Vous êtes sur la page 1sur 12

LeBlanc, C., Houlsby, G. T. & Byrne, B. W. (2010). Géotechnique 60, No. 2, 79–90 [doi: 10.1680/geot.7.

00196]

Response of stiff piles in sand to long-term cyclic lateral loading


C . L E B L A N C  , G . T. H O U L S B Y † a n d B. W. B Y R N E †

The driven monopile is currently the preferred founda- Le monopieu battu est actuellement le type de fondation
tion type for most offshore wind farms. While the static préféré pour la plupart des parcs éoliens en mer. Bien
capacity of the monopile is important, a safe design que la capacité statique du monopieu soit importante,
must also address issues of accumulated rotation and une conception sans danger doit également tenir compte
changes in stiffness after long-term cyclic loading. des problèmes de la rotation accumulée et de variations
Design guidance on this issue is limited. To address this, dans la rigidité, à la suite d’efforts cycliques de longue
a series of laboratory tests were conducted where a stiff durée. Les conseils et indications conceptuels relatifs à
pile in drained sand was subjected to between 8000 and cette question étant limités, on a procédé à une série
60 000 cycles of combined moment and horizontal load- d’essais en laboratoire, dans le cadre desquels on a
ing. A typical design for an offshore wind turbine soumis un pieu rigide, enfoncé dans du sable drainé, à
monopile was used as a basis for the study, to ensure un nombre de cycles de moment de charge et de charges
that pile dimensions and loading ranges were realistic. horizontales allant de 8000 à 60 000. On a utilisé, comme
A complete non-dimensional framework for stiff piles in élément de base de cette étude, un modèle typique de
sand is presented, and applied to interpret the test monopieu pour éolienne en mer, afin d’assurer que les
results. The accumulated rotation was found to be dimensions du pieu et les plages de charge étaient
dependent on relative density, and was strongly affected réalistes. Pour interpréter les résultats de cet essai, on
by the characteristics of the applied cyclic load. Parti- présente et on applique un cadre non dimensionnel
cular loading characteristics were found to cause a complet pour pieux rigides dans le sable. La rotation
significant increase in the accumulated rotation. The cumulée, qui s’est avérée tributaire du poids spécifique,
pile stiffness increased with number of cycles, which était affectée dans une grande mesure par les caractéris-
contrasts with the current methodology where static tiques de la charge cyclique appliquée. On s’est aperçu
load–displacement curves are degraded to account for que certaines caractéristiques particulières de la charge
cyclic loading. Methods are presented to predict the engendraient une augmentaiton significative de la rota-
change in stiffness and the accumulated rotation of a tion cumulée. La rigidité du pieu augmentait avec le
stiff pile due to long-term cyclic loading. The use of the nombre de cycles, contrairement à la méthodologie ac-
methods developed is demonstrated for a typical full- tuelle, dans laquelle les courbes charge statique/déplace-
scale monopile. ment sont dégradées pour tenir compte des charges
cycliques. La communication présente des méthodes per-
mettant de prédire les variations de la rigidité et la
rotation cumulée d’un pieu rigide due à des charges
cycliques à long terme. L’application des méthodes dével-
KEYWORDS: laboratory tests; piles; repeated loading; sands; oppées est démontrée pour un monopieu typique gran-
settlement; stiffness deur nature.

INTRODUCTION the proportions of a typical offshore wind turbine on a


Wind power currently offers a very competitive source of monopile foundation.
renewable energy, and therefore the market for onshore and The design of monopiles relies on standards and empirical
offshore wind farms is projected to expand rapidly within data originating from the offshore oil and gas sector. How-
the next decade. There are strong political and industrial ever, the loading of an offshore wind turbine is very differ-
forces, especially in northern Europe, supporting the ent, in both magnitude and character, from that of oil and
development of offshore wind power to reduce reliance on gas installations. It is characteristic for offshore wind tur-
fossil fuels and control greenhouse gas emissions. bines that the substructure will be subjected to strong cyclic
There are several foundation concepts for offshore wind loading, originating from the wind and wave loads. This
farms. The cost-effectiveness of a particular concept depends occurs not only during extreme conditions but also during
to a large extent on the site conditions. Most current normal service conditions. This can lead to accumulated
foundations are ‘monopiles’, which are stiff piles with large rotation of the wind turbine tower, adversely affecting its
diameters, driven 20–30 m into the seabed. Recently in- ultimate strength or fatigue life. The long-term movements
stalled monopiles have diameters in the range of 4–6 m and of the foundation may significantly impact on all parts of
a length/diameter ratio of approximately 5. Fig. 1 illustrates the wind turbine, including the support structure, machine
components and blades. Therefore it is of great importance
to investigate the effects of cyclic loading.
The primary design drivers for offshore wind turbine
Manuscript received 6 November 2007; revised manuscript accepted foundations are deformation and stiffness rather than ulti-
8 April 2009. Published online ahead of print 15 December 2009. mate capacity. Modern offshore wind turbines are designed
Discussion on this paper closes on 1 July 2010, for further details
see p. ii.
as ‘soft-stiff’ structures, meaning that the first natural fre-
 Department of Civil Engineering, Aalborg University, and quency is in the range between the excitation frequency
Department of Offshore Technology, DONG Energy, Copenhagen, bands, 1P and 3P, in order to avoid resonances. 1P and 3P
Denmark. denote the frequency bands of the rotor rotation and the
† Department of Engineering Science, University of Oxford, UK. blade passing, typically in the range of 0.3 Hz and 1 Hz

79

Géotechnique 2010.60:79-90.
80 LEBLANC, HOULSBY AND BYRNE
vibrations, so that the displacements of the pile measured
were related solely to the applied loading conditions.

CURRENT METHODOLOGY
Piles are widely used for various structures, such as
bridges, high-rise structures and offshore oil and gas installa-
tions. The interactions between soil and laterally loaded piles
are typically accounted for by the use of p–y curves,
originally introduced by Reese & Matlock (1956) and
McClelland & Focht (1958). The p–y curves adopt the
H Winkler approach by uncoupling the response of various
layers in the soil, and can therefore easily include effects of
non-linearity, soil layering and other soil properties. A p–y
curve defines the relationship p(y) between the soil resis-
tance p arising from the non-uniform stress field surrounding
the pile mobilised in response to the lateral pile displace-
e ment y, at any point along the pile. The implementation of
p–y curves requires a numerical procedure to solve the
fourth-order differential equation for beam bending with
appropriate boundary conditions
y d4 y
Ep I p  pð yÞ ¼ 0, z 2 ½0; L (1)
dz 4
z in which Ep and Ip denote the elastic modulus and second
moment of area of a pile respectively.
L The p–y curves evolved primarily from research in the oil
and gas industry, as the demand for large pile-supported
offshore structures increased during the 1970s and 1980s.
Research has included testing of full-sized piles in sand
under both static and cyclic loading conditions. An overview
of the important tests and results is given by Reese & Impe
D (2001). The p–y curves for piles in sand described by Reese
et al. (1974) and O’Neill & Murchison (1983) led to
Fig. 1. Typical offshore wind turbine installed on monopile
recommendations in the standards (DNV, 1977; API, 1993)
foundation
for oil and gas installations. In 2004 these recommendations
were adopted in the standard Design of offshore wind turbine
structures (DNV, 2004), which represents the current state of
respectively. Long-term cyclic loading of the foundation is the art for design of monopiles in the offshore wind indus-
likely to change the stiffness of the surrounding soil and try.
therefore the interaction of the foundation and the soil, The method adopted in the standards uses a procedure to
owing to the accumulation of irreversible deformations. Any construct non-linear p–y curves for monopiles in sand sub-
significant change in stiffness may result in interference jected to cyclic loading as a function of the static ultimate
between the first natural frequency and the excitation fre- lateral resistance pu ,
quencies, 1P or 3P, which would be highly problematic. Thus  
Bz
it is important to assess the concepts of stiffness and/or p ¼ Apu tanh (2)
strength changes during long-term cyclic loading. Apu y
The performance of monopiles subjected to long-term
cyclic loading must therefore be addressed to achieve a safe where A ¼ 0.9 for cyclic loading, and B is an adjustment
design of an offshore wind turbine. Although many useful parameter to account for the relative density of the sand.
methods have been proposed to predict the response of piles This method was originally developed by O’Neill &
to lateral cyclic loading, methods predicting the accumulated Murchison (1983), and has some theoretical basis. However,
rotation and resulting stiffness due to long-term cyclic load- it relies to a high degree on empiricism, using data obtained
ing are limited. This paper explores the load–displacement primarily from two full-scale load tests reported by Cox et
behaviour of stiff monopiles in sand subjected to long-term al. (1974). These tests were conducted using two slender
cyclic loading. The objective is to provide information for piles, with diameter 0.61 m and length 21 m. The piles were
the development of a conceptual model capable of predicting subjected to static and cyclic lateral load. To assess the
the response of monopiles to this loading. validity of the method, systematic studies were conducted by
Laboratory tests were conducted to simulate a driven Murchison & O’Neill (1984), which proved the method to
monopile in drained conditions subjected to 8000–60 000 be superior to other methods. However, the validity of the
load cycles of combined moment and horizontal loading. In method relies on very few tests on relatively flexible driven
comparison, a typical offshore wind turbine is designed for a steel piles subjected to cyclic loading.
fatigue load with 107 cycles. The laboratory tests were
carried out on the laboratory floor, with due consideration of
issues of scaling of the results. The main advantage of Shortcomings of current methodology
performing the experiments at 1g was the capability to apply The current design methodology, based on p–y curves,
up to 60 000 load cycles in a realistic time frame while has gained broad recognition, owing to the low failure rate
maintaining high-quality displacement measurements. It was of piles over several decades. However, when applied to
possible to isolate the testing rig from the effect of external offshore wind turbine foundations, the design methodology

Géotechnique 2010.60:79-90.
RESPONSE OF STIFF PILES IN SAND TO LONG-TERM CYCLIC LATERAL LOADING 81
is being used outside its verified range, and several design determined degradation parameter that depends on the in-
issues are not properly taken into account. stallation method, soil density and load characteristics.
First, current standards rely on methods built upon empiri- By investigating a subset of the full-scale tests, Lin &
cal data obtained from long, slender and flexible piles. When Liao (1999) proposed that the accumulated displacement of
scaling to large-diameter piles, a distinction must be made piles can be predicted by
between a pile that behaves in an almost rigid fashion and u N  u0
one that is relatively flexible, since this affects the soil–pile ¼  ln ð N Þ (5)
u0
behaviour (Briaud et al., 1984). A rigid pile rotates without
flexing significantly, and develops a ‘toe kick’ under moment in which u0 and u N denote the pile-head deflection in the
and lateral loading. Criteria for rigid or flexible behaviour first and Nth load cycle respectively, and  is an empirical
have been proposed by various researchers (e.g. Dobry et degradation parameter, similar to Æ, depending on the in-
al., 1982; Budhu & Davies, 1987; Carter & Kulhawy, 1988). stallation method, soil density and load characteristics.
The range of transition from flexible to rigid pile behaviour The methods proposed by Long & Vanneste (1994) and
may, according to Poulos & Hull (1989), be evaluated by Lin & Liao (1999) provide simple means for predicting the
effects of cyclic loading. However, the determination of the
Es L4
4:8 , , 388:6 (3) empirical degradation parameters relies on a small number
Ep I p of tests carried out on long, flexible piles subjected to fewer
than 50 cycles of loading. Further investigations are needed
in which Es denotes the elastic modulus of the soil. to verify the form of the models, and to extend them for use
A typical monopile has a diameter of 4 m, wall thickness of in predicting the long-term behaviour of stiff, driven piles.
0.05–0.07 m and penetration depth of 18 m. According to Other investigations include small-scale tests on stiff piles
equation (3), the transition from rigid to flexible pile be- subjected to 10 000 cycles, as reported by Peng et al.
haviour occurs in the range from Es  14 MPa to (2006). However, only a few tests are reported, and the data
Es  1121 MPa. Thus for most sands encountered the mono- interpretation is limited. A more theoretical approach is
pile behaviour tends toward the rigid case. given by Lesny & Hinz (2007), who attempt to predict
Second, the recommended p–y curves for cyclic loading accumulated displacements using data from cyclic triaxial
are designed primarily for evaluation of the ultimate lateral tests and a finite element model incorporating Miner’s law.
capacity. Important design issues, such as accumulated rota- The method is theoretical, and still requires validation
tion and stiffness changes due to long-term cyclic loading, against experimental data.
are poorly accounted for. Long-term cyclic loading is likely
to densify, or in some circumstances possibly loosen, the
surrounding soil, resulting in changes to the stiffness of the DIMENSIONLESS EQUATIONS FOR SCALING OF
foundation. Additionally, an accumulated rotation during LABORATORY TESTS
the lifetime of an offshore wind turbine is expected, since The basis of this paper is a set of laboratory floor
the cyclic loading often occurs from one direction. The experiments on stiff monopiles in sand. Results from labora-
current design methodology is not capable of predicting tory tests of foundations in sand must be carefully scaled to
either effects of soil densification or long-term movements predict the behaviour of a full-scale structure. As is well
of the monopile. recognised for structures on sand, the loading response is
Finally, the current methodology accounts for cyclic load- governed by the frictional behaviour of the sand, which in
ing in an incomplete manner. Repetitive lateral load tests on turn is governed by the isotropic stress level. In the labora-
two offshore piers in Tampa Bay, reported by Long & tory the isotropic stress level controlling the test behaviour
Vanneste (1994), showed much greater displacements than is low, resulting in higher friction angles but lower shear
predicted using the p–y curves proposed by Reese et al. stiffnesses, in comparison with a full-scale test. These issues
(1974). The reason for this discrepancy, according to Long of scaling can be addressed by choosing appropriate scaling
& Vanneste (1994), is that the cyclic p–y curves do not methods, as presented in the following.
account for such factors as installation method, load charac- To ensure that the peak friction angle in a laboratory test
teristics or number of load cycles. corresponds to the value in a full-scale test, the soil sample
is prepared at a lower relative density. This is straightfor-
ward, but the issues of stiffness are more complex. An
METHODS FOR PREDICTING THE RESPONSE TO attempt to account for the influence of isotropic stress level
LONG-TERM CYCLIC LOADING is made by expressing the shear modulus G as
The inadequacy of the current methodology for predicting  n
G  v9
the cyclic loading response of piles means that new models, ¼ c1 (6)
incorporating factors affecting the cyclic behaviour, must be pa pa
developed. Results from 34 full-scale cyclic lateral load tests
of piles in sand were collected by Long & Vanneste (1994) in which pa is the atmospheric pressure, c1 is a dimension-
to identify the factors affecting the cyclic behaviour. These less constant,  v9 is an appropriate effective vertical stress,
included soil density, pile type, installation method and, most and n is the pressure exponent (Kelly et al., 2006). Evalua-
importantly, the characteristics of the cyclic load. Long & tion of the shear modulus using equation (6) requires the
Vanneste (1994) adopted a method, originally introduced by determination of a representative vertical effective stress,  v9 .
Little & Briaud (1988), to account for cyclic loading. The The vertical effective stress around a pile varies with depth.
method is based on the deterioration of static p–y curves, Thus an appropriate choice is to use the vertical stress at a
which is taken into account by reducing the static soil depth c2 L below the sea-bed, given by
reaction modulus according to
 v9 ¼ c2 Lª9 (7)
RN
¼ N Æ (4) in which ª9 is the effective unit weight and c2 is a
R0 dimensionless constant. The pressure exponent in equation
in which R0 and RN denote the soil reaction modulus on the (6) is reported to vary from 0.435 at very small strains to
first and Nth load cycle respectively, and Æ is an empirically 0.765 at very large strains (Wroth et al., 1979). The shear

Géotechnique 2010.60:79-90.
82 LEBLANC, HOULSBY AND BYRNE
strain range of greatest interest is likely to fall in the range son between tests can be obtained by plotting M ~ against
103 to 104 (Simpson, 2002). In this range n is reported to be Ł~ while e~ and other parameters influencing k~ are kept
around 0.75 for sands (Park & Tatsuoka, 1994; Porovic & constant.
Jardine, 1994). However, a value of 0.5 may capture most of Several other parameters may influence the M ~ – Ł~ rela-
the important features of increased shear stiffness with tionship. These can be understood by investigating the static
pressure (Wroth & Houlsby, 1985). This is confirmed by moment resistance of a monopile. Consider the idealised
Kelly et al. (2006), who successfully adopted a value of 0.5 horizontal stress distribution, in the ultimate lateral limit
to compare the results of laboratory and full-scale tests of state, along a stiff pile in sand, as illustrated in Fig. 2. The
suction caisson foundations. resulting distributed horizontal load along the pile is deter-
For the case of a monopile subjected to a horizontal load mined by KD v9 ¼ KDª9z, in which K is a factor depend-
H and moment M at the sea-bed, the resulting lateral ing on the friction angle (e.g. Broms, 1964) and d is the
displacement u and rotation Ł can be obtained from an depth of the pivot point. The pile toe is assumed to shear
elastic stiffness relation, expressed as at the critical state friction angle cr . Thus the shear
2 3 resistance at the bottom of the pile is governed by the
M    vertical effective force arising from the overburden sand,
6 L7 k1 k2 LŁ (/4)D2 Lª9, plus a contribution arising from the structure,
4 5 ¼ DG (8)
k2 k3 u c3 V, in which V is equal to the gravity force acting on the
H
structure, and c3 is a dimensionless constant between 0 and
in which k1 , k2 and k3 are dimensionless constants. To 1. On the basis of these assumptions, the equations of
obtain the moment–rotation relationship, u can be elimi- horizontal equilibrium and moment equilibrium at the pile
nated to give head are given by
   
"  #  1
GL2 D k 1 k 3  k 22 H þ c3 V þ D2 Lª9 sin cr ¼ d 2  L2 KDª9
M ¼ Ł (9) 4 2
k 3  k 2 ð HL=M Þ (11)
 
The issue of scaling is addressed by incorporating equations  ð 3 3Þ
1 L  2d KDª9
(6) and (7) in equation (9) to obtain a moment–rotation M  c3 V þ D2 Lª9 L sin cr ¼
4 3
relationship given entirely in terms of non-dimensional para-
(12)
meters,
pffiffiffiffiffiffiffiffi  rffiffiffiffiffiffiffiffi respectively. These equations can be combined to eliminate
M c1 ð c2 Þ k 1 k 3  k 22 pa
¼ Ł (10) d and give the interaction equation
DL3 ª9 k 3  k 2 ð HL=M Þ Lª9  2  3
|fflffl{zfflffl} |fflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl{zfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl} |fflfflfflffl{zfflfflfflffl} 1 3 3 M 1 H
~
M k~ Ł~ þ Æ 3 ¼ þ Æ þ 2
2 2 2 L KDª9 2 L KDª9
in which k~ is the non-dimensional stiffness, and M ~ and Ł~ (13)
are the non-dimensional values of moment and rotation  
 2
respectively. The non-dimensional moment/force ratio arising c3 V þ D Lª9 sin cr
in k~ will be denoted e~ ¼ M= HL, called the non-dimensional Æ ¼
4
(14)
load eccentricity. This suggests that a satisfactory compari- L2 KDª9

where Æ is introduced for simplicity. This relation can be


V rearranged to obtain an expression given entirely in non-
dimensional parameters,
M
 3
3 M 1 1 H 2
¼Æ þ 12 þ Æ þ (15)
H K DL3 ª9 2 K L2 Dª9
|fflffl{zfflffl} |fflfflffl{zfflfflffl}
~
M ~
H
 
V  D sincr
Ƽ c3 þ (16)
DL2 ª9 4 |{z}L K
|fflffl{zfflffl}
V~ 1=
KDγz d
introducing the pile aspect ratio  and the non-dimensional
vertical and horizontal loads V~ and H ~ respectively. The non-
dimensional horizontal load H ~ can be replaced by H ~¼M ~ = e~.
Thus it follows from equations (15) and (16) that the static
moment capacity, in terms of M ~ , is uniquely determined by
the non-dimensional parameters V~, e~ and . This suggests
that the non-dimensional moment–rotation relationship in
KDγz
Ld equation (10) could be written as
 
M~ ¼ k~ V~, e~,  Ł~ (17)
è
æ π 2
çc3V  4 D Lγ sin φcr
ç
è
æ Thus a satisfactory comparison of both stiffness and
strength, between laboratory and full-scale tests, is likely to
Fig. 2. Horizontal stress distribution in ultimate limit state for ~ against Ł~, while keeping V~, e~ and
be obtained by plotting M
laterally loaded stiff pile in sand  constant. This scaling law derived for monotonic loading

Géotechnique 2010.60:79-90.
RESPONSE OF STIFF PILES IN SAND TO LONG-TERM CYCLIC LATERAL LOADING 83
is also assumed to cover the cyclic response of stiff piles in The experiments were conducted using unsaturated yellow
sand. The non-dimensional parameters are listed in Table 1. Leighton Buzzard silica sand. The characteristics of the sand
are summarised in Table 2, and further information is given
by Schnaid (1990).
The container for the sand was carefully filled by pouring
EXPERIMENTAL EQUIPMENT sand from a low drop height to achieve a very loose state. A
A simple and efficient mechanical load rig is used to denser state was also obtained using a hammer drill to
apply loads to the pile. The rig was originally developed by vibrate the bottom plate of the sand container.
Rovere (2004) for testing of caisson foundations. The rig The tests were conducted using a stiff copper pile. The
consists of a 550 mm 3 600 mm 3 600 mm container for outer dimensions of the pile are scaled to approximately
sand, a steel frame with pulleys, three weight hangers, and a 1:50, in relation to a typical monopile. The pile properties
lever with a driving motor, as illustrated in Fig. 3. The lever are listed in Table 3. The pile was driven into the sand by
is attached to the steel frame through a pivot, and carries a gentle driving with a plastic hammer from a fixed drop
motor, which rotates a mass m1 to cause cyclic loading. The height. The number of strokes needed to reach the final
motor is a geared single-phase AC motor rotating at a penetration depth varied from 460  20 to 740  30 for
frequency of 0.106 Hz. The pulley ropes are 3 mm low- loose and medium dense sand respectively. The monopile
stretch spectral ropes. was fixed horizontally during installation using side supports.
The load rig is a simple static system. Initially, when Horizontal deflections were measured by two dial gauges.
m1 ¼ m2 ¼ 0, the weight of the mass m3 is chosen to The load rig is illustrated in Fig. 4.
balance any force acting on the lever. Thus, assuming that
Ł  /2, as is the case for minor lever deflections, any
sinusoidal load in the form f (t ) ¼ f0 + fa sin(øt ) can be TEST PROGRAMME
applied to the pile by appropriately choosing m1 ¼ The test programme was designed to investigate the re-
[(l2 /la )fa ]/g and m2 ¼ [(lc /la )fa  f0 ]/g. Since m2 . 0, it sponse of the pile and its dependence on the relative density
follows that la or lc must fulfil the condition la /lc , fa /f0 . The of the sand and the characteristics of the applied cyclic load.
load rig is very stable, and can accurately provide a sinus- The average relative densities were Rd ¼ 4% and
oidal loading for more than 1 000 000 cycles. Rd ¼ 38%, corresponding to a loose and a medium-dense
state respectively. A relationship between effective stress,
relative density and peak friction angle for Yellow Buzzard
Table 1. Non-dimensional parameters Sand is given by Schnaid (1990), and is used to compare
peak friction angles between the laboratory tests and a full-
Moment loading ~ ¼ M scale monopile. For the calculation it is assumed that a
M
L3 Dª9 representative effective stress can be taken at 0.8L beneath
the sea-bed. Fig. 5 illustrates that the peak friction angles
Vertical force V
V~ ¼ used in the laboratory were estimated as 358 and 438, which
L2 Dª9 equate to field conditions of Rd ¼ 8% and Rd ¼ 75%, corre-
H sponding to a loose and a dense state respectively.
Horizontal force ~¼
H
L2 Dª9 The characteristics of the applied cyclic load must be
rffiffiffiffiffiffiffiffi
uniquely defined. In the following, load levels are referred to
Rotation: degrees pa in terms of the applied moment M. The corresponding
Ł~ ¼ Ł
Lª9 horizontal force follows from H ¼ M/e. Two independent
parameters are defined to characterise the applied sinusoidal
Load eccentricity M loading,
e~ ¼
HL
L
Aspect ratio ¼
D
Table 2. Characteristics of yellow Leighton Buzzard Sand
(Schnaid, 1990)

Property Value

Particle sizes, D10, D30, D50, D60 : mm 0.63/0.70/0.80/0.85


Specific gravity, Gs 2.65
Minimum dry unit weight, ªmin : kN/m3 14.65
430 mm
Maximum dry unit weight, ªmax : kN/m3 17.58
Critical angle of friction, cr : degrees 34.3

m2 l2
360 mm
Motor

π
Table 3. Properties of the copper monopile
θ⯝
m3 2
80 mm Property Value

m1
Pile diameter, D: mm 80.0
la Wall thickness: mm 2.0
lc Penetration depth, L: mm 360.0
Load eccentricity, e: mm 430.0
Fig. 3. Mechanical load rig used to investigate response of stiff Pile weight, V: N 35.0
piles to long-term cyclic loading

Géotechnique 2010.60:79-90.
84 LEBLANC, HOULSBY AND BYRNE
a b

Fig. 4. Experimental set-up: (a) mechanical load rig; (b) side supports used during installation; (c) installed pile with two dial gauges
measuring horizontal deflections; (d) driving motor

Laboratory cyclic loading, normalised with respect to the static moment


Full-scale
capacity. It follows that 0 , b , 1. The ratio c 2 [1; 1]
48 quantifies the characteristics of the cyclic load, and takes the
46 Rd  100% value 1 for a static test, 0 for one-way loading, and 1 for
φ: degrees

44 two-way loading. A visual interpretation of the load ratios is


Rd  75%
42
38% 75% given in Fig. 6.
40 Rd  50% Initially, static load tests were performed to determine the
38 static moment capacity in terms of M ~ R , as shown in Fig. 7.
Rd  25%
36 4% 8% From the moment–rotation curves it is not possible to
34 Rd  0% identify a distinct point of failure. Thus failure is defined by
0 50 100 150 Ł~ ¼ 48 ¼ 0.0698 rad. The moment–rotation curves in Fig.: 7
p: kPa
are for convenience fitted by :Ł~  0:038 rad 3 ( M ~ =M~ R )2 33
~
and Ł  0:042 rad 3 ( M ~ =M~ R) 1 92
for Rd ¼ 4% and Rd ¼
Fig. 5. Friction angles of yellow Leighton Buzzard Sand as 38% respectively, valid in the range 0:25 M ~R , M ~ ,
function of effective isotropic stress p9 and relative density Rd 0:50 M~ R.
It is important to select appropriate values of b so that
M max the experiments reflect realistic loading conditions. Typical
b ¼ design loads for an offshore wind turbine are shown in Table
MR 4. The limit state ULS refers to the ultimate load-carrying
(18)
M min capacity and ULS/1.35 to the worst expected transient load.
c ¼ SLS and FLS are the serviceability and fatigue limit states
M max
occurring 102 and 107 times during the lifetime of the wind
in which MR refers to the static moment capacity of the pile, turbine respectively. Further information is given by DNV
and Mmin and Mmax are the minimum and maximum in a (2004).
load cycle. The ratio b is a measure of the size of the The design loads are compared with the laboratory load-

Géotechnique 2010.60:79-90.
RESPONSE OF STIFF PILES IN SAND TO LONG-TERM CYCLIC LATERAL LOADING 85
M Bearing capacity: test results
1·0
MR Bearing capacity: theory
ζb  0·75 ζb  0·5
Scaled design loads
0·5 1·0
0·25 Cyclic loading ranges: ζb
0·5
0
0·0
ζc  0·0 ζc  0·8
0·5
1·0
ULS K  10
Fig. 6. Characteristics of cyclic loading defined in terms of æb
and æc 0·6
~ ULS/1·35
M

1·4 0·4
SLS
60%
1·2
~ 50%
MR  1·24 FLS
1·0 40%
0·2
Rd  38%
30%
~ 0·8
M
20%
0·6
~ 0
0·4 MR  0·6 0 0·2 0·4 0·6 0·8 1·0 1·2
Rd  4% ~
H
0·2
Fig. 8. Cyclic loading ranges, in terms of æb , in relation to
0 design loads of a typical offshore wind turbine
0 0·02 0·04 0·06 0·08 0·1
~
θ

Fig. 7. Moment capacity determined from static load tests for cycles. The evolution of the accumulated rotation is evalu-
e~ 1:19 ated in terms of the dimensionless ratio
˜Łð N Þ Ł N  Ł0
¼ (19)
Łs Łs
Table 4. Typical design loads for a 2 MW turbine
which expresses the magnitude of the rotation ˜Ł(N) caused
N M: MN m H: MN V: MN by cyclic loading in terms of the rotation Łs that would
occur in a static test when the load is equivalent to the
ULS 1 95 4.6 5.0
ULS/1.35 1 70 3.4 5.0 maximum cyclic load (as defined by b 3 M R ). The non-
SLS 102 45 2.0 5.0 dimensional stiffness k~ is obtained substituting measured
FLS 107 28 1.4 5.0 values of k ¼ M/Ł (see Fig. 9) into equation (10), which is
rearranged to give
k
k~ ¼ pffiffiffiffiffiffiffiffiffi (20)
ing, in terms of non-dimensional parameters M ~ and H ~ , by L5=2 D pa ª9
scaling the design loads such that ULS coincides with the
static moment capacity of the laboratory pile. Fig. 8 shows
this static capacity, determined from equation (15), com- Accumulated displacements
pared with the capacities determined from five static load The method proposed by Lin & Liao (1999) suggests that
tests in loose sand, conducted at different values of e~. Also accumulated rotation is proportional to ln(N). This approach
shown on the figure are ranges of b for e~ ¼ 1:19. The was investigated by plotting ˜Ł/Łs as a function of ln(N). A
comparison suggests that 30% , b , 50% is the range of good fit was obtained for N , 100, but extrapolation beyond
primary interest for piles with designs governed by the ULS. N . 500 underestimated the accumulated rotation. A better
The value of c is expected to vary between 0.5 and 1, fit was found if the accumulated rotation was modelled as
since the response of a wind turbine is governed by large increasing exponentially with N rather than logarithmically,
aerodynamical damping, resulting in one-way cycling rather as is in agreement with the method proposed by Little &
than two-way cycling. For completeness, c was investigated Briaud (1988) and Long & Vanneste (1994). The exponential
in the full range from 1 to 1. The chosen test programme behaviour appears as straight lines in double logarithmic
is summarised in Table 5. axes, as shown in Fig. 10.
The results for one-way loading, plotted in Figs 10(a) and
10(b), show a very good fit with an exponential expression.
DISCUSSION The results include approximately 104 load cycles, whereas
The results of the laboratory tests are investigated by the fatigue limit state is governed by 107 load cycles. The
plotting the angular rotation of the pile Ł in response to closeness of fit up to 104 cycles indicates that, in the
the applied moment M for both static and cyclic tests. The absence of further experimental data, it might be reasonable
method for data extraction is outlined in Fig. 9. The to extrapolate to N ¼ 107 . Further data are, of course,
gathered data provides information on both stiffness and required to confirm this hypothesis.
accumulated rotation as functions of N, the number of load The results obtained by varying c , plotted in Figs 10(c)

Géotechnique 2010.60:79-90.
86 LEBLANC, HOULSBY AND BYRNE
Table 5. Test programme

No. Type e~ Rd : % : degrees b c N

1 Static 0.10 4 35 – – –
2 Static 0.42 4 35 – – –
3 Static 0.78 4 35 – – –
4 Static 1.19 4 35 – – –
5 Static 3.33 4 35 – – –
6 Cyclic 1.19 4 35 0.20 0 8 200
7 Cyclic 1.19 4 35 0.27 0 18 200
8 Cyclic 1.19 4 35 0.34 0 8 400
9 Cyclic 1.19 4 35 0.40 0 17 700
10 Cyclic 1.19 4 35 0.53 0 8 600
11 Cyclic 1.19 4 35 0.40 0.98 8 510
12 Cyclic 1.19 4 35 0.40 0.67 7 400
13 Cyclic 1.19 4 35 0.40 0.33 8 800
14 Cyclic 1.19 4 35 0.40 0.33 65 370
15 Static 1.19 38 43 – – –
16 Cyclic 1.19 38 43 0.27 0 8 090
17 Cyclic 1.19 38 43 0.40 0 7 423
18 Cyclic 1.19 38 43 0.52 0 17 532
19 Cyclic 1.19 38 43 0.40 0.50 9 003
20 Cyclic 1.19 38 43 0.40 0.80 9 814
21 Cyclic 1.19 38 43 0.40 0.50 9 862

accepted, and the majority of lateral load tests reported in


k0 kN
the literature are conducted at c ¼ 0, with some at c ¼ 1.
∆θ(N) However, the results presented here clearly illustrate that
θs
Mmax loading with c  0.6 causes an accumulated rotation that
is more than four times larger than for one-way loading,
M M
that is, c ¼ 0. The authors are not aware of similar observa-
tions in cyclic loading tests, but clearly this result has
profound implications for assessing the results of cycling.
Mmin
θ0 θN
θ θ Variation of pile stiffness
(a) (b) Interpretation of the stiffness results involves a greater
scatter of the data. This is partly because the measurement
Fig. 9. Method for determination of stiffness and accumulated of secant stiffness in a cycle involves differences of small
rotation: (a) cyclic test; (b) static test
displacements. Plotting k~N as a function of ln(N) indicates
that the stiffness evolves approximately logarithmically with
and 10(d), exhibit a more volatile behaviour, particularly for cycle number, as shown in Fig. 12. This suggests that the
c , 0. However, the trend in the data also follows the evolution of stiffness can be approximated by
exponential behaviour shown in the one-way load tests.
Based on these observations, it is proposed that displace- k~N ¼ k~0 þ A k ln ð N Þ (22)
ments due to cyclic loading can be predicted by
where A k is a dimensionless constant. It is observed from
˜Łð N Þ :
¼ Tb ðb , Rd ÞTc ðc Þ  N 0 31 (21) Fig. 12 that all slopes are almost equal. This suggests that
Łs Ak is independent of both relative density and load charac-
teristics within the observed range. The expression in equa-
in which Tb and Tc are dimensionless functions, depending tion (22) was fitted to the data in Fig. 12 (the dotted lines)
on the load characteristics and relative density. The function using the value Ak ¼ 8.02, and values of k~0 were determined
Tc is defined such that Tc (c ¼ 0) ¼ 1. The expression in from the point of intersection with the k~-axis where N ¼ 1.
equation (21) was fitted to the data in Fig. 10 (the dotted The empirically determined values of k~0 can be expressed
lines) to empirically determine values of Tb and Tc. These by
values are plotted in Fig. 11 as functions of b and c
respectively. The behaviour of the functions Tb (b , Rd ) and k~0 ¼ K b ðb ÞK c ðc Þ (23)
Tc (c ) is clearly apparent, and curves were easily fitted.
Generally, the loose sand results in low values of Tb as in which Kb and Kc are dimensionless functions, depending
compared with the denser sand. The value of Tc is found to on the load characteristics and relative density. The function
be independent of relative density. Kc is defined such that Kc (c ¼ 0) ¼ 1. The empirically
The Tc curve in Fig. 11 shows a remarkable result. determined values of Kb and Kc , as functions of b and c
Clearly, when c ¼ 1, then Tc must be zero, since no respectively, are illustrated in Fig. 13. The behaviour of the
accumulated displacement will occur under static load. Also, functions Kb (b ) and Kc (c ) was easily determined and
when c ¼ 1, then it is expected that Tc will be zero, since curves fitted. It is not possible to make a clear distinction
the force applied is equal in both directions. The maximum between the results for Rd ¼ 4% and Rd ¼ 38%. This indi-
one-way load is obtained when c ¼ 0, and intuitively it cates that values of stiffness are somewhat independent of
seems reasonable to expect that this loading will cause the the relative density, at least for the low to medium densities
largest accumulated rotation. This assumption is commonly tested. However, this is unlikely to hold for Rd ! 100%,

Géotechnique 2010.60:79-90.
RESPONSE OF STIFF PILES IN SAND TO LONG-TERM CYCLIC LATERAL LOADING 87
1 1
10 10
ζc  0·00 ζb  0·53 ζc  0·00
Rd  4% Rd  38%
ζb  0·40

100 ζb  0·34 100 ζb  0·52

∆θ ∆θ
ζb  0·27 ζb  0·40
θs θs

ζb  0·20 ζb  0·27
101 101

102 0 102
10 101 102 103 104 105 100 101 102 103 104 105
N N
(a) (b)

101 ζb  0·40 101 ζb  0·40


Rd  4% ζc  0·33 Rd  38%
ζc  0·50
ζc  0·00
0 0
10 10 ζc  0·00
ζc  0·34
∆θ ∆θ
θs θs ζc  0·50
ζc  0·67

101 101 ζc  0·81


ζc  0·98

102 102 0
100 101 102 103 104 105 10 101 102 103 104 105
N N
(c) (d)

Fig. 10. Measured displacements as a function of N, Rd , æb and æc . Dotted lines obtained using equation (21)

since no increase in stiffness is expected for sand in its Given Łs and k~0, it is possible to estimate the accumulated
densest state. rotation and stiffness change due to long-term cyclic load-
The most important outcome of the results is that stiffness ing. Values of Tb and Tc are determined from Fig. 11, with
always tends to increase. This observation opposes the the representative relative density chosen as Rd ¼ 4%, since
current methodology of degrading static p–y curves to the angle of friction is 358. The resulting increase in
account for cyclic loading. stiffness follows from equation (22), as

k~N ¼ k~0 þ 8:02ln ð107 Þ  345


Example
An example is given to demonstrate the use of the k~N  k~0 (24)
proposed methods. Consider a stiff monopile, with L ¼ 18 m ) ¼ 60%
k~0
and D ¼ 4 m, driven into sand with a friction angle of 358.
The task is to determine the increase in stiffness and
This result indicates that the stiffness can be expected to
accumulated rotation due to 107 load cycles characterised by
increase by approximately 60% during the lifetime of the
b ¼ 0.3 and c ¼ 0.2.
wind turbine. The accumulated rotation follows from equa-
Initially, values of Łs and k~0 must be determined. These
tion (21), as
can be calculated by various methods, for example using
p–y curves, finite element models or, alternatively, using the ˜Ł 0:31
non-dimensional framework which has been presented here. ¼ 0:047 3 1:5 3 ð107 Þ  10:4
The non-dimensional approach requires that the full-scale Łs (25)
structure and laboratory pile have comparable values of ~, V~ ) ˜Ł ¼ 0:0344 rad ( 28)
and e~, as is the case in this example. The non-dimensional
static rotation for b ¼ 0.3 gives a non-dimensional moment The accumulated rotation is estimated to be 28, which is a
of M ~ ¼ 0:3 3 M ~ R ¼ 0.3 3 0.6 ¼ 0.18, at which point the value that would breach the tolerance criterion. It should be
non-dimensional rotation is Ł~ ¼ 0.0023. The corresponding noted that the accumulated rotation is calculated on the basis
static rotation of the full-scale monopile follows from the of 107 load cycles, acting in the same direction. Less
definition of Ł~, which gives Łs ¼ 0.0033. The initial non- rotation must be expected, since the actual loading would be
dimensional stiffness is determined from equation (23), by multidirectional. Of course, if b is in the range between
evaluating Kb and Kc from Fig. 13, to obtain k~0 ¼ 240 3 0.7 and 0.4, then much higher rotation is predicted.
0.9  216. The non-dimensional stiffness can optionally be If the monopile is more conservatively designed, say by
transformed to the absolute value of the full-scale stiffness using a static design capacity equal to 1.5 times ULS, then
using the relationship in equation (20). b will be approximately 0.2. In this case, the predicted

Géotechnique 2010.60:79-90.
88 LEBLANC, HOULSBY AND BYRNE
0·25
to be less severe in terms of accumulated rotation, as
Rd  38%
compared with unidirectional loading.
The tests showed that cyclic loading always increased the
0·20
pile stiffness, and the increase was found to be independent
of relative density. This contrasts with the current method-
ology of degrading static p–y curves to account for cyclic
0·15
Rd  4% loading. A method, based on the experimental work carried
Tb out, is presented to predict changes in stiffness due to long-
term cyclic loading.
0·10
The results in this paper lay out a basic framework to
incorporate effects of cyclic loading in a simple manner.
Further work should be carried out to investigate piles
0·05 installed in very dense sand, the effect of altering pile
dimensions, and how a representative cyclic load is chosen.
The effect of the loading frequency on the drained response
0 of laterally loaded piles in sand is limited. However, owing
0 0·2 0·4 0·6 0·8
ζb
to the scale of field monopiles it is possible that the
(a) response in saturated sand may not be completely drained,
and further work on the effects of loading frequency will
5 need to be undertaken. Finally, comparisons with full-scale
Rd  4% measurements should be carried out to ensure that the
proposed methods are reliable and valid.
4 Rd  38%

NOTATION
3 c1 , c2 , c3 dimensionless constants
Tc D pile diameter
D10, D60 particle sizes
2 d pile pivot point
Ep elastic modulus of pile
Es elastic modulus of soil
e load eccentricity
1
f, f0 , fa load rig forces
G shear modulus
Gs specific gravity
0 g gravitational acceleration
1·0 0·5 0 0·5 1·0
ζc
H horizontal load at sea-bed
Ip moment of inertia of pile
(b)
K Broms factor
Fig. 11. Functions relating (a) Tb and (b) Tc to relative density Kb , Kc dimensionless functions
Rd , and characteristics of cyclic load in terms of æb and æc k pile stiffness
k1 , k2 , k3 dimensionless parameters
k0 pile stiffness in first cycle
accumulated rotation and increase in stiffness are 0.278 and kN pile stiffness in Nth cycle
L penetration depth of pile
42% respectively. l2 , la , lc load rig dimensions
M, Mmin , Mmax moment at sea-bed
MR static moment resistance of pile
CONCLUSION m1 , m2 , m3 load rig masses
A series of tests was conducted on small-scale driven N number of load cycles
piles subjected to long-term cyclic loading. A typical design pa atmospheric pressure
of a monopile was adopted and used to quantify realistic p9 effective isotropic stress
pile dimensions and loading ranges. Furthermore, a complete Rd relative density
Tb, Tc dimensionless functions
non-dimensional framework for stiff piles in sand is pre-
t time
sented and applied to interpret the test results. V gravity force acting on the structure
The accumulated rotation of a stiff pile is largely affected y horizontal deflection
by the characteristics of the applied cyclic load. Thus z depth below sea-bed
parameters characterising the load, other than maximum load Æ dimensionless parameter
levels, are required for accurate predictions. For example, ªmin , ªmax dry unit weight
results for one-way loading were found to differ by a factor ª9 effective unit weight
of four as compared with two-way loading. A very signifi- b , c load characteristic parameters
cant result was that the most onerous loading condition was  pile aspect ratio
found to be between one-way and two-way loading. A Ł pile rotation
Ł0 pile rotation in first cycle
method to predict the accumulated rotation during the life- ŁN pile rotation in Nth cycle
time of an offshore wind turbine foundation is presented. Łs static pile rotation
When applied, the method predicts that typical tolerances for  v9 effective vertical stress
accumulated rotation are breached if the foundation is  angle of friction
designed so that the design capacity is equal to the ULS cr critical state angle of friction
load. This suggests that considerations of accumulated rota- ø rotational frequency
tion are the primary design driver. The proposed method Note: , above parameters indicates corresponding dimen-
does not account for multidirectional loading, which is likely sionless values: see Table 1.

Géotechnique 2010.60:79-90.
RESPONSE OF STIFF PILES IN SAND TO LONG-TERM CYCLIC LATERAL LOADING 89
450 450
ζb  0·53 ζb  0·52

400 ζc  0·0
400 ζb  0·40 ζb  0·40
Rd  38%
350 ζb  0·34 350
ζb  0·27

300 ζb  0·27 300


~ ~
kN kN
250 ζb  0·20 250

200 200

ζc  0·0 150
150
Rd  4%
100 100
100 101 102 103 104 105 100 101 102 103 104 105
N N
(a) (b)

450 450 ζc  0·50


ζc  0·33
ζb  0·40 ζb  0·40 ζc  0·50
400
Rd  4% ζc  0·34 400
Rd  38%
ζc  0·81
350 ζc  0·67 350

300 ζc  0·98 300


~ ~
kN kN
250 250

200 200

150 150

100 100
100 101 102 103 104 105 100 101 102 103 104 105
N N
(c) (d)

Fig. 12. Measured stiffness as a function of N, Rd , æb and æc . Dotted lines are obtained using equation (22)

400 1·6

Rd  4%
350
1·4
Rd  38%
300

1·2
250

Kb 200 Kc 1·0

150
0·8
100

0·6
50

0 0·4
0 0·1 0·2 0·3 0·4 0·5 0·6 1·0 0·5 0 0·5 1·0
ζb ζc
(a) (b)

Fig. 13. Values of (a) Kb and (b) Kc as function of relative density Rd , and characteristics of cyclic load in terms of æb and æc

REFERENCES Broms, B. B. (1964). Lateral resistance of piles in cohesionless soils.


API (1993). Recommended practice for planning, designing, and J. Soil Mech. Found. Engng Div. ASCE 90, No. 3, 123–156.
constructing fixed offshore platforms: working stress design, Budhu, M. & Davies, T. (1987). Nonlinear analysis of laterally
RP2A-WSD, 20th edn. Washington, DC: American Petroleum loaded piles in cohesionless soils. Can. Geotech. J. 24, No. 2,
Institute. 289–296.
Briaud, J. L., Smith, T. D. & Meyer, B. J. (1984). Using pressure- Carter, J. & Kulhawy, F. (1988). Analysis and design of drilled shaft
meter curve to design laterally loaded piles. Proc. 15th Ann. foundation socketed into rock. Ithaca, NY: Cornell University.
Offshore Technol. Conf., Houston, TX, paper no. 4501. Cox, W. R., Reese, L. & Grubbs, B. R. (1974). Field testing of

Géotechnique 2010.60:79-90.
90 LEBLANC, HOULSBY AND BYRNE
laterally loaded piles in sand. Proc. Offshore Technol. Conf., deformation of sands in plane strain compression. Proc. 13th
Houston, TX, paper no. 2080. Int. Conf. Soil Mech. Found. Engng, New Delhi 1, 1–6.
DNV (1977). Rules for the design, construction and inspection of Peng, J. R., Clarke, B. & Rouainia, M. (2006). A device for cyclic
offshore structures. Hovek, Norway: Det Norske Veritas. lateral loaded model piles. Geotech. Test. J. 29, No. 4, 1–7.
DNV (2004). Offshore standard: Design of offshore wind turbine Porovic, E. & Jardine, R. J. (1994). Some observations on the
structures, DNV-OS-J101. Hellerup, Denmark: Det Norske Ver- static and dynamic shear stiffness of ham river sand. Proc. IS-
itas. Hokkaido 1994, 1, 25–30.
Dobry, R., Vincente, E., O’Rourke, M. & Roesset, J. (1982). Poulos, H. & Hull, T. (1989). The role of analytical geomechanics
Stiffness and damping of single piles. J. Geotech. Engng 108, in foundation engineering. In Foundation engineering: Current
No. 3, 439–458. principles and practices, Vol. 2, pp. 1578–1606. Reston, VA:
Kelly, R. B., Houlsby, G. T. & Byrne, B. W. (2006). A comparison ASCE.
of field and laboratory tests of caisson foundations in sand and Reese, L. C. & Impe, W. F. (2001). Single piles and pile groups
clay. Géotechnique 56, No. 9, 617–626. under lateral loading. London: Taylor & Francis.
Lesny, K. & Hinz, P. (2007). Investigation of monopile behaviour Reese, L. & Matlock, H. (1956). Nondimensional solutions for
under cyclic lateral loading. Proc. 6th Int. Conf. on Offshore laterally loaded piles with soil modulus assumed proportional to
Site Investigation and Geotechnics, London, 383–390. depth. Proc. 8th Texas Conf. on Soil Mech. and Found. Engng,
Lin, S. S. & Liao, J. C. (1999). Permanent strains of piles in sand Austin, TX.
due to cyclic lateral loads. J. Geotech. Geoenviron. Engng 125, Reese, L., Cox, W. R. & Koop, F. D. (1974). Analysis of laterally
No. 9, 798–802. loaded piles in sand. Proc. 6th Offshore Tech. Conf., Houston 2,
Little, R. L. & Briaud, J. L. (1988). Cyclic horizontal load tests on paper no. 2079.
six piles in sands at Houston Ship Channel. Research Report Rovere, M. (2004). Cyclic loading test machine for caisson suction
5640 to USAE Waterways Experiment Station. Civil Engineer- foundations. Ecole Centrale de Lille/Politecnico di Milano.
ing, Texas A&M University. Schnaid, F. (1990). A study of the cone-pressuremeter test in sand.
Long, J. & Vanneste, G. (1994). Effects of cyclic lateral loads on PhD thesis, University of Oxford.
piles in sand. J. Geotech. Engng 120, No. 1, 225–244. Simpson, B. (2002). Engineering needs. In Pre-failure deformation
McClelland, B. & Focht, J. (1958). Soil modulus for laterally loaded characteristics of geomaterials (eds M. Jamiolkowski, R. Lan-
piles. Trans. ASCE 123, 1049–1086. cellotta and D. Lo-Presti), pp. 1011–1026. Torino: Swets &
Murchison, J. M. & O’Neill, M. W. (1984). Evaluation of p–y Zeitlinger.
relationships in cohesionless soils. Proceedings of the ASCE Wroth, C. P. & Houlsby, M. F. (1985). Soil mechanics: property
symposium on analysis and design of pile foundations, San characterisation and analysis procedures. Proc. 11th Int. Conf.
Francisco, pp. 174–191. Soil Mech. Found. Engng, San Francisco, 1–50.
O’Neill, M. W. & Murchison, J. M. (1983). An evaluation of p–y Wroth, C. P., Randolph, M. F., Houlsby, G. T. & Fahey, M. (1979).
relationships in sands, Research Rep. No. GT-DF02-83. Depart- A review of the engineering properties of soils with particular
ment of Civil Engineering, University of Houston, TX. reference to the shear modulus, Report CUED/D-SOILS TR75.
Park, C. S. & Tatsuoka, F. (1994). Anisotropic strength and University of Cambridge.

Géotechnique 2010.60:79-90.

Vous aimerez peut-être aussi