique, Vol 40, pp. 467- 488.
Machado, S. L. & Vilar, O. M. 2003. Geotechnical characteristics of an unsaturated soil deposit at São Carlos,
Brazil. Characterisation and Engineering Properties of Natural Soils – Tan et al. (eds.), Swets & Zeitlinger, Lisse,
ISBN 90 5809 537 1
Mantaras, F.M. & Schnaid, F. 2002. Cavity expansion in dilatant cohesive-frictional soils. Géotechnique, 52 (5),
337-348.
Marsland, A. and Randolph, M.F.R. 1977. Comparisons of the results from pressuremeter tests and large in situ
plate tests in London clay. Géotechnique, 27(2), 217–243.
Mayne, P.W. & Schneider, J.A. 2001. Evaluating axial drilled shaft response by seismic cone. Foundations &
Ground Improvement. Geot. Sp. Pub. GSP 113, ASCE, Reston/VA, 655–669.
Mayne, P.W. 2001. Stress-strain-strength-flow parameters from enhanced in-situ tests. Proc. Intl. Conf. on In-Situ
Measurement of Soil Properties & Case Histories, Bali, Indonesia, 27–48. Parayangan Catholic University.
MELT 1993. Design Rules for Foundations, Tender Documents for Public Works, Fasc. No.62, Titre V (in French),
Imprimerie Nationale Paris.182 pages.
Ménard L. 1963. Estimation of load bearing capacity of foundations from pressuremeter tests’ results. Sols Soils 5,
9-32. (in French)
Ménard, L., Bourdon, G. & Gambin, M. 1969. General method for design of raft and isolated pile horizontally loaded
using pressuremeter results. Sols Soils No 22/23 (6), 16-29. (In French)
Pinyol, N., Vaunat, J. & Alonso, E. E. 2007. A constitutive model for soft clayey rocks that includes weathering
effects. . Géotechnique 57 (2), 137–151.
Poulos, H.G. & Davis, E.H. 1980. Pile Foundation Analysis and Design. New York, Wiley & Sons.
Poulos, H.G. (1989) Pile behavior: theory and application, 29th Rankine Lecture, Geotechnique, 39 (3), 363-416.
Randolph, M.F. & Wroth, C.P. (1978) Analysis of deformation of vertically loaded piles. ASCE Journal of the
Geotechnical Engineering Division, 104 (GT12), 1465-1488.
Randolph, M.F. & Wroth, C.P. 1979. A simple approach to pile design and the evaluation of pile tests. Behavior of
Deep Foundations, ASTM STP 670, 484-499.
Robertson, P.K. & Campanella, R.G. 1983. Interpretation of cone penetration tests. Part I: Sand; Part II: Clay.
Canadian Geot. J., Vol. 20, Nº 4: 718-745.
Schnaid, F. & Coutinho, R.Q. 2005. Pressuremeter Tests in Brazil (National Report). In: International Symposium
50 Years of Pressuremeters, (2) 305-318.
Schnaid, F. 2005. Geo-characterisation and properties of natural soils by in situ tests. Keynote Lecture. 16th
ICSMGE, Osaka, (1), 3-45. Millpress, Rotterdam.
Schnaid, F., & Mántaras, F.M. 2004. Interpretation of pressuremeter tests in a gneiss residual soil from São Paulo,
Brazil. In: Viana da Fonseca, A. & Mayne, P.W. (eds.) Geotechnical & Geophysical Site Characterization.
Rotterdam, Millpress. 1, 49-74.
Schnaid, F., Fahey, M. & Lehane, B. 2004a. In situ test characterisation of unusual geomaterial. Keynote Lecture.
In: Viana da Fonseca, A. & Mayne, P.W. (eds.) Geotechnical & Geophysical Site Characterization. Rotterdam,
Millpress. 1, 49-74.
Schnaid, F.; Kratz de Oliveira, L.A. and Gehling, W.Y.Y. 2004b. Unsaturated constitutive surfaces from
pressuremeter tests. J. Geotech. and Geoenvironmental Engng, ASCE, 130 (2), 174-185.

69
ISP7-PRESSIO 2015. Frikha, Varaksin & Gambin (Eds.) © 2015

Schnaid, F.; Ortigão, J.R.; Mántaras, F.M.; Cunha, R.P. & McGregor, I. 2000. Analysis of self-boring pressuremeter
(SBPM) and Marchetti dilatometer (DMT) tests in granite saprolites. Can Geotech. J. 37,1-15.
Souza Neto, J.B.; Coutinho, R.Q; Lacerda, W.A 2005. Evaluation of the Collapsibility of a Sandy by In Situ Collapse
Test. In: 16th ICSMGE, Osaka, (2). 735-738.
Topa Gomes, A. 2009. Elliptical Shafts by the Sequential Method of Excavation on Vertical Direction. The Case of
Metro do Porto. PhD Thesis, Faculdade de Engenharia, University do Porto. (In Portuguese)
Topa Gomes, A, A. S. Cardoso, J. Almeida e Sousa, J. C. Andrade & C. Campanhã 2008a. Design and behavior of
Salgueiros Station for Porto Metro. Proc. of the 6th Int. Conf. on Case Histories in Geotechnical Engineering, 11-
16 August, 2008 2008. Arlington, VA, Vol. 1. 8.04C.
Topa Gomes, A., Viana da Fonseca, A. & Fahey, M. 2008b. Self-boring pressuremeter tests in Porto residual soil:
results and numerical modelling. In: Geotechnical and Geophysical Site Characterisation, Huang, A–B & Mayne,
P.(eds). Taylor & Francis, 1193-1021.
Tuna, C., 2006. Tests and Analysis of the Behaviour of Laterally Loaded Piles in Residual Soils from Granite. MSc
Thesis. FEUP, Univ. of Porto (in Portuguese)
Tuna,C., Viana da Fonseca, A. & Santos, J. A 2008. Data interpretation and analysis of the behavior of laterally
loaded piles in ISC’2 experimental site by recourse of PMT and DMT based methods. In: Geotechnical and
Geophysical Site Characterisation, Huang, A–B & Mayne, P.(eds). Taylor & Francis, 639-645.
Viana da Fonseca, A. & Santos, J. 2008. International Prediction Event. Behaviour of Bored, CFA and Driven Piles
in Residual Soil. ISC’2 experimental site. Book edited by the University of Porto (FEUP) & Technical University
of Lisbon (IST-UTL) www.fe.up.pt/sgwww/labgeo/pdf/Book-IppE-Piles-ISC2.pdf
Viana da Fonseca, A. 2003. Characterizing and deriving engineering properties of a saprolitic soil from granite, in
Porto. Characterization and Engineering Properties of Natural Soils. Eds. Tan et al., 1341–1378. Swets &
Zeitlinger, Lisse.
Viana da Fonseca, A. 1988. Characterisation of Residual Soils from Granite in the Region of Porto. MSc Thesis,
University of Porto. (in Portuguese)
Viana da Fonseca, A. 1996. Geomechanics in Residual Soils from Porto Granite. Criteria for the Design of Shallow
Foundations. PhD Thesis, University of Porto. (in Portuguese)
Viana da Fonseca, A. 1998. Identifying the reserve of strength and stiffness characteristics due to cemented
structure of a saprolitic soil from granite. Proc. 2nd International Symposium on Hard Soils – Soft Rocks.
Naples. Vol.1: pp.361-372. Balkema, Rotterdam.
Viana da Fonseca, A. 1999. Surface loading tests for mechanical characterisation of a saprolitic soil from granite of
Porto. Proceedings 9th Panamerican Conference on Soil Mechanics and Geotechnical Engineering, Foz do
Iguassu, Brasil. 1, 403-409.
Viana da Fonseca, A. 2003. Characterizing and deriving engineering properties of a saprolitic soil from granite, in
Porto. Characterization and Engineering Properties of Natural Soils. Eds. Tan et al. Vol.2, pp.1341-1378. Swets
& Zeitlinger, Lisse.
Viana da Fonseca, A. and Silva Cardoso, A. 1998. Linearly increasing elastic analysis of surface loading tests on a
saprolitic soil from granite. Proc. XI Panamerican Conference on Soil Mechanics and Geotechnical – Foz de
Iguassu, Brasil, Vol.3, 1527-1535.
Viana da Fonseca, A. and Almeida e Sousa, J. 2001, At-rest coefficient of earth pressure in saprolitic soils from
granite. Proc. XIV ICSMFE, Istanbul, Vol. 1, 397–400.
Viana da Fonseca, A. and Coutinho, R. Q. 2008. Characterization of residual soils, Keynote Lecture. In:
Geotechnical and Geophysical Site Characterisation, Huang, A–B & Mayne, P.(eds). Taylor & Francis, 195–248.
Viana da Fonseca, A. and Topa Gomes, A. 2010. Project and construction of Underground stations and tunnels
(TBM and NATM) in heterogeneous masses for Metro do Porto. In: Excavations and Tunnels in Granite
(TÚNELES Y EXCAVACIONES EN GRANITO). Edited by Aula PAYMACotas - Ingeniería de Túneles, da
Universitat Politècnicade Catalunya. 79-123.
Viana da Fonseca, A., Buttling, S. and Coutinho, R. Q. 2012. Foundations: Shallow and deep foundations,
unsaturated conditions, heave and collapse, monitoring and proof testing. Pgs. 283-412, in A Handbook of
Tropical Residual Soil Engineering. Ed.s Bujang B. K. Huat, David G. Toll & Arun Prasad. 548 pp., A.A. Balkema
book, CRC Press, Taylor & Francis Group. (ISBN 978 0 415 45731 6).
Viana da Fonseca, A., Carvalho, J., Ferreira, C., Santos, J. A., Almeida, F., Pereira, E., Feliciano, J., Grade, J. &
Oliveira, A., 2006. Characterization of a profile of residual soil from granite combining geological, geophysical,
and mechanical testing techniques. Geotechnical and Geological Engineering, Springer, Netherlands. Vol. 14
(5), 1307-1348.
Viana da Fonseca, A., Carvalho, J., Ferreira, C., Tuna, C., Costa, E. & Santos, J. 2004. Geotechnical
characterization of a residual soil profile: the ISC’2 experimental site, Porto. Geotechnical & Geophysical Site
Characterizaton. In: Viana da Fonseca, A. & Mayne, P.W. (eds.) Geotechnical & Geophysical Site
Characterization. Rotterdam, Millpress.2, 1361-1370.
Viana da Fonseca, A., Coop, M.T., Fahey, M. and Consoli, N. 2011. The interpretation of conventional and non-
conventional laboratory tests for challenging geotechnical problems. Keynote Lecture IS’Seoul, in ‘Deformation
Characteristics of Geomaterials’. IOS Press, Amstredam. 1: 84-119.
Viana da Fonseca, A., Ferreira, C. & Carvalho, J. 2004b. Tentative evaluation of K0 from shear waves velocities
determined in Down-Hole (Vsvh) and Cross-Hole (Vshv) tests on a residual soil. In: Viana da Fonseca, A. &
Mayne, P.W. (eds.) Geotechnical & Geophysical Site Characterization. Rotterdam, Millpress. 2, 1755-1764.

70
ISP7-PRESSIO 2015. Frikha, Varaksin & Gambin (Eds.) © 2015

Viana da Fonseca, A., Matos Fernandes, M. & Cardoso, A. S. 1998a. Characterization of a saprolitic soil from Porto
granite by in situ testing, First Int. Conf. on Site Characterization –ISC’98. Atlanta, Vol.2, pp.1381-1388.
Balkema, Rotterdam.
Yu, H. S., Tan, S. M. and Schnaid, F. 2007. A critical state framework for modelling bonded geomaterials.
Geomechanics and Geoengineering, 2(1), 61-74.
Yu, P. and Richart, F.E. 1984. Stress ratio effects on shear modulus of dry sands. Journal of Geotechnical
Engineering, ASCE, 110(3), 331–345.

71
ISP7-PRESSIO 2015. Frikha, Varaksin & Gambin (Eds.) © 2015

72

Vous aimerez peut-être aussi