You are on page 1of 14

Ocean Engineering 226 (2021) 108299

Contents lists available at ScienceDirect

Ocean Engineering
journal homepage: www.elsevier.com/locate/oceaneng

Centrifuge testing on monotonic and cyclic lateral behavior of


large-diameter slender piles in sand
Huan Wang a, Lizhong Wang a, *, Yi Hong a, Davis Mašín b, Wei Li c, Ben He c, Hualin Pan d
a
Key Laboratory of Offshore Geotechnics and Material of Zhejiang Province, College of Civil Engineering and Architecture, Zhejiang University, 866 Yuhangtang Rd,
Xihu, Hangzhou, Zhejiang, 310027, China
b
Faculty of Science, Charles University in Prague, Albertov 6, 128 43, Praha, Czechia
c
Key Laboratory for Far-shore Wind Power Technology of Zhejiang Province, Power China Huadong Engineering Corporation Limited, 201 Gaojiao Rd, Yuhang,
Hangzhou, Zhejiang, 311122, China
d
Construction Management Department, Zhejiang Province Energy Group Company LTD., 152 Tianmushan Rd, Xihu, Hangzhou, Zhejiang, 310007, China

A R T I C L E I N F O A B S T R A C T

Keywords: The existing studies have been primarily focused on the lateral behavior of large-diameter stubby pile or small-
Centrifuge modelling diameter slender pile in sand, with little attention paid to large-diameter slender pile. This study presents a
Large-diameter slender pile unique series of centrifuge tests on monotonic and cyclic lateral behavior of heavily instrumented large-diameter
Medium dense sand
slender piles in medium dense sand. Two typical length to diameter ratios (L/D) are considered with the same
p-y curve
length (L = 60 m) but different diameters (D = 4 and 6 m). It is found that the lateral behaviors of large-diameter
Cyclic loading
Lock-in bending moment slender pile, including its monotonic p-y response, cyclic accumulation of lateral displacement and cyclic stiffness
evolution, are marginally different from those of small-diameter slender piles, but significantly deviate from the
large-diameter stubby piles. This may suggest the longstanding argument of ‘diameter effect’ is relatively minor,
while the lateral behavior of monopile in sand is more significantly governed by the relative pile-soil stiffness.
The API (2011) non-conservatively predicts both stiffness and capacity of the large-diameter slender piles,
leading to development of a new p-y formulation. These centrifuge testing results form a unique database to
support development of new design methods for large-diameter slender piles, and to verify advanced numerical
analyses involving cyclic models.

1. Introduction aspects: (a) the ultimate lateral capacity, which aims to avoid collapse in
an extreme loading event; (b) cyclic stiffness degradation, which affects
Offshore wind power industry has been experiencing a rapid devel­ repeated accumulation of pile deformation and therefore fatigue dam­
opment in the recent decades (Kaynia, 2019; The EWEA, 2020). Up to age within the pile and the turbine structure (Jeng, 2001; Liu et al.,
2018, more than 20 GW offshore wind turbines have been installed 2015; Markou and Kaynia, 2018; Wang et al., 2020). The lateral
worldwide and more than 520 GW of total installed capacity is antici­ behavior of monopile in sand may be potentially affected by pile
pated by 2050 (Cohen, 2019). One key factor that would limit the diameter D (Sørensen, 2012; Doherty and Gavin, 2012), or the ratio of
economic development of offshore wind turbines lies in the foundation, pile embedment length over diameter L/D (Li et al., 2017; Thieken et al.,
which takes up to 30–40% of the total project cost (Byrne and Houlsby, 2015; Hong et al., 2017).
2003; Sun et al., 2012). Of all different foundation types, the monopile Extensive experimental and theoretical investigations have been
originating from the laterally loaded single pile in oil and gas industry is performed to understand the monotonic and cyclic lateral behavior of
the most widely adopted for its relatively low fabrication and installa­ piles (Klinkvort et al., 2010; Alderlieste, 2011; Leth, 2013; Choo and
tion cost (Doherty and Gavin, 2012; Wang et al., 2018). By 2018, more Kim, 2015; Zhu et al., 2016; Qi et al., 2016 and Wang et al., 2019). These
than 81% of 5258 installed offshore wind turbines in European countries studies mainly focus on the behavior of small-diameter slender piles
are supported by the monopile foundations (The EWEA, 2020). (typically with D < 3 m and L/D > 20) and large-diameter stubby piles
The design of laterally loaded monopile principally considers two (typically with D > 4 m and L/D < 8), as illustrated in Fig. 1 (Gilbert

* Corresponding author.
E-mail address: wlzzju@163.com (L. Wang).

https://doi.org/10.1016/j.oceaneng.2020.108299
Received 10 April 2020; Received in revised form 27 August 2020; Accepted 25 October 2020
Available online 26 March 2021
0029-8018/© 2020 Elsevier Ltd. All rights reserved.
H. Wang et al. Ocean Engineering 226 (2021) 108299

large-diameter slender piles in sand.

2. Centrifuge modelling

All of the centrifuge model tests reported in this study were carried
out in the geotechnical centrifuge facility (GCF) at the Hong Kong
University of Science and Technology (HKUST). The beam centrifuge,
with a diameter of the rotating arm of 8.4 m, has the maximum carrying
capacity of 400g-ton (Ng et al., 2001).
For the monopile foundation, existing study suggests that the
drainage condition around the monopile is usually fully drained due to
the high permeability of sand and low frequency of lateral loads (Lai
et al., 2019). Klinkvort (2012) proved that tests performed using dry
sand offer consistent results as compared to those performed in satu­
rated sand under the same effective stress level. Therefore, the centri­
fuge tests reported herein were all performed in dry sand at a centrifugal
acceleration of 100g, aiming to save the test preparation time and pre­
vent the potential entry of water into the coating that protects strain
gauges. Scaling factors for various physical quantities between the
centrifuge model and the intended prototype relevant to these centrifuge
tests are summarized in Table 1. Derivations for these scaling laws are
given in Taylor (1995).

2.1. Experimental program and objectives

The unique geometric and geotechnical properties of monopiles


Fig. 1. Monopile geometries of this study and those in the literatures.
practiced in offshore China are considered in this experimental program,
i.e., large-diameter slender piles in medium dense sand. A total of four
et al., 2018; Truong et al., 2018). The former pile is usually designed for
series of centrifuge tests were carried out, including monotonic and
foundations that support jacket foundations, while the latter pile is
multi-stage cyclic loading tests for two large-diameter long slender
mainly for foundations supporting offshore wind turbines in relatively
monopiles. The two piles have different diameters (i.e., D = 4 and 6 m in
stiff seabed (e.g., dense sand or stiff clay), as frequently encountered in
prototype) but the same embedded length (i.e., L = 60 m), leading to
the North Sea (Byrne et al., 2015).
relative length-over-diameter ratios, namely L/D = 15 and 10.
In addition to the two afore-mentioned types of piles, there has been
Two monotonic loading tests were performed first to investigate the
emerging a new pile type with a unique geometrical form (typically with
lateral behavior of the larger-diameter long slender piles. These tests aim
D > 4 m and L/D > 10), i.e., large-diameter slender pile. This type of pile
to (a) understand the lateral stiffness and capacity of large-diameter long
has been widely used for offshore wind sites with presence of relatively
slender piles, and to (b) examine the applicability of the existing p-y
soft seabed (e.g., medium dense sand or soft clay) along with typhoon
models, that were derived based on small-diameter slender piles or
loadings, such as offshore China, to ensure sufficient lateral resistance
large-diameter stubby piles.
(Lai et al., 2019). At the same time, turbines of up to 15 MW with rotor
Following the monotonic tests, two series of multi-stage load
diameters of up to 230 m are expected to be installed in the future
controlled cyclic loading tests were performed to investigate the long-
around the world. These giant turbines are anticipated to be supported
term cyclic response of the two large-diameter slender monopiles. The
by large-diameter slender pile to ensure sufficient foundation stiffness
primary objectives of these tests are to (a) assess how do the cumulative
and resistance (Steelwind Nordenham, 2020). In addition to the offshore
rotation characteristics of these unique piles compared with that of the
wind farm, the large-diameter slender pile can also be used as breasting
reported small-diameter slender piles and larger-diameter stubby piles,
and mooring dolphins for berthing or mooring a vessel (Winkel, 2016).
and to (b) understand the evolution of cyclic lateral stiffness and
Since the geometries of the large-diameter slender pile fall significantly
bending moments (including peak and lock-in moment), which deter­
beyond the parameter space of all the existing studies, questions arise
mine the fatigue damage of the monopile. Details on the centrifuge test
concerning its monotonic p-y response, cyclic characteristics including
program in this study are summarized in Table 2.
deformation accumulation and stiffness degradation. This research gap
To the authors’ best knowledge, this work is the first to study the
has necessitated high quality experiments that offer fresh insights into
lateral behavior of large-diameter slender piles (D > 4 m and L/D > 10)
the lateral behavior of large-diameter slender pile, for supporting the
development of a vast number of wind farms being planned along
Table 1
offshore China.
Scaling factors relevant to centrifuge tests in this study (Taylor, 1995).
In light of these discussions, this study presents a unique series of
centrifuge model tests simulating large-diameter slender piles in me­ Physical quantity Scaling factor (Model/Prototype)

dium dense sand subjected to monotonic and multi-stage cyclic lateral Gravitational acceleration n
loadings. The tests were performed on two heavily instrumented Length 1/n
Area 1/n2
monopiles with different diameters (D = 4 and 6 m in prototype) but the
Volume 1/n3
same embedded depth (L = 60 m in prototype), resulting L/D of 15 and Settlement n
10, respectively. The experimental program was specifically designed to Stress 1
conform the likely conditions of monopiles in offshore China. The Strain 1
measured results concerning the various monotonic and cyclic responses Force 1/n2
Density 1
of the pile were interpreted, and compared with those of the reported
Mass 1/n3
small-diameter slender piles and large-diameter stubby piles. A new p-y Flexural rigidity 1/n4
relationship is proposed to describe the lateral load-transfer behavior of Bending moment 1/n3

2
H. Wang et al. Ocean Engineering 226 (2021) 108299

Table 2
Centrifuge test objectives and program.
Test Soil type Pile Geometry Loading condition
ID

T1 Toyoura sand (Dr D = 4 m, L = Monotonic


T2 = 65%) 60 m Cyclic (cyclic amplitude# = 10, 20,
30, 40 and50%Fu*)

T3 D = 6 m, L = Monotonic
T4 60 m Cyclic (cyclic amplitude# = 10, 20,
30, 40 and50%Fu*)
#
For each cyclic amplitude, 1000 cycles of repeated loading were applied; the
load period was 5 s, which was slow enough to eliminate the dynamic effect.
*
Fu denotes the ultimate lateral capacity of the pile, as determined from the
corresponding monotonic test.

in sand, while the previous work focused on lateral response of either


large-diameter stubby piles (typically with D > 4 m and L/D < 8) or
small-diameter slender piles (typically with D < 3 m and L/D > 20).

2.2. Centrifuge modelling package

Fig. 2(a) shows the plane view of the centrifuge model package. All
the dimensions are in prototype (in meter) except for numbers in pa­
rentheses, which denote model scale (in millimeter). The model box has
an internal plan area of 124.5 m by 124.5 m in prototype at a centrifugal
acceleration of 100g, which is sufficiently large to accommodate all the
four tests (as summarized in Table 2) with minimum boundary effect.
The horizontal spacing between each two piles was designed to be no
smaller than 8 D, while the distance of the model pile to each boundary
was no smaller than 6 D. Numerical (Ahmed and Hawlader, 2016) and
experimental (Klinkvort, 2012; Kirkwood, 2016) studies showed that
the distance between the model piles and pile to the container boundary
in this study is larger enough to eliminate any boundary effect.
Fig. 2(b) presents the elevation view of the centrifuge model box. The
seabed has a thickness of 720 mm (72 m in prototype). The model
monopiles were prefixed with an aluminum frame to the target position
and elevation before raining the sand into the model box. The loading
eccentricity in each test is 10 m (in prototype) above the ground surface.
For each test, three linear variable differential transformers (i.e., LVDTs)
were installed at different heights to deduce the rotation and displace­
ment at ground surface of each model pile. The monotonic or cyclic load
in each test was applied through a servo-controlled hydraulic actuator.
The extension rod of the actuator was attached with a load cell to offer
feedbacks to the closed-loop system in the load-controlled tests, along
with a LVDT to record the induced displacement at the loading height.
The hydraulic actuator was mounted on a rigid frame, which can be
moved to fit the location of four tested piles in this study. The loading
head pieces was curved to guarantee a point interaction between the
lateral loading and the pile without any moment. Fig. 2(c) shows a photo
illustrating the details of the pile head instrumented with LVDTs while
subjected to a lateral load.
Fig. 2. Centrifuge model package: (a) elevation view; (b) plane view; (c) a local
2.3. Model pile and instrumentation view of the instrumented pile head under lateral load.

The model piles in this study were fabricated from 750 mm long strain gauges that were arranged as a full Wheatstone bridge, aiming to
7075-T6 aluminum alloy pipes. The material has a compression yield minimize the temperature effect on the measured bending moment. To
strength of 572 MPa, a tension yield strength of 503 MPa, an elastic protect the strain gauges, each instrumented model pile was externally
modulus of 71.7 GPa and a poison’s ratio of 0.33. The wall thickness of coated with a 1 mm thick epoxy (with an elastic of 2 GPa). The contri­
each model pile was manufactured to be 2 mm, so as to meet the bution of the epoxy coating on the bending stiffness of the pile has been
equivalent section bending stiffness of typical monopile in offshore wind taken into account, as summarized in Table 3.
projects (Negro et al., 2017). Detail of the model piles and the corre­
sponding prototype are summarized in Table 3. 2.4. Centrifuges model preparation and test procedure
Fig. 3 shows the typical model piles of two different diameters used
in this study. Each model pile was instrumented with 27 levels of foil Toyoura sand, a standard testing material, was used in all the ex­
type strain gauges to measure the bending moment along the pile, which periments reported herein. The sand consists of sub-angular particles
form the basis for deriving the p-y curve. Each level consists of four with a mean diameter (D50) of 0.17 mm and a uniformity coefficient (Uc)

3
H. Wang et al. Ocean Engineering 226 (2021) 108299

Table 3 aluminum frame. Toyoura sand was then “rained” into the model box
Summary of model pile properties in this study. from a big sand hopper that was lifted by a crane at a constant falling
Model pile 1 Model pile 2 height of 0.5 m above the rising surface of the sand. The average relative
density of the sand bed in this study is approximately 65%, as summa­
Dmodel (m) 0.04 0.06
Dprototype (m) 4 6 rized in Table 2. After preparing the sandy seabed, the aluminum frame
tmetal (m) 0.002 0.002 that constrained the four model piles were dismantled. This was fol­
texpoxy (m) 0.001 0.001 lowed by installation of the lateral loading system and the instrumen­
D/t 20 30 tation (i.e., LVDTs) for measuring lateral pile displacement. The
EIprototype (GNm2) 311 1106
L/D 15 10
centrifuge model package was then transferred to the swinging platform
e/D 2.5 1.67 and operated to the centrifugal acceleration of 100g in 10 steps, with a
Toe condition Open Open constant increment of 10g for each. Upon reaching 100g, a waiting
Installation Prefixed (wished-in place) Prefixed (wished-in place) period of 2 h was imposed to enable the stabilization of ground settle­
ment, before the commencement of each load-controlled lateral pile test.
The loading condition for each test (including amplitude and frequency)
is summarized in Table 2.
The monotonic tests were performed first on model piles P1 and P2
(as shown in Fig. 2(a)) by displacement control at a loading rate of 0.1
mm/s. The loading rate is slow enough to eliminate any dynamic effect.
After the completion of the monotonic tests, a series of load-controlled
cyclic tests were performed on model piles P3 and P4 (as shown in
Fig. 2(a)) with a period of 5 s for each loading cycle. Fig. 4(a) and (b)
shows the measured episodes of cyclic loads in prototype applied to the
monopiles with D = 4 and 6 m, respectively. A total of five episodes of

Fig. 3. Photo of typical instrumented model piles.

of 1.7. Its maximum and minimum void ratios are 0.977 and 0.597,
respectively. The critical state friction angle of the sand is 31◦ . Selected
properties of the Toyoura sand are summarized in Table 4 (Ishihara,
1993).
The air pluvial deposition method was employed to prepare the
model seabed. Before the preparation of the sand seabed, the model piles
were prefixed at their target position and elevated by a customized

Table 4
Selected properties of Toyoura sand (Ishihara, 1993).
Parameters Values

Mass median diameter D50 (mm) 0.17


Coefficient of uniformity, Uc 1.7
Specific gravity, Gs 2.65
Minimum void ratio, emin 0.597
Maximum void ratio, emax 0.977
Critical state friction angle, ϕ (degree) 31
Fig. 4. Applied load episodes of: (a) D = 4 m; (b) D = 6 m.

4
H. Wang et al. Ocean Engineering 226 (2021) 108299

one cycling with increasing amplitudes (10–50% Fu) were applied to


each monopile. Each loading episode consists of 1000 cycles of repeated
loading at a constant amplitude. It is evident that the servo-controlled
loading system has satisfactorily produced the designed multi-stage
cyclic loads for the two piles.

2.5. Post-processing procedures to derive the p-y curve

According to the Timoshenko beam theory, the bending moment,


pile deflection and the soil reaction follow the relationships below:
∫ (∫ ) ∫ (∫ )
M
y= φdz dz = dz dz (1)
EI

d2 M
p= − (2)
dz2

where y is the lateral deflection of monopile at depth z; φ is the curvature


of the monopile at depth z; M is the bending moment at depth z; EI is the
section flexural modulus of the monopile; and p is the soil reaction per
unit length at depth z. Therefore, the soil reaction p and the deflection y
can be derived by double differentiation and double integration of the
bending moment profile, respectively. Following this theory, the
measured bending moment profile consisting of limited data points
needed to be best-fitted by a certain mathematic function. In this study,
the 7th degree global polynomial approximation was adopted to fit the
Fig. 5. Beam on nonlinear Winkler foundation model (BNWF) for analyzing a
bending moment data, following the common practice for deriving the p-
laterally loaded monopile.
y curve of slender piles (Zhu et al., 2016; Haiderali and Madabhushi,
2016). In deriving the pile deflection profile, additional boundary con­
4.1. Monotonic load-displacement response
ditions are needed to determine the two integration constants that arise
from the double integration of the bending moment profile. The calcu­
Fig. 6 shows the measured load-deflection responses of the two
lated rotation and deflection at the ground surface (from the LVDTs, see
monopiles with varying diameters (D = 4 and 6 m). As illustrated, each
Fig. 2(b)) were taken to serve as the purpose.
pile exhibited a continuous hardening response under the lateral loading
with no apparent yielding point within the range of induced lateral
3. Beam on nonlinear Winkler foundation model (BNWF)
displacement (up to 0.14 D). This trend is consistent with the behavior
observed experimentally for laterally loaded monopiles by many other
The p-y method is the mostly adopted method for analysing the
researches (Choo and Kim, 2015; Kirkwood, 2016; Byrne et al., 2015;
response of laterally loaded monopile. In this method, the monopile is
Zhu et al., 2016). Given the lack of yielding point for the
modelled as a beam on nonlinear Winkler foundation (BNWF) with the
load-displacement curve, a displacement-based criterion that defines the
soil described by a serial of uncoupled nonlinear springs (i.e. p-y curves),
monopile to yield at a threshold lateral displacement of 10% D (Byrne
as shown in Fig. 5. According to the Euler-Bernoulli beam theory, the
et al., 2015) is employed to determine the lateral capacity (Fu) of each
behavior of the BNWF model is governed by the following fourth-order
pile. The deduced lateral capacities of the piles with D = 4 and 6 m are
differential equation:
25 MN and 77 MN, respectively.
d4 y In the same figure, the calculated load-displacement curves based on
EI − p=0 (3)
dz4 BNWF model using the p-y curve suggested by API (2011) are also

where y = lateral deflection of the monopile any give depth of z; p = soil


reaction per unit length; EI = bending stiffness of the monopile.
To solve the fourth-order differential equation of the BNWF model,
either finite difference method or finite element method can be used to
discretize the soil-monopile system. In this study, a beam-spring model
is built in the finite element program ABAQUS (Systèmes, 2007).
Three-dimensional beam elements (B31) are adopted to model the
monopile, while the “Spring1” element in ABAQUS is used to model the
nonlinear springs. The nonlinear spring behavior is simulated by an
external user-defined file, which contains pairs of force-displacement
relations calculated with any given p-y model, such as the one sug­
gested by API (2011).

4. Lateral monotonic behavior of the large-diameter slender


piles

In this and the subsequent sections, all results are interpreted in


prototype scale, unless stated otherwise.

Fig. 6. Monotonic lateral pile response at ground surface.

5
H. Wang et al. Ocean Engineering 226 (2021) 108299

presented for comparison. It is seen that the current practice of API


(2011) has led to an overestimation of the initial stiffness for the piles
with D = 4 and 6 m by 90% and 240%, respectively. This overestimation
could be non-conservative in the sense that, it would drift the design for
natural frequency of the wind turbine to the 1P limit that is susceptible
to resonance and thus fatigue failure. In addition, the adoption of API
(2011) has resulted in an overestimation of the lateral capacity (deduced
based on the 10% D criterion) for the piles with D = 4 and 6 m by 69%
and 52%, respectively.

4.2. Comparison of the p-y curve between this study and the API (2011)

To understand the difference between the measured and the calcu­


lated (by API (2011)) load-displacement behavior, p-y curves were
derived from the experiments at five typical depths and compared to
those suggested by API (2011), as shown in Fig. 7. In the figure, the
deduced soil reaction p was normalized by the pile diameter D and
vertical effect stress (i.e. σ’v) at the corresponding depth, while the lateral
displacement y was normalized by the pile diameter D. By doing so,
identical values of the initial slopes of the normalized p-y curves at
different depths mean the initial lateral stiffness kini = p/y of these p-y
curves (without normalization) increases linearly with depth, so does
the ultimate lateral capacity pult.
It is seen from Fig. 7(c) that for each pile (D = 4 or 6 m), the slope of
the initial part (kini = p/yσ’v) of the experimentally deduced normalized
p/Dσ ’v-y/D curves reduce with depth. Considering the vertical effective
stress σ ’v increase linearly with depth, the decrease of the initial slope
with depth, as shown in Fig. 7(c), implies that the initial stiffness kini
does not increase linearly with depth. Instead, it increases nonlinearly
with depth at decreasing rate. Meanwhile, as shown in Fig. 7(c) for the
two piles, the experimentally deduced normalized p-y curves at different
depths tend to converge to the same ultimate value, suggesting the
normalized ultimate capacity pult/(Dσ’v) is insensitive to pile diameter.
This also means pult of the two piles increases with depth at a constant
rate. These experimental observations are significantly different from
the p-y curves suggested by API (2011), where kini value is assumed to
increase linearly with depth, while pult value is assumed to increase
nonlinearly with depth at an increasing rate. Quantitatively, the p-y
curve suggested by API (2011) has overestimated the experimentally
deduced initial stiffness by more than five times. This explains its over-
prediction for the lateral pile head stiffness (as shown in Fig. 6) and poor
description of the p-y curve (as shown in Fig. 7 (a) and (b)).

4.3. New p-y formulation for large-diameter slender monopile

In view of the inadequacy of the API (2011)’s p-y curve to capture the
behavior of large-diameter slender piles (as shown in Figs. 6 and 7), a
new p-y formulation is proposed based on the unique database on such
piles reported herein. Following the routine practice, the proposed p-y
curve consists of three parts: the overall curve shape function, the initial
stiffness kini and the ultimate soil resistance pult. It has been well
recognized that the shape of p-y curve in many occasions could be
adequately captured by the following hyperbolic function (Georgiadis
et al., 1992; Choo and Kim, 2015; Klinkvort, 2012; Kirkwood, 2016; Zhu
et al., 2016).
y
p= (4)
1
kini
+ pyult

Having selected the curve shape function, the remaining issues are to
formulate kini and pult, in light of the experimental observations of large-
diameter slender piles.
The experimentally deduced distributions of kini with depth along Fig. 7. p-y curves measured in this study and computed by API (2011) of model
piles with D = 4 and 6 m are shown in Fig. 8. The initial stiffness kini is piles: (a) D = 4 m; (b) D = 6 m; (c) D = 4 and 6 m.
determined as the initial slope of the p-y curves, as shown in Fig. 7(c).
For comparison, the deduced kini distribution from experiments of large-
diameter stubby piles (D = 6 m, L = 31–42.8 m), as reported by Choo

6
H. Wang et al. Ocean Engineering 226 (2021) 108299

Where Patm is the atmospheric pressure.


Fig. 9 presents the distributions of ultimate soil pressure calculated
by some commonly used approaches, as proposed by API (2011), Broms
(1964) and Zhang et al. (2005). Note that the test results of the two piles
(D = 4 and 6 m) reported herein in Fig. 7(c) reveal the normalized ul­
timate capacity pult/(Dσ’v) is insensitive to pile diameter, and the pult
value increases with depth at a constant rate. The trends of calculated
pult distributions using the three approaches above are in line with these
experimental observations, except API (2011)’s approach. The pult dis­
tributions proposed by Broms (1964) and Zhang et al. (2005) were both
attempted, in conjunction with the readily determined curve shape
function of p-y (Eq. (4)) and non-linear distribution function of kini (Eq.
(5)). It is found that the adoption of Zhang et al. (2005)’s distribution
function of pult has led to better predictions for the load-displacement
responses of the two piles reported in this study (evident in the
following sub-section). Thus, the following equation proposed by Zhang
et al. (2005) is employed to formulate distribution of pult for
large-diameter slender piles:

(6)

pult = KP2 γ zD

Where KP is the Rankine’s passive earth pressure coefficient, γ′ is


effective unit weight, z is the depth and D is the pile diameter.

4.4. Validation of the proposed p-y formulation for large-diameter slender


pile

This sub-section aims to verify the ability of proposed p-y approach in

Fig. 8. The initial stiffness of the p-y curve.

and Kim (2015), is included in the figure. In addition, the figure also
contains the calculated distributions of kini suggested by API (2011), by
Klinkvort and Hededal (2014) based on their large-diameter stubby pile
tests, and by Zhu et al. (2016) based on their small-diameter slender pile
tests.
As illustrated, the kini distributions of the two piles with different D (i.
e., 4 and 6 m) in this study fall within a narrow range, which increase
with depth but at a decreasing rate. No obvious “diameter effect” on the
kini distributions is evident from the two centrifuge test results reported
herein. At each given depth, the deduced kini values of these large-
diameter slender piles are generally smaller than those of large-
diameter stubby piles (Choo and Kim, 2015). Indeed, Choo and Kim
(2015) suggested a linear distribution of kini with depth for the
large-diameter stubby piles.
Compared to the measured kini distributions in this experimental
study, API (2011) significantly overestimates the kini values at all the
depths. Klinkvort and Hededal (2014)’s equation derived based on
large-diameter stubby piles well captures the measured kini values within
the top 10 m, but over-predicts the measurement at greater depth.
Comparatively, Zhu et al. (2016)’s equation proposed based on
small-diameter slender piles gives the best approximation for the
measured distributions of kini of the large-diameter slender piles re­
ported herein, although it tends to slightly under-predict the initial
lateral stiffness at great depth. The comparisons above appear to imply
the value of kini is more governed by pile rigidity, but relatively insen­
sitive to pile diameter.
Based on the centrifuge test results reported herein, the following
exponent function is found to best describe the distribution of kini for the
large-diameter slender piles:
( )0.7
′/
kini = 350 σv (5)
Patm Fig. 9. Typical distributions of ultimate soil resistance with depth.

7
H. Wang et al. Ocean Engineering 226 (2021) 108299

predicting various observed lateral behavior of the large-diameter


slender piles, including the load-displacement relation and bending
moment profile. This was achieved by implementing the proposed p-y
formulations into the BNWF model built in ABAQUS (Wang et al., 2020).
Fig. 10 compares the measured and computed load-displacement
relations of the two large-diameter slender piles (D = 4 and 6 m). The
computed results were obtained based on three sets of p-y formulations,
as proposed by this study, by Zhu et al. (2016) and by Klinkvort and
Hededal (2014). Among the three sets of selected p-y formulations, the
one proposed in this study yields the best predictions for both the initial
stiffness and the lateral capacity (i.e., lateral load at 10%D) of the two
piles with different diameters. The p-y formulation proposed based on
small-diameter slender piles (Zhu et al., 2016) offers closer approxi­
mation to the measured results, than that proposed on the basis of
large-diameter stubby piles (Klinkvort and Hededal, 2014). This may
suggest that the lateral behavior of monopile is more dependent on the
relatively pile-soil stiffness than the pile diameter.
The predictive capabilities of the three sets of p-y formulations (as
used in Fig. 10) for the measured bending moment profiles along the two
large-diameter slender piles at typical loading stages are evaluated, as
shown in Fig. 11. Despite the different formulations adopted in the three
p-y approaches, they all predict quite similar bending moment profiles
for the two piles at different loading stages, which show reasonable
agreements with the measured data It has been recognized that the
bending moment profile of laterally loaded pile is not sensitive to the
variable p-y curves (Hetenyi, 1946; Barton et al., 1983; Augustesen et al.,
2009).
Compared with the bending moment profile, the stiffness and ca­
pacity of laterally loaded pile, however, are significantly affected by the
p-y curves. It is evident from Figs. 6 and 10 that the p-y curves suggested
by API (2011) and Klinkvort and Hededal (2014) lead to
non-conservative predictions for the lateral pile stiffness and capacity,
while the p-y curves proposed by Zhu et al. (2016) and this study appear
to result in more satisfactory predictions. It is worth noting that the p-y
curves of Zhu et al. (2016) and this study were deduced from lateral
loading tests on small-diameter slender piles and large-diameter slender
piles, respectively. The similarity between the two may suggest the
lateral behavior of pile is more dependent on relative soil-pile stiffness,
rather than pile diameter It is also inferred by the two figures that the
laterally loaded large-diameter slender pile behaves more similarly to
small-diameter slender piles than the large-diameter stubby piles. In
other words, the lateral behavior of monopile is more dependent on the
relatively pile-soil stiffness than the pile diameter.

Fig. 11. Comparison of the bending moment profile between the measured and
computed by the modified p-y curve: (a) D = 4 m; (b) D = 6 m.

5. Lateral cyclic behavior of the large-diameter slender piles

5.1. Cumulative behavior of lateral displacement and rotation

Fig. 12(a) and (b) presents the measured cumulative lateral


displacement and rotation with number of loading cycles in a typical
large-diameter slender pile (D = 6 m), respectively. It is seen that both
the displacement and rotation accumulate with loading cycles but at a
decreasing rate, exhibiting a shakedown behavior for all the cyclic
amplitudes applied (up to 50% Fu). Despite the progressively stabilized
response of the pile under each episode of cycling, the cumulative
rotation remains to be a major concern. The allowable rotation limit
(0.25◦ ), as specified by DNV GL, 2014, only holds for the smallest cyclic
amplitude (10% Fu).
Fig. 10. Comparison of the force-deflection response at ground surface be­ Fig. 13(a)–(b) show, respectively, the normalized cumulative lateral
tween the measured and computed by the modified p-y curve.

8
H. Wang et al. Ocean Engineering 226 (2021) 108299

Fig. 12. Measured cumulative response of a typical monopile (D = 6 m) with Fig. 13. Normalized cumulative behavior of a typical pile (D = 6 m) with
loading cycles: (a) lateral displacement; (b) rotation. number of cycles.
(a) normalized lateral displacement; (b) normalized rotation.

displacement and rotation with number of loading cycles in a typical


episode, respectively. αy and αθ are cyclic accumulation indices that
large-diameter slender pile (D = 6 m). In the figure, each measured
displacement or rotation at an arbitrary Nth cycle is normalized with that gauge the rate of cumulative lateral displacement and rotation with
induced by the 1st cycle for a given load episode. The physical quantities number of cycles.
in both X-axis and Y-axis of the figure are presented in log-scale. It is These two cyclic accumulation indices (i.e., αy and αθ ) serve as the
seen that both the normalized lateral displacement and rotation in­ basis to compare the cyclic cumulative behavior of the large-diameter
creases with the log-scaled number of loading cycles in an approxi­ slender piles in medium dense sand, as reported herein, with the exist­
mately linear form. Similar observations have also been made by ing relevant database. For this purpose, six sets of databases were
Leblanc et al., 2010, Klinkvort (2012), Kirkwood (2016) and Truong collected, based on the test results in this study and those reported
et al. (2018) based on their model tests simulating laterally loaded piles elsewhere (Little and Briaud, 1988; Rosquoët et al., 2007; Klinkvort and
in sand. To quantify the rate of cyclic accumulation for a given pile, it is Hededal, 2013; Li et al., 2015; Truong et al., 2018). The tests involved in
a routine practice to fit the relationships in Fig. 13(a)–(b) with the power the six sets of databases were all related to cyclically loaded lateral piles
functions below (Truong et al., 2018): in sand, considering different values of relatively density of sand (Dr),
( ) pile diameter (D) and relative pile-soil stiffness (EPIP)/(EsL4). As pro­
posed by Poulos and Hull, 1989, the relative soil–pile stiffness can be
log yN/y = αy log(N) (7)
1 quantified by a nondimensional coefficient EPIP/EsL4 (EP IP , Es , and L
( ) denote the flexural rigidity of the pile, soil modulus, and embedded
log θN/θ = αθ log(N) (8) depth of the pile, respectively). Some key features of the tests involved in
1 these databases are summarized in Table 5.
Fig. 14(a) and (b) show the dependency of the cyclic accumulation
where y1 and yN denote cumulative lateral deflection at the 1st and the index αy on pile diameter and relative soil-pile stiffness, respectively. As
Nth cycle of a given loading episode, respectively. θ1 and θN denote shown in Fig. 14(a), the cyclic accumulation index αy exhibits no clear
cumulative rotation at the 1st and the Nth cycle of a given loading

9
H. Wang et al. Ocean Engineering 226 (2021) 108299

79 (Oztoprak and Bolton,

Note: Es ⋅Es is soil Young’s modulus. Apart from Rosquoët et al. (2007), Li et al. (2015) and Zhu et al. (2016), the QUOTE Es ⋅Es was estimated based on the suggested value in Oztoprak and Bolton (2013).
This study

Flexible
2013)
0.001
1106
60

65
6

79 (Oztoprak and Bolton, (a)


This study

Flexible
0.0003
2013)
311
60

65
4

82-99 (Oztoprak and Bolton,


Truong et al. (2018)

0.022–0.024
50–99

2013)
3.92

74.4

Stiff
24

(b)
13.5/20.4/27.7 (Sun,
Zhu et al. (2016)

0.47/7.46/56.66

0.0007/0.0005
0.75/1.5/2.5
15/30/50

/0.0003
Flexible
2016)
60
77.5 (Oztoprak and Bolton,
Choo and Kim (2015)

82–87

2013)
824.2

0.012

Stiff
31
6

102 (Oztoprak and Bolton,


Klinkvort and Hededal

Fig. 14. Dependency of the cyclic accumulation index αy on: (a) pile diameter
D; (b) relative soil-pile stiffness (EPIP)/(Esl4); (c) relative density of sand Dr.

dependency on the pile diameter. For example, strong divergence can be


observed for the piles with diameter of approximately 4 m. On the other
(2013)

2013)
0.048

hand, the values of cyclic accumulation index αy can be better correlated


Stiff
522
Key features of the tests involved in the collected database.

18

90
3

to the relative soil-pile stiffness (EPIP)/(EsL4) as shown in Fig. 14(b)). For


flexible piles (EPIP)/(EsL4)≤0.0025 (Poulos and Davis, 1980), the cyclic
42.5 (Yang et al.,

accumulation index αy lies within a narrow range between 0.04 and


Li et al. (2015)

0.08, while αy of the stiff piles falls within a broader range between 0.08
and 0.19. Fig. 14(b) also reveals that the αy values (αy =0.04–0.07) of the
2016)
0.038

0.038
0.34

large-diameter slender piles reported herein are close to those


Stiff
100
2.2

(αy =0.05–0.07) of the small-diameter slender piles reported by


Rosquoët et al. (2007), but much smaller than the αy values
(αy =0.08–0.22) of the large-diameter stubby piles (Klinkvort and
Rosquoët et al.

67.6 (Li et al.,

Hededal, 2013; Li et al., 2015; Truong et al., 2018).


0.00034

As far as the dependency of αy on Dr is concerned (see Fig. 14(c)), the


Flexible
(2007)

2014)
0.476

αy values of flexible piles with either large-diameter (this study) or


0.72
12

86

small-diameter (Rosquoët et al., 2007; Truong et al., 2018) fall within a


narrow range of 0.04–0.086, with no obvious dependency Dr. On the
(EPIP/EsL4)
(GNm2)

Pile type
Es (MPa)

other hand, the αy values of large-diameter rigid piles (Rosquoët et al.,


Table 5

D (m)
L (m)

2007; Truong et al., 2018) are more strongly dependent on Dr, i.e., less
EPIP

Dr

10
H. Wang et al. Ocean Engineering 226 (2021) 108299

cyclic accumulation (smaller αy ) for piles in denser sand (larger αy ). This consequence of resonance of the structure. These observed trends from
is probably attributed to the different failure mechanism between flex­ test in sand differ from that in clay, where the foundation stiffness de­
ible and rigid piles. The rigid piles experience rigid rotation under lateral creases under constant-amplitude cycling due to the soil stiffness
loading with the soil resistance along the whole length mobilized, when reduction accompanied by the cumulative excess pore pressure (Hong
the mobilized soil resistance of flexible piles is concentrated at shallow et al., 2017; Lai et al., 2019).
depth with little mobilized at deep, such that the former is more sensible To capture the evolution of unloading stiffness with loading cycles, a
to lateral soil resistance that increases with Dr. logarithm function is usually adopted:
Based on the readily collected database, as shown in Fig. 14, it may /
kθ, N k = 1 + μ ln(N) (9)
be concluded that the cyclic accumulation index αy of piles in medium θ,1

dense sand is mostly affected by relative pile-soil stiffness, while exhibit


little dependency on pile diameter. Consequently, the αy values of large- where kθ, N and kθ, 1 denote rotational stiffness at the 1st and the Nth cycle
diameter flexible piles are similar to those of small-diameter flexible of a given loading episode, respectively. μ is cyclic stiffness degradation
piles, but differs significantly from those of large-diameter stubby piles. index that characterizes the rate of rotational stiffness increase with
number of cycles.
The deduced ranges of stiffness degradation index μ from the tests on
5.2. Evolution of dynamic foundation stiffness
the large-diameter slender piles are illustrated in Fig. 16, accompanied
with results from the published database on large-diameter stubby piles
In addition to the cumulative deformation behavior, another major
(Leblanc et al., 2010; Klinkvort and Hededal, 2013; Li et al., 2015) and
concern for cyclically loaded monopile lies in dynamic foundation
small-diameter slender pile (Zhu et al., 2016). The abscissa of the figure
stiffness, due to the dynamically sensitive nature of the wind turbine
is the non-dimensional relatively soil-pile stiffness EPIP/EsL4, as defined
structure. The dynamic foundation stiffness may be decomposed into
by Poulos and Davis (1980).
lateral and rotation components. It is evident analytically that the dy­
As shown in Fig. 16, the rotational stiffness shows little dependency
namic behavior of an offshore wind turbine structure is mainly governed
on pile diameter with same range of 0.05–0.19 for the 4 m and 6 m
by the rotational stiffness, but is relatively insensitive to the lateral
diameter piles in this study. On the other hand, the values of rotational
stiffness (Arany et al., 2016). For example, Zania (2014) demonstrated
stiffness index μ are more governed by the relative soil-pile stiffness
that a 60% increase in rotational stiffness may lead to a 5–10% increase
(EPIP)/(EsL4).
in the Eigen-frequency of a wind turbine structure. Therefore, this
It can be seen that for the flexible piles (where EPIP/EsL4≤0.0025
sub-section is focused on the interpretation of rotational stiffness, which
(Poulos and Davis, 1980), including the large-diameter slender piles and
is calculated by taking the secant stiffness of each cycle of the
the small-diameter slender piles, stiffness evolution indices μ similarly
moment-rotation relationship.
lie within the range between 0.05 and 0.19, while smaller stiffness
Fig. 15 presents the development of rotational unloading stiffness
evolution indices (μ = 0.02 and 0.07) are resulted in the large-diameter
with number of cycles (in log scale) for a typical large-diameter flexible
stubby piles. These observations may suggest that the cyclic stiffness
pile (D = 6 m) subjected to the five loading episodes. In the figure, the
evolution could be significantly affected by relative soil-pile stiffness,
rotational stiffness at each arbitrary Nth cycle (kθ, N) is normalized by
while the “diameter effect” is relatively minor.
that at the 1st cycle (kθ, 1) of the corresponding loading episode. It is seen
In view of the scattered trends of the rotational stiffness data at
that the rational stiffness decreases with the increasing amplitude of
different cyclic amplitude, an attempt was made to generalize the trends
cyclic loading, as anticipated.
by re-plotting these data against the normalized residual rotation
For each given cyclic amplitude, however, the rotational stiffness
θResidual, N/θResidual, 1, as shown in Fig. 17. The symbols θResidual, N and
increases with the number of cycle (in log scale) following an approxi­
θResidual, 1 denote the residual rotations at the Nth and the 1st cycle of a
mately linear relationship. Because the stiffness of the sand around the
given pile under a corresponding loading episode, respectively. It can be
pile is likely to increase as the constant-amplitude cycling proceeds, due
seen that the normalized rotational stiffness kθ, N/kθ, 1 of the two
to the densification of the sand. The most significant stiffness increase
monopiles (D = 4 and 6 m) exhibits an approximately linear correlation
occurs at the lowest amplitude of cycling (10% Fu), with a maximum
to the normalized residual rotation θResidual, N/θResidual, 1, irrespective of
percentage increase of approximately 150%. This may shift an initially
the cyclic amplitude. This is probably because the residual pile rotation
soft-stiff designed wind turbine into the stiff zone, with potential
signifies the level of plastic soil strains developed around the pile, which
is related to the stiffness of the soil, and thus the rotational stiffness of
the pile. The encouraging correlation between the two quantities, as
evident in Fig. 17, may have suggested an alternative way to quantify
the evolution of pile stiffness through the cumulative residual deflection.

Fig. 15. Evolution of rotational unloading stiffness of a typical large-diameter Fig. 16. Dependency of the unloading stiffness index on relative soil-pile
flexible pile (D = 6 m) with cycle numbers. stiffness (EPIP)⁄(EsL4).

11
H. Wang et al. Ocean Engineering 226 (2021) 108299

Substantial lock-in bending moment was induced after each


unloading, which can account for up to 40% of the peak bending
moment developed during the preceding loading cycle. The presence of
lock-in moment after unloading (at zero lateral load) is a consequence of
plastic strains in the soil around the pile, as signified by the measured
residual lateral displacement and rotation (Fig. 12(a) and (b)). The
figure also shows that as the cycling proceeds, the locations of maximum
peak and lock-in moments gradually move downward. This implies a
downward load transfer resulting from the cyclic degradation of soil
resistance at shallow depth.
The difference between the maximum peak moment (MP,max) and
maximum lock-in moments (ML,max), which governs the fatigue of the
monopile, is further quantified by plotting the development of ML,max/
MP,max ratio during cycling at varying amplitudes, as shown in Fig. 19. It
is seen that at a given cycling number, ML,max/MP,max ratio generally
increases with cyclic amplitude. For each given cyclic amplitude, the ML,
max/MP,max ratio increases almost linearly with number of cycles (in log
scale), suggesting the lock-in moment accumulates faster than the peak
moment.
Fig. 17. Dependency of the unloading stiffness index μ on relative soil-pile
Considering that the lock-in bending moment is likely to be associ­
stiffness (EPIP)/(Esl4).
ated with the residual rotation of the pile, attempts were also made to
generalize the correlation between these two physically quantities in
5.3. Development of peak and lock-in bending moment
dimensionless forms. By trial-and-error, it is found that there exists a
unique correlation between a dimensionless form involving lock-in
The heavy instrumentation along each model pile has enabled a
moment (ML, max, N/MP, max, N)/(ML, max, 1/MP, max, 1) and the normal­
detailed examination of the development of peak moment and residual
ized residual rotation θResidual,N/θResidual, 1, as illustrated in Fig. 20. It is
moments, which jointly determines the fatigue damage that could have
worth noting that the fitted correlation shown in the figure is indepen­
been accumulated in the pile during the multi-stage cycling. The resid­
dent of pile diameter, cyclic amplitude and number of cycles. This may
ual moment is also alternatively defined as lock-in moment (Kirkwood
suggest the lock-in moment is inherently related to the level of plastic
and Haigh, 2014), which refers to the bending moment being ‘locked’
soil strain around the pile, which could be gauged by the residual
into the pile after unloading (i.e., zero lateral load at the pile head).
rotation at the pile head.
Fig. 18 shows the development of peak and lock-in bending moment
profiles of a typical large-diameter flexible pile (D = 6 m) subjected to
constant amplitude cycling (10% Fu). For clarity, the figure only in­ 6. Summary and conclusions
cludes the measured data at three typical loading cycles, namely the 1st,
100th and 1000th cycles. The calculated monotonic and cyclic bending This study presents a unique program of centrifuge model tests,
moment profiles based on the p-y approach suggested by API (2011) are aiming to offer fresh insight into the lateral monotonic and cyclic
also shown in the figure for comparison. behavior of large-diameter slender piles (D > 4 m and L/D > 10) in
It can be seen that both the peak bending moment of the pile in­ medium dense sand. Particular emphasizes are placed on the compara­
creases with number of loading cycles. This is consistent with the tive lateral behavior between the large-diameter slender piles reported
observed increase in rotation of the pile during constant amplitude herein, and those of the small-diameter slender piles (typically with D <
cycling (see Fig. 12(b)). The API (2011)’s p-y approach is shown to 3 m and L/D > 20) and large-diameter stubby piles (typically with D > 4
underestimate the peak bending moment induced by the 1st loading m and L/D < 8) published in the literature. Interpretation of the
cycle, and the percentage increase of the peak bending moment during centrifuge testing results has led to the following major conclusions:
the subsequent cycling. Because the API (2011)’s approach over­
estimates the stiffness of the p-y response.

Fig. 19. Development of the ratio between the maximum peak and lock-in
Fig. 18. Development of peak and lock-in bending moment profiles during bending moment (ML,max/MP,max) in a typical pile (D = 6 m) during cyclic
constant amplitude cycling at 10% Fu. loading at different amplitudes.

12
H. Wang et al. Ocean Engineering 226 (2021) 108299

editing, Funding acquisition. Yi Hong: Writing - original draft, Writing -


review & editing, Funding acquisition. Davis Mašín: Writing - review &
editing. Wei Li: Writing - review & editing, Funding acquisition. Ben
He: Writing - review & editing, Funding acquisition. Hualin Pan:
Writing - review & editing, Funding acquisition.

Declaration of competing interest

The authors declare that they have no known competing financial


interests or personal relationships that could have appeared to influence
the work reported in this paper.

Acknowledgement

The authors gratefully acknowledge the financial supports provided


by National Key Research and Development Program
(2018YFE0109500), National Natural Science Foundation of China
(51939010, 51779221 and 51909249), the Key Research and Develop­
ment Program of Zhejiang Province (2018C03031), Zhejiang Provincial
Fig. 20. Relationship between the normalized maximum locked-in bending Natural Science Foundation (LHZ20E090001 and LQ19E090001), and
moment (ML,max, N/MP,max, N)/(ML,max, 1/MP,max, 1) and normalized residual the research grant Inter-Excellence (Inter-Action) of the Czech Ministry
rotation (θResidual,N /θResidual,1 ).
of Education, Youth and Sports (LTACH19028).

1. The monotonic p-y load transfer behavior of the large-diameter


References
slender pile is marginally different from those of small-diameter
slender piles, but significantly deviate from the large-diameter Ahmed, S.S., Hawlader, B., 2016. Numerical analysis of large-diameter monopiles in
stubby piles. This may suggest the longstanding argument of dense sand supporting offshore wind turbines. Int. J. GeoMech. 16 (5), 04016018.
“diameter effect” is indeed relatively minor, while the lateral Alderlieste, E.A., 2011. Experimental Modelling of Lateral Loads on Large Diameter
Monopile Foundations in Sand. Delft University of Technology, the Netherlands. M.
behavior of monopile in sand is more significantly affected by the Sc Thesis.
relatively pile-soil stiffness. API American Petroleum Institute, 2011. Geotechnical and Foundation Design
2. The p-y curve recommended by API (2011) tends to offer Considerations. API RP 2GEO. API, Washington, DC.
Arany, L., Bhattacharya, S., Macdonald, J.H., Hogan, S.J., 2016. Closed form solution of
non-conservative predictions for the large-diameter slender piles in Eigen frequency of monopile supported offshore wind turbines in deeper waters
medium dense sand. Quantitatively, API (2011)’s p-y curve has incorporating stiffness of substructure and SSI. Soil Dynam. Earthq. Eng. 83, 18–32.
overestimated the initial stiffness and ultimate lateral capacity Augustesen, A.H., Brødbæk, K.T., Møller, M., Sørensen, S.P.H., Ibsen, L.B., Pedersen, T.S.,
Andersen, L., 2009. Numerical modelling of large-diameter steel piles at horns rev.
(corresponding to a lateral head displacement of 10%D) by 240% Proceedings of the Twelfth International Conference on Civil, Structural and
and 69%, respectively. A new p-y relation, with properly selected Environmental Engineering Computing, pp. 1–14.
forms for initial stiffness and lateral capacity, is therefore proposed Barton, Y.O., Finn, W.D.L., Parry, R.G.H., Towhata, I., 1983. Lateral pile response and p-y
curves from centrifuge tests. In: Proc. 15th Offshore Technol. Conf., Houston, Texas,
to specifically capture the lateral load-transfer behavior of
pp. 503–508.
large-diameter slender piles. Validations have been made against the Broms, B.B., 1964. Lateral resistance of piles in cohesionless soils. J. Soil Mech. Found
centrifuge testing results. Div. 90 (3), 123–158.
Byrne, B.W., Houlsby, G.T., 2003. Foundations for offshore wind turbines. Phil. Trans.
3. The large-diameter slender piles exhibit lower rates of cyclic
Roy. Soc. Lond.: Math. Phys. Eng. Sci. 361 (1813), 2909–2930.
displacement and rotation accumulation than those of the large- Byrne, B.W., McAdam, R., Burd, H.J., Houlsby, G.T., Martin, C.M., Gavin, K., Doherty, P.,
diameter stubby piles, but behave similarly as the small-diameter Igoe, D., Zdravković, L., Taborda, D.M.G., Potts, D.M., Jardine, R.J., Sideri, M.,
slender piles in these regards. This may suggest the cyclic behavior Schroeder, F.C., Muir Wood, A., Kallehave, D., Skov Gretlund, J., 2015. Field Testing
of Large Diameter Piles under Lateral Loading for Offshore Wind Applications. Proc
of laterally loaded piles in medium dense sand depends primarily on 16th European Conference on Soil Mechanics and Geotechnical Engineering
relative soil-pile stiffness, while the “diameter effect” is less (ECSMGE). Edinburgh, UK.
dominating. Choo, Y.W., Kim, D., 2015. Experimental development of the p-y relationship for large-
diameter offshore monopiles in sands: centrifuge tests. J. Geotech. Geoenviron. Eng.
4. Substantial lock-in bending moment was induced along the large- 142 (1), 04015058.
diameter slender pile, which can account for up to 40% of the peak Cohen, A., 2019. As Global Energy Demands Grows, So Does Appetite for Offshore Wind
bending moment developed within the same loading cycle. The ra­ accessed. https://www.forbes.com/sites/arielcoh
en/2019/03/26/as-global-energy-demands-grows-so-does-appetite-for-offsh
tios between the lock-in and peak moments, which determine the ore-wind/#3247073265e7. (Accessed 1 March 2020).
fatigue damage cumulated along the pile, exhibit a single trend of DNV GL, 2014. Design of Offshore Wind Turbine Structures (DNV-OS-J101). DNV GL,
correlation with the normalized residual rotation at the pile head, Oslo, Norway.
Doherty, P., Gavin, K., 2012. Laterally loaded monopile design for offshore wind farms.
irrespective of pile diameter, cyclic amplitude and number of cycles.
Proc. Ins. Civil Eng. Energy 165 (1), 7–17.
Georgiadis, M., Anagnostopoulos, C., Saflekou, S., 1992. Centrifugal testing of laterally
These centrifuge testing results form a unique database to support loaded piles in sand. Can. Geotech. J. 29 (2), 208–216.
Gilbert, R., Stokoe, K., Huang, Y., Munson, J., Bauer, J., Hosseini, R., Hussien, A., 2018.
development of new design methods for large-diameter slender piles,
Laboratory testing of lateral load response for monopiles in sand. BSEE/BOEM TAP
and to verify advanced numerical analyses involving cyclic models, 769.
which in turn would facilitate development of design methods for large- Haiderali, A.E., Madabhushi, G., 2016. Evaluation of curve fitting techniques in deriving
diameter slender piles. p–y curves for laterally loaded piles. Geotech. Geol. Eng. 34 (5), 1453–1473.
Hetenyi, M., 1946. Beams on Elastic Foundations. The University of Michigan Press, Ann
Arbor, Michigan, p. 94.
CRediT authorship contribution statement Hong, Y., He, B., Wang, L.Z., Wang, Z., Ng, C.W.W., Mašín, D., 2017. Cyclic lateral
response and failure mechanisms of semi-rigid pile in soft clay: centrifuge tests and
numerical modelling. Can. Geotech. J. 54 (6), 806–824.
Huan Wang: Conceptualization, Methodology, Writing - original Ishihara, K., 1993. Liquefaction and flow failure during earthquakes. Geotechnique 43
draft. Lizhong Wang: Project administration, Writing - review & (3), 351–451.

13
H. Wang et al. Ocean Engineering 226 (2021) 108299

Jeng, D.S., 2001. Numerical modeling for wave-seabed-pipe interaction in a non- Oztoprak, S., Bolton, M.D., 2013. Stiffness of sands through a laboratory test database.
homogeneous porous seabed. Soil Dynam. Earthq. Eng. 21 (8), 699–712. Geotechnique 63 (1), 54–70.
Kaynia, A.M., 2019. Seismic considerations in design of offshore wind turbines. Soil Poulos, H.G., Davis, E.H., 1980. Pile Foundation Analysis and Design. Wiley, New York.
Dynam. Earthq. Eng. 124, 399–407. Poulos, H., Hull, T., 1989. The role of analytical geomechanics in foundation
Kirkwood, P.B., Haigh, S.K., 2014. Centrifuge testing of monopiles subject to cyclic engineering. Foundation engineering: Current principles and practices. 2,
lateral loading. In: Proceedings of the 8th International Conference on Physical 1578–1606.
Modelling in Geotechnics 2014 (ICPMG2014), pp. 827–831. Perth, Australia. Qi, W.G., Gao, F.P., Randolph, M.F., Lehane, B.M., 2016. Scour effects on p–y curves for
Kirkwood, P.B., 2016. Cyclic Lateral Loading of Monopile Foundations in Sand. shallowly embedded piles in sand. Geotechnique 66 (8), 648–660.
University of Cambridge, UK. PhD. Thesis. Rosquoët, F., Thorel, L., Garnier, J., Canepa, Y., 2007. Lateral cyclic loading of sand-
Klinkvort, R.T., Hededal, O., 2013. Lateral response of monopile supporting an offshore installed piles. Soils Found. 47 (5), 821–832.
wind turbine. Proc. Inst. Civ. Eng. Geotech. Eng. 166 (2), 147–158. Sørensen, S.P.H., 2012. Soil-structure Interaction for Non-slender Large Diameter
Klinkvort, R.T., Hededal, O., 2014. Effect of load eccentricity and stress level on Offshore Monopoles. Ph.D. Thesis, Aalborg University, Denmark.
monopile support for offshore wind turbines. Can. Geotech. J. 51 (9), 966–974 Sun, Y.X., 2016. Experimental and Numerical Studies on a Laterally Loaded Monopile
(Chicago). Foundation of Offshore Wind Turbine. Ph.D. Thesis, Zhejiang University.
Klinkvort, R.T., 2012. Centrifuge Modelling of Drained Lateral Pile - Soil Response: Sun, X., Huang, D., Wu, G., 2012. The current state of offshore wind energy technology
Application for Offshore Wind Turbine Support Structures. Technical University of development. Energy 41 (1), 298–312.
Denmark, Denmark. PhD. Thesis. Systèmes, D., 2007. Abaqus Analysis User’s Manual. Simulia Corp. Providence, R.I., USA.
Klinkvort, R.T., Leth, C.T., Hededal, O., 2010. In: Springman, S., Laue, J., Seward, L. Taylor, R.N., 1995. Centrifuges in modelling: principles and scale effects. Geotechnical
(Eds.), Centrifuge Modelling of a Laterally Cyclic Loaded Pile. Physical Modelling in Centrifuge Technol. 19–33.
Geotechnics. CRC Press, London, pp. 959–964. The European Wind Energy Association (EWEA), 2020. Offshore Wind Europe Key
Lai, Y.Q., Wang, H., Wang, L.Z., Hong, Y., 2019. Experimental investigation on Trends and Statistics 2019 accessed. https://windeurope.org/about-wind/statisti
monotonic and cyclic lateral response of large diameter monopiles in sand and soft cs/offshore/european-offshore-wind-industry-key-trends-statistics-2019/. (Accessed
clay. A Report to Power China Huadong Engineering Limited Corporation. Zhejiang 1 March 2020).
University. Thieken, K., Achmus, M., Lemke, K., 2015. A new static p-y approach for piles with
Leblanc, C., Houlsby T., G., Byrne W., B., 2010. Response of stiff piles in sand to long- arbitrary dimensions in sand. Geotechnik 38 (4), 267–288.
term cyclic lateral loading. Géotechnique 60 (2), 79–90. Truong, P., Lehane, B.M., Zania, V., Klinkvort, R.T., 2018. Empirical approach based on
Leth, C.T., 2013. Improved Design Basis for Laterally Loaded Large Diameter Pile: centrifuge testing for cyclic deformations of laterally loaded piles in sand.
Experimental Based Approach. PhD Thesis. Aalborg University, Denmark. Geotechnique 69 (2), 133–145.
Li, Z., Kotronis, P., Escoffier, S., 2014. Numerical study of the 3D failure envelope of a Wang, X., Zeng, X., Li, J., Yang, X., 2018. Lateral bearing capacity of hybrid monopile-
single pile in sand. Comput. Geotech. 62, 11–26. friction wheel foundation for offshore wind turbines by centrifuge modelling. Ocean
Li, W., Igoe, D., Gavin, K., 2015. Field tests to investigate the cyclic response of Eng. 148, 182–192.
monopiles in sand. Proc. Institution of Civil Eng. Geotechnical Eng. 168 (5), Wang, X., Zeng, X., Li, X., Li, J., 2019. Investigation on offshore wind turbine with an
407–421. innovative hybrid monopile foundation: an experimental based study. Renew.
Li, W., Zhu, B., Yang, M., 2017. Static response of monopile to lateral load in Energy 132, 129–141.
overconsolidated dense sand. J. Geotech. Geoenviron. Eng. 143 (7), 04017026. Wang, H., Lehane, B.M., Bransby, M.F., Wang, L.Z., Hong, Y., 2020. A simple approach
Little, R.L., Briaud, J.L., 1988. Full Scale Cyclic Lateral Load Tests on Six Single Piles in for predicting the ultimate lateral capacity of a rigid pile in sand. Géotechnique
Sand. Miscellaneous Paper GL-88-27. Texas AM University College Station, Texas. Letters 10 (3), 429–435. https://doi.org/10.1680/jgele.20.00006.
Liu, B., Jeng, D.S., Ye, G.L., Yang, B., 2015. Laboratory study for pore pressures in sandy Wang, H., Wang, L.Z., Hong, Y., He, B., Zhu, R.H., 2020. Quantifying the influence of pile
deposit under wave loading. Ocean Eng. 106, 207–219. diameter on the load transfer curves of laterally loaded monopile in sand. Appl.
Markou, A.A., Kaynia, A.M., 2018. Nonlinear soil-pile interaction for offshore wind Ocean Res. 101, 102196.
turbines. Wind Energy 21 (7), 558–574. Winkel, J., 2016. Large Diameter Dolphin Piles: the Effect of the Inner Soil on Their Local
Negro, V., López-Gutiérrez, J.S., Esteban, M.D., Alberdi, P., Imaz, M., Serraclara, J.M., Buckling Resistance. M.Sc Thesis. Delft University of Technology, the Netherlands.
2017. Monopiles in offshore wind: preliminary estimate of main dimensions. Ocean Yang, M., Ge, B., Li, W., Zhu, B., 2016. Dimension effect on p-y model used for design of
Eng. 133, 253–261. laterally loaded piles. Procedia Eng. 143, 598–606.
Ng, C.W.W., Van Laak, P., Tang, W.H., Li, X.S., Zhang, L.M., 2001. The Hong Kong Zania, V., 2014. Natural vibration frequency and damping of slender structures founded
geotechnical centrifuge. Proceedings of the 3rd International Conference on Soft Soil on monopiles. Soil Dynam. Earthq. Eng. 59, 8–20.
Engineering, pp. 225–230. Zhang, L., Silva, F., Grismala, R., 2005. Ultimate lateral resistance to piles in cohesionless
Nordenham, Steelwind, 2020. Beyond XXL – Slim Monopiles for Deep-Water Wind Farms soils. J. Geotech. Geoenviron. Eng. 131 (1), 78–83.
accessed. https://www.offshorewind.biz/2020/05/11/beyond-xxl-slim-monopile Zhu, B., Li, T., Xiong, G., Liu, J.C., 2016. Centrifuge model tests on laterally loaded piles
s-for-deep-water-wind-farms/. (Accessed 16 August 2020). in sand. Int. J. Phys. Model. Geotech. 16 (4), 160–172.

14

You might also like