You are on page 1of 18

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/369698677

Cumulative cyclic response of offshore monopile in sands

Article  in  Applied Ocean Research · April 2023


DOI: 10.1016/j.apor.2023.103481

CITATIONS READS

0 103

5 authors, including:

Yilong Sun Mohamed Hesham El Naggar


Wenzhou University The University of Western Ontario
9 PUBLICATIONS   42 CITATIONS    425 PUBLICATIONS   8,186 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Performance Assessment and Design Guidelines for Three-Sided Precast Concrete Culverts View project

Investigation of Hybrid Foundation System for Offshore Wind Turbine View project

All content following this page was uploaded by Yilong Sun on 12 April 2023.

The user has requested enhancement of the downloaded file.


Applied Ocean Research 133 (2023) 103481

Contents lists available at ScienceDirect

Applied Ocean Research


journal homepage: www.elsevier.com/locate/apor

Cumulative cyclic response of offshore monopile in sands


Yilong Sun a, b, Chengshun Xu a, *, M. Hesham El Naggar c, Xiuli Du a, Pengfei Dou d
a
The Key Laboratory of Urban Security and Disaster Engineering of Ministry of Education, Beijing University of Technology, Beijing 100124, China
b
Key Laboratory of Engineering and Technology for Soft Soil Foundation and Tideland Reclamation of Zhejiang Province, Wenzhou University, Wenzhou 325035, China
c
Department of Civil and Environmental Engineering, Western University, London, N6A 3K7, Canada
d
Institute of Geotechnical Engineering, Tsinghua University, Beijing 100084, China

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

Keywords: Satisfactory performance of offshore wind turbines is governed by the cyclic response of its foundation under
Offshore monopile wind and wave. Most methods of evaluating the lateral cumulative deformation of monopile due to cyclic loading
Lateral cyclic loading are based on small-diameter pile tests installed in dry sand. This paper investigates the cumulative deformation
Cumulative deformation
of large-diameter monopiles installed in saturated sand. A numerical model was developed employing FLAC3D
Sand
and was validated by comparing its predictions with the results of triaxial tests and centrifuge tests. The validated
numerical model was used to evaluate and compare the cyclic responses of monopiles between saturated case
and no pore pressure case. To evaluate the cumulative deformation for offshore monopile due to cyclic loading,
the numerical model was refined to account for the effects of number and amplitude and frequency of cyclic
loading, soil permeability coefficient, soil relative density and pile diameter. It was found that the larger sub­
sidence and the less capacity of soil around the pile is caused by accumulation and dissipation of transient excess
pore pressure, thus the monopile in saturated sand would experience the larger cumulative deformation.
Correspondingly, based on this parametric study, an analytical model was developed to calculate the pile-top
cumulative displacement of offshore large-diameter monopile under lateral cyclic loading considering the ef­
fects of pile diameter. This analytical model was validated against the results of centrifuge tests and model pile
tests in saturated sands.

1. Introduction developed (Burd et al., 2020a; 2020b). In addition, some modified p-y
curve approaches were built by conducting numerical calculation and
The long-term performance of offshore monopiles is governed by its loaded pile tests (Thieken et al., 2015; Sørensen et al., 2010; Ashour,
response to environmental loads (i.e., wind and waves) that are cyclic in 2014). These improvements are in order to consider the effect of pile
nature. The number of load cycles applied on a monopile supporting diameter. However, these approaches do not properly account for the
offshore wind turbine may exceed 108 during its service life (Achmus variation of soil stiffness caused by cyclic loading. So, the cumulative
et al., 2009). This cyclic loading may induce cumulative deformation of deformation of monopile due to cyclic loading can’t been well obtained
the monopile, which will adversely impact the wind turbine operation. by these approaches.
Therefore, it is necessary to evaluate the monopile performance char­ Several studies investigated the cumulative lateral deformation of
acteristics under lateral cyclic loading. model piles under long-term lateral cyclic loading, and some analytical
Although the p-y curve approach is widely used to evaluate the models were developed to evaluate the lateral displacement and rotation
lateral response for large-diameter monopiles that are now employed for of the monopile head under lateral cyclic loading (Bienen et al., 2011;
offshore wind turbine applications (API, 2014), there is an awareness Klinkvort et al., 2012, 2013; Lin et al., 1999; Li et al., 2015). These
that p-y curve method may not properly provide the realistic behavior of analytical models consider explicitly the number of load cycles on the
offshore monopiles (Bouzid, 2018). It is principally that because p-y monopile response. For example, Long et al. (1994) proposed an
curve method was developed mostly based on lateral load tests of flex­ analytical model to evaluate the cyclic response of monopiles consid­
ible small diameter (≤ 610 mm) (Reese et al., 1974). To improve the ering the degradation of soil resistance, and reported that the cumula­
performance of the p-y method, the PISA design model has recently been tive deformation of monopiles installed in sand is strongly influenced by

* Corresponding author.
E-mail address: xuchengshun@bjut.edu.cn (C. Xu).

https://doi.org/10.1016/j.apor.2023.103481
Received 15 May 2022; Received in revised form 9 January 2023; Accepted 19 January 2023
Available online 17 February 2023
0141-1187/© 2023 Published by Elsevier Ltd.
Y. Sun et al. Applied Ocean Research 133 (2023) 103481

both load amplitude and soil relative density. Subsequently, Albiker effect of pile diameter on soil resistance. The developed model was
et al. (2017) developed a modified analytical model of the pile cumu­ validated by comparing its predictions with the observed development
lative deformation by the results of model pile tests. Richards et al. trend of pile cumulative displacement in centrifuge tests and model pile
(2021) analyzed the influence of stress level on the pile cumulative tests.
deformation, based on conducting the small-scale tests at 1 g and in the
centrifuge at 9 g and 80 g. However, these model pile load tests and 2. Validation of numerical model
analytical models have certain limitations that impacted their practical
value, especially the small diameter of the model piles and scaling to Proper selection of the parameters of the soil constitutive model and
prototype as well as carrying out the test in dry sand beds. Therefore, pile-soil interaction is critical for reliable evaluation of the monopile
their suitability for evaluating the cumulative deformation of offshore dynamic response. Therefore, these parameters and the soil constitutive
large diameter monopiles is questionable. model are calibrated to evaluate the performance of this developed
Numerical models were established to evaluate the cumulative de­ numerical model by comparing the results of numerical model with
formations of offshore monopiles. For example, Achmus et al. (2009) measured results from cyclic triaxial tests and centrifuge pile load tests.
established a sand degradation stiffness model (DSM) based on the re­
sults of drained cyclic triaxial tests on sand. Employing the DSM model, 2.1. Cyclic triaxial tests
Kuo et al. (2011) investigated the impact of embedded depth on the pile
performance. Meanwhile, Luo et al. (2018) and Yang et al. (2017) The SANISAND soil constitutive model is stress-ratio controlled,
conducted parametric studies and proposed some modifications to the critical state elastic-plastic sand model (Yang et al., 2020). In this model,
model to evaluate the cumulative pile deformation more accurately. the stress-ratio bounding surface, shown in Fig. 1, which simulates the
Chong et al. (2017; SH 2018; 2019) developed a semi-empirical nu­ soil kinematic hardening and is characterized by the lode angle, θ, the
merical model based on soil cumulative strain function, and used it to lode angle of the current back-stress ratio tensor, θα, and the image
analyze the cyclic response of monopiles. Their results revealed that back-stress ratio tensors, αbθ and αdθ . The model also adopts the state
most of the cumulative deformation occurs during the first few cycles of parameter ψ; thus, it can simulate the change in effective stress
loading (N<100). However, these constitutive models and correspond­ confinement and soil void ratio and consequently captures the response
ing numerical models were based on either drained tests or load tests on of sand in loose to dense states. The constitutive model accounts for
model piles also installed in dry sand. As such, the behavior of saturated changes in the sand fabric- dilatancy due to load reversal and simulates
two-phase seabed sand and its effect on pile-soil interaction are not sand contractive and dilative behaviors. The implementation of the
considered. Therefore, Cuéllar et al. (2011; 2014) developed a saturated model is simple through a set of material parameters that can be applied
two-phase medium numerical model based on u-pw model proposed by to different confining stresses and sand relative densities (Dafalias et al.,
Zienkiewicz et al. (1980) and reported that the transient cumulative 2004; Cheng et al., 2013; Verdugo et al., 1996). This constitutive model
pore pressure around the pile was significantly produced under extreme is incorporated in the FLAC3D model employed in the current study.
storm loading. In addition, the results of their model tests revealed that Table 1 presents the equations of the SANISAND soil constitutive model.
the long-term cyclic loading would induce an inverted cone-shaped Table 2
subsidence zone in the soil adjacent to the pile (Li et al., 2021). Li Ramirez et al. (2018) conducted a series of strain-controlled un­
et al. (2020) discussed the influence of the applied force and cyclic drained cyclic triaxial tests on Ottawa sand. The sand maximum and
loading frequency on the soil deformation process and proposed the minimum void ratios were 0.82 and 0.53 in these tests, and the relative
empirical equations for depicting the subsidence zone around the pile, density, Dr = 40% and 90%. The test samples were consolidated to a
based on a cyclic monopile experimental study. In addition, Damgaard confining pressure of 100 kPa, then subjected to an axial strain of 30%
et al. (2014) and Bayat et al. (2016; 2017) evaluated the effects of cyclic and a cyclic strain of 0.22%.
loading on soil in the vicinity of monopile employing a two-dimensional The above triaxial tests of Ottawa sand were simulated by employing
fluid-structure interaction model and reported that the transient pore the FLAC3D software and SANISAND constitutive model. This present
pressure caused by the cyclic loading would decrease the soil capacity in study is performed to compare with the results of these triaxial tests and
the vicinity of monopile. At the same time, Li et al. (2019) and Liu et al. the calculated results of Ramirez et al. (2018). It is because the rigid
(2022) and Barari et al. (2017) established 3D FEMs of offshore
monopiles installed in saturated sand and used this numerical model to
investigate the transient response of offshore monopiles under wind and
wave. They reported that there is a remarkable difference on the
development of the monopile deformation between in dry sand site and
saturated sand site.
The surveyed literature clearly indicates that the saturated two-
phase medium behavior of the saturated sand and densification of
sand in the vicinity of pile should be accounted for when evaluating the
long-term performance of monopiles supporting offshore wind turbines.
However, most of analytical models for the monopile cumulative
deformation is based on the test results of small pile in dry sands.
Therefore, this study evaluates the cyclic response of an offshore large
diameter monopile employing numerical models developed using the
program FLAC3D incorporating the SANISAND soil constitutive model
(Dafalias et al., 2004; Cheng et al., 2013), to simulate the sand dilatancy
and saturated seabed sand behavior, and hence account for the densi­
fication of soil around the monopile. Various analyses were then con­
ducted to reveal the difference in cumulative deformation of offshore
monopiles installed in saturated and no pore pressure cases. The effects
of soil relative density, pile-diameter and loading amplitude on the cu­ Fig. 1. Schematic of surfaces in the π plane (bounding surface (red); dilatancy
mulative deformation are evaluated to complement previous studies. surface (blue); critical state surface (green dash); maximum stress-ratio surface
Consequently, an analytical model was developed to account for the (black dash); yield surface (black circle)) (Yang et al., 2020).

2
Y. Sun et al. Applied Ocean Research 133 (2023) 103481

Table 1 of the initial stiffness, the peak deviatoric stress and the softening of sand
SANISAND constitutive model equations (Yang et al., 2020). as well as the developed excess pore pressure Δu in saturated sand
Description Equations Constants reasonably agreed with the experimental results. Here excess pore
pressure Δu is equal to pore water pressure under the cyclic loading
Elastic relations ε̇ev e
= ṗ/K; ė = ṡ/(2G)
Plastic relations
minus the lateral confining pressure. It is indicated that a set of material
ε̇pv = 〈L〉D; ėP = 〈L〉R

Hypoelastic G0 parameters can be applied to the different relative density.


G = G0 pat (2.97 − e) /(1 + e)(p/pat )1/2
2

K = 2(1 + v)G/[3(1 − 2v)] ν


√̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅ √̅̅̅̅̅̅̅̅
Yield surface f = (s − pα) : (s − pα) − 2/3pm = 0 m
2.2. Simulation of centrifuge pile load tests in dry sand
Dilatancy D = A0 (1 + 〈z : n〉)(αdθ − α) : n A0, ng
Deviatoric flow rule R = Bn − C[n2 − (1 /3)I]

Rosquoët et al. (2004) performed three centrifuge tests at 40 g on a


n = (r − α)/ ‖ r − α ‖
B = 1 + 3(1 − c)/(2c)g(θ, c)cos3θ c steel pile to investigate the cyclic lateral response of pile installed in
√̅̅̅̅̅̅̅̅
C = 3 3/2(1 − c)g(θ, c)/c dense Fontainebleau sand with Dr = 86%. Table 3 presents the soil
g(θ, c) = 2c/[(1 + c) − (1 − c)cos3θ] parameters. Considering the scaling factor of 40, the prototype pile
Kinematic hardening α̇ = 〈L〉(2 /3)h(αbθ − α) length, diameter and bending stiffness were 14.6 m and 0.72 m and
Fabric-dilatancy rate ż = − cz 〈L〉( − ε̇pv )(zmax n − z) cz, zmax 505MN/m2, respectively. The pile was embedded 12.0 m into the sand
Hardening coefficient h = b0 /[(α − αin ) : n]
and was subjected to cyclic lateral loading applied at eccentricity above
b0 = G0 h0 (1 − ch e)(p/pat )− 1/2 h0 , ch
√̅̅̅̅̅̅̅̅ ground surface, e = 1.6 m. The lateral cyclic load time histories of cases
Image point on DS αdθ = 2/3[g(θ, c)Mexp(nd ψ ) − m]n nd, M
Image point on BS
√̅̅̅̅̅̅̅̅
nb
P32 and P330 are shown in Fig. 4.
αbθ = 2/3[g(θ, c)Mexp( − nd ψ ) − m]n
Critical state line eref
The parameters of the constitutive model used in the numerical
c , λc, ξ
ref
b0 = ec − λc (p/pat )ξ
model were inferred based on the sand relative density, and most other
constitutive model parameters are default values in Table 4 for most
sands described in the literature 0. Pile-soil interaction in this numerical
Table 2 model is simulated by using interface element. The friction and cohesion
SANISAND constitutive model parameters for Ottawa sand (Ramirez et al.,
in the parameters of interface element are taken as 0.55 times the
2018).
strength parameters of adjacent soil, and normal stiffness and shear
Parameter Otawa sand stiffness of interface element are both specified as 10 times the stiffness
Non-dimensional elastic modulus constant/Gr 125(kPa) of the neighboring soil zone (Itasca, 2012).
Poisson’s ratio/υ 0.05 These centrifuge tests are simulated by employing the above
Critical-state stress ratio/M 1.26
modeling approach. In the tests and numerical models, the pile is free-
Ratio of Critical-state stress ratio in extension and compression/c 0.735
State line constant/λc 0.0287 headed. Fig. 5 compares the calculated and measured displacement of
Void ratio at p = 0/ec0 0.78 the pile-top under lateral cyclic loading. Reasonable agreement can be
State line constant/ξ 0.7 observed between the calculated and measured responses. In particular,
Yield surface constant/mm 0.02 the calculated cumulative pile deformation is in excellent agreement
h0 5
ch 0.968
with the measured response, which validates the numerical model
nb 0.6 developed in the current study. However, some differences are observed
A0 0.5 between the calculated and measured data, which is attributed to the
nd 0.5 lack of calibration of the constitutive model parameters for the Fontai­
zmax 11
nebleau sand.
cz 500
Cut-off factor to deal with low pressures/kcut 0.01
N 1
2.3. Simulation of centrifuge pile load tests in saturated sand

boundary of the numerical model is not exactly the same as the rubber Fuqiang et al., (2011) conducted the centrifuge tests in a saturated
boundary of soil tests. The boundary displacement of numerical model is sand prepared from Fujian standard sand. Maximum void ratio and
fixed, but some small boundary displacement is inevitably produced in minimum void ratio are 0.848 and 0.519, respectively. The dry density
the soil tests. So there are some differences between the calculated and of the sand was 1.6 g/cm3, and its relative density was about 60%. The
measured test results in Figs. 2 and 3. Overall, the numerical predictions soil permeability is 2.1 × 10− 5 m/s. Diameter of the model pile was 50
mm and its length was 420 mm. With the scaling factor of 50 g, the

Fig. 2. Comparison of calculated and measured results of cyclic undrained triaxial tests of Ottawa sand with Dr=40%.

3
Y. Sun et al. Applied Ocean Research 133 (2023) 103481

Fig. 3. Comparison of calculated and measured results of cyclic undrained triaxial tests of Ottawa sand with Dr=90%.

Table 3 Table 4
Parameters of soil used in centrifuge tests (Gerolymos et al., 2009; Gian­ SANISAND constitutive model parameters of this numerical model.
nakos et al., 2012). Parameter value
Parameter Sand
Non-dimensional elastic modulus constant/Gr 125(kPa)
√̅̅̅
Strain shear modulus/(kPa), z soil depth/m 30 z Poisson’s ratio/υ 0.05
Friction angle/(◦ ) 41.8 Critical-state stress ratio/M 1.25
Critical-state angle/(◦ ) 33 Ratio of Critical-state stress ratio in extension and compression/c 0.712
Relative density/(%) 86 State line constant/λc 0.019
Mass density/(g/cm3) 1.63 Void ratio at p = 0/ec0 0.934
State line constant/ξ 0.7
Yield surface constant/mm 0.01
h0 7.05
ch 0.968
nb 1.25
A0 0.704
nd 2.1
zmax 2.0
cz 600
Cut-off factor to deal with low pressures/kcut 0.01
N 1

numerical results and the test results is to embody the application of


numerical model. The comparisons of numerical results and test results
are shown in Fig 7. Fig 7(a) shows the comparison of excess pore pres­
sure Δu for PPT1, which is a 2 m buried depth point and 1 m away from
the pile. Here Δu is equal to pore water pressure under cyclic loading
minus the lateral consolidation pressure. Fig. 7(b) shows the comparison
of pile-top lateral peak displacement under cyclic loading between test
data and numerical results. It can be seen that the numerical results are a
little more than the test data, but on the whole the calculated excess pore
Fig. 4. Cyclic load time histories of cases P32 and P330 (Giannakos pressure and calculated cumulative deformation are basically consistent
et al., 2012). with the test results. The difference of pile-top lateral displacement in­
creases with increasing the number of cyclic loading. It is due to the
prototype diameter was 2.5 m and the embedded depth was 16 m, and SANISAND constitutive may overestimate the displacement with
the pile segment above ground was 5 m. Lateral cyclic loading with increasing the number of cyclic loading. But the maximum difference is
frequency = 1 Hz and an amplitude of 170 N was applied to the pile top less than 18%. It is indicated that the developed numerical model can
as shown in Fig. 6. To clearly present the cyclic loading, Fig. 6 shows the basically obtain the pore pressure response and the pile response under
25 s of cyclic loading. Viscous fluid prepared by HPMC powder was used cyclic loading.
in this test. The prototype cyclic loading frequency was 0.02 Hz and its
amplitude was 425 kN. 3. Comparison of accumulated deformation for monopile in
Most of parameters for constitutive model are default values in saturated and no pore pressure cases
Table 4. The void ratio of sand in this numerical model is input and
calculated according to the relatie density and maximum void ratio and To evaluate the influence of pore water pressure in saturated sand on
minimum void ratio. The interface element is also used to simulate pile- the deformation of monopile, this verified numerical model is used in
soil interaction in this numerical model. the analysis of monopiles installed in no pore pressure (with effective
Pile response caused by accumulation and dissipation of excess pore unit weight) and saturated cases. the results for the saturated and no
pressure is time-dependent (Ong et al., 2006; D.E.L 2018). The com­ pore pressure cases are compared in terms of the monopile cumulative
parison between the excess pore pressure and pile-top displacement of deformation at its head, deflection along its shaft and the subsidence of

4
Y. Sun et al. Applied Ocean Research 133 (2023) 103481

Fig. 5. Comparison of pile-top displacement and lateral loading between test results and numerical model (Giannakos et al., 2012).

tests. In addition, the seawater level is considered to be 30.0 m above


ground level (mudline). The soil domain is 80 m in length and 50 m in
depth. Mesh size sensitivity was conducted to ensure accurate results
and acceptable computational efficiency. Fig. 8 presents an overview of
different size mesh models: mesh 1 (mesh size 2 m), mesh 2 (radiative
mesh size ratio 1.1), mesh 3 (mesh size 3 m), mesh 4 (mesh size 4 m).
Details of numerical model with the different mesh size are summarized
in Table 6. The lateral cyclic loading is applied to the monopile head as
shown in Fig. 9. At the first positive peak of cyclic loading, the lateral
deformation of the different mesh size numerical models is compared in
Fig. 8. It can be seen that the results from models with mesh 1 and mesh
2 and mesh 3 and mesh 4 are basically identical. The errors of mesh 1
and mesh 2 are less than 4%. In addition, when the static lateral force is
applied to the monopile head, the response of monopile obtained from
the different meshes are compared in Fig. 10. The max error of mesh 2
with mesh 3 and mesh 1 is 5%. Thus mesh 2 is sufficiently accurate,
while providing acceptable computational efficiency.
The offshore monopile is subjected to frequent wave and wind loads,
Fig. 6. Lateral cyclic loading of model pile (Wang et al., 2011). which act at low frequency (about 0.1 Hz) (Arany et al., 2017).

soil adjacent to the monopiles as well as the produced pore water Table 5
pressure. Parameters for steel pipe pile.
Overall Embed Diameter Wall Poisson Elastic
3.1. Numerical model of offshore monopile length /m depth /m /m thickness ratio modulus
/m /GPa

The parameters of SANISAND model are shown in Table 4, while 60 30 7.5 0.09 0.25 210
Table 5 presents the properties of the steel pile used in the centrifuge

Fig. 7. Comparison of excess pore pressure and pile-top displacement between the tests and numerical model.

5
Y. Sun et al. Applied Ocean Research 133 (2023) 103481

Fig. 8. Mesh sensitivity analysis: lateral deformation of different mesh size at the first positive load peak.

properly consider the effect of developed excess pore water pressure,


Table 6
such as the degradation soil model (Achmus et al., 2009; Luo et al.,
Details of numerical model with the different mesh size.
2018; Yang et al. 2017) or semi-empirical model (Chong et al., 2017; SH
No Gird type Number of elements Number of nodes 2018; 2019). Therefore, the developed numerical model is employed to
Mesh 1 Uniform mesh 642,294 625,900 investigate the cumulative deformation of offshore monopiles installed
Mesh 2 Radiative mesh 20,160 19,558 in saturated sand for both with and without transient excess pore pres­
Mesh 3 Uniform mesh 50,860 47,024
sure, and the results are compared to evaluate the effect of transient
Mesh 4 Uniform mesh 23,820 21,552
excess pore pressure on the calculated monopile response. The effective
unit weight 9.76 kN/m3 of sand is used in the numerical model of no
pore pressure case to consider the buoyancy effect. In the numerical
model of saturated case (considering pore pressure), the saturated unit
weight of sand is 19.76 kN/m3 and fluid-structure interaction are
considered. Both sand profiles have a relative density of 55% and the
hydraulic permeability of the medium dense saturated sand was set in
this study to 4 × 10− 5 m/s (Cuéllar et al., 2011). In both numerical
models, the cyclic loading is the same to capture the same wave loading.

3.2. Discussion of cumulative monopile deformation differences in


saturated and no pore pressure cases

The results of the comparative analyses are elaborated to explore the


difference in cumulative deformation of monopiles in saturated and no
pore pressure cases. Both cases are the same in stress sate, boundary
state and load. The different point is that no pore pressure case doesn’t
include drainage effect (accumulation and dissipation of excess pore
pressure) and deformation due to drainage. In particular, the pattern of
Fig. 9. Schematic diagram of applied cyclic lateral loading. pile deflection, subsidence depth of the soil around the pile, the devel­
opment of transient pore pressure in the soil along the pile shaft and the
cumulative displacement of pile-top.
Therefore, the frequency of cyclic loading considered in the analysis is
0.1 Hz. Based on the design procedure proposed by Arany et al. (2017)
3.2.1. Cumulative pile-top deformation
and the study by Bhattacharya et al. (2017), the amplitude of the cyclic
The calculated responses are normalized to explore the variation of
lateral loading for 7.5 m monopile is determined. While the water depth
the cumulative pile-top deformation with the number of load cycles.
is 30 m, the amplitude of cyclic loading applied to the 7.5 m pile head is
Fig. 12 compares the peak lateral displacement of pile head yN and the
about 5MN to represent wave loads. The total loading time considered in
normalized lateral displacement yyN0 (where y0 and yN are the pile lateral
the analysis is 2000s, corresponding to 200 load cycles. Fig. 9 displays
the time history of the applied lateral cyclic load, in which the dashed displacement after the first and N cycles, respectively) for the monopile
line means that the cyclic loading continues for a total of 2000s. Fig. 11 installed in saturated case and no pore pressure case.
shows the details of this numerical model. Fig. 12(a) shows that the lateral displacement both increases with
These existing analytical methods of offshore monopile cumulative increasing the number of cyclic loading in no pore pressure and satu­
deformation only consider the submerged unit weight but do not rated cases. To compared the development of cumulative displacement,
the normalized lateral displacement of the pile head is shown in Fig. 12

6
Y. Sun et al. Applied Ocean Research 133 (2023) 103481

Fig. 10. Comparison of response of the pile under different mesh size ((a)Displacement of pile at seabed line, (b)Deflection of pile (16MN), (c) Bending moment of
pile (16MN)).

Fig. 11. Monopile numerical calculation model ((a)Lateral view, (b)Front view).

Fig. 12. Comparison of normalized pile-top lateral displacement in saturated and no pore pressure cases.

(b). It is demonstrated from Fig. 12(b) that the normalized displacement both saturated and no pore pressure cases but the rate of increase in
in no pore pressure case increases gradually with the number of load saturated case is larger than that of the case for monopile in no pore
cycles. Meanwhile, the normalized displacement in saturated case in­ pressure case. And significant growth of cumulative displacement occurs
creases at a greater rate with the number of load cycles than the no pore during the first load cycles (N<100). This is also reported by Chong
pressure case. The rate of increase in pile head displacement both de­ (2017). These results clearly indicate that the lateral resistance of the
creases slightly after 20 cycles of loading, but the normalized displace­ saturated sand decreases due the development of transient pore
ment in saturated case continues to increase up to 1.88 and the pressure.
normalized displacement in no pore pressure case increases up to 1.35 Based on principle of effective stress, deformation of saturated sand
after 200 load cycles. This represents an increase of 39% compared to included two parts. One is deformation due to seepage, and the other is
the response of the monopile in the no pore pressure case due to the deformation due to loading. In both case, the same soil initial stress state
effect of transient pore pressure. Fig. 12(b) also shows that the upward is considered to obtain the same buoyancy effect. But in the no pore
trend of the normalized displacement continues after 100 load cycles in pressure case, the seepage effect is not included. In other word, the

7
Y. Sun et al. Applied Ocean Research 133 (2023) 103481

deformation due to the accumulation and dissipation of water pore Table 7


pressure is not caused. So the deformation in the no pore pressure case is Comparison of parameters for accumulated displacement prediction model.
less than the saturated case. Influence parameter Influence parameter of cyclic
Based on the numerical results, the difference of predicting the pile- Tb loading c
top cumulative displacement by using the existing analytical models is No pore pressure 0.045 0.42
investigated. Luo et al. (2018) and Yang et al. (2017) evaluated the case
existing analytical models for predicting cumulative monopile de­ Saturated case 0.066 0.51
formations based on the degradation stiffness model (DSM) and pointed
out that the power function of the number of load cycles can predict well
3.2.2. Deformation of the pile at the different depth
the cumulative deformation. Similarly, Lai et al. (2020) indicated that
Figure 14 compares the lateral displacement of the different loaded
the power function of number of load cycles can predict well the pile
number along the pile shaft. It is observed from Fig. 14 that most of the
cumulative deformation by validating it against centrifuge model pile
displacement at the pile head is due its rotation in addition to some
tests. Therefore, the power function proposed by Leblanc et al. (2010), i.
flexural deflection in both saturated and no pore pressure cases. The
e., Eq. (1), is used to elucidate the difference of the predicted cumulative
depth of rotation point is about 22 m in both cases and remains basically
deformation in saturated and no pore pressure cases. The considered
unchanged as the number of load cycles increases. However, the amount
loading is one-way cyclic loading applied at the pile-top, so Tc is 1. The
of rotation is larger in the saturated case and the cumulative pile lateral
tilt angle of the pile head is converted into lateral displacement, i.e.:
displacement increases as the number of load cycles increases. In addi­
Δθ(N) tion, for the same number of load cycles, the cumulative lateral
= Tb Tc N c (1)
θ0 displacement for the monopile in saturated case is greater than that for
the monopile in no pore pressure case due to the reduced stiffness of
yN = y0 (1 + Tb N c ) (2) saturated sand along the pile shaft. The significant increase in lateral
displacement for the monopile in saturated sand indicates significant
where N is the number of load cycles; Δθ(N) is the increment of pile tilt degradation of soil stiffness, which is attributed to the increase in pore
angle (rad) after N cycles; θ0 is the pile tilt angle (rad) of the pile after the water pressure due to the cyclic loading as will be discussed later.
first cycle; Tb and Tc are the influence parameters; only one-way cyclic
loading is considered in this study, so Tc is 1; c is the influence index of 3.2.3. Subsidence depth of soil around pile
the number of load cycles. The comparison of soil subsidence contour around the pile is shown
To present the difference of the predicted model applied to calculate in Fig. 15. which demonstrates that soil subsidence occurs in the vicinity
the cumulative deformation between the saturated and no pore pressure of the pile in both saturated and no pore pressure cases, and the subsi­
cases, the cumulative pile-top displacement values calculated from the dence depth and influence region in saturated sand are more than the
numerical models are curve fitted into Eq. (2) to establish the correlation results of no pore pressure case. It is due to the bearing capacity of
parameters, Tb and c, and the results are presented in Table 5. Fig. 13 saturated sand is low than the no pore pressure case. In addition, the
compares the normalized pile-top displacement obtained from the nu­ subsidence zone of numerical results is similar to the study of Cuéllar 0.
merical models with those obtained employing Eq. (2) with the curve The reasonability of numerical model is indicated again.
fitting parameters. Inspecting the curve fitting parameters presented in Figure 16 compares the maximum depth of soil subsidence around
Table 7, it is clear that the parameters are quite different for the no pore the pile, which demonstrates that soil subsidence occurs in both satu­
pressure and saturated cases. This is attributed to the effect of the rated and no pore pressure cases. However, the depth of subsidence in
transient excess pore pressure generated in the vicinity of the pile during saturated sand is more than twice that of the no pore pressure case. With
the cyclic loading, which further confirms that the effect of transient 100 cycles, the subsidence depth of the soil around pile is about 0.03 m
pore pressure should be considered in calculating the cumulative in no pore pressure case and 0.065 m in saturated case. This is attributed
displacement of offshore monopiles. to the fact that the lateral cyclic loading leads to shrinkage of the soil
Because significant growth of cumulative displacement occurs dur­ around the pile. In addition, the cumulative and dissipation of the
ing the early number of cycle (N<100), the following comparisons are transient pore pressure causes the progressive subsidence of the soil
focused on the early number of cycle (N<100). around the pile in saturated sand.

Fig. 13. Comparison of fitting values for normalized pile-top lateral displace­ Fig. 14. Comparison of the lateral displacement of the different loaded number
ment in no pore pressure and saturated cases. along the pile shaft.

8
Y. Sun et al. Applied Ocean Research 133 (2023) 103481

Fig. 15. Comparison of the subsidence contour of soil around pile.

pore pressure for − 1 m is low than the others. It is due to the soil strain of
shallow zone is large, and the point of − 1 m is near the mudline so the
excess pore pressure may easily dissipate.
To further explore the influence of transient excess pore pressure, the
pore pressure ratio is calculated. The pore pressure ratio is the value of
excess pore water pressure divided by the lateral effective consolidation
stress of soil. Fig. 19 displays the time history of the transient excess pore
pressure ratio at depths of − 1 m, − 4 m, − 8 m, − 16 m, − 20 m and − 28 m
after 20 load cycles. It is observed from Fig. 19 that the transient excess
pore pressure ratios at depths of − 1 m and − 4 m are higher than at other
depths. The peak value of transient excess pore pressure ratio is about
0.45. This is because soil deformations at depths of − 1 m and − 4 m are
larger than at other depths. It is indicated that the soil resistance drops
significantly. At the buried depth of − 8 m, the peak value of transient
excess pore pressure ratio is about 0.2, so there is a slight reduction in
the soil resistance. In addition, the deformation of offshore monopile
under lateral loading is primarily a rigid rotation (Ahmed et al., 2016;
Georgiadis et al., 1992; Prasad et al., 1999; Sun et al., 2020). In this
study the rotation point of the pile is found to be about − 22 m. The
Fig. 16. Comparison of subsidence depth of soil around pile.
deformation of soil near and below the rotation point is small. But the
transient excess pore pressure at − 28 m is more fluctuation than the
3.2.4. Pore-pressure at the different depths point of − 20 m. From Fig. 14, it is due to the point of − 28 m is near the
The variation of pore pressure along the pile shaft is evaluated. pile end and the lateral deformation of this point is more than the point
Fig. 17 compares the pore pressure before and after the application of of − 20 m. Overall, the transient excess pore pressure ratio at the
lateral cyclic loading. It is obvious from Fig. 17 that transient excess pore embedded depths of − 16 m, − 20 m, and − 28 m is basically unchanged,
pressure develops near the pile. These monitoring points of transient about zero. That is, there is no reduction in the soil resistance at these
pore pressure are 1 m away from the monopile. The history curves of the locations.
transient excess pore pressure for the shallow zone around the pile (− 1 In summary, accumulative and dissipation of transient excess pore
m, − 4 m, − 8 m) are shown in Fig. 18. It is observed that the transient pressure in the shallow soil around the pile is especially noticeable,
pore pressure fluctuates significantly, but the peak of transient excess which reduces the resistance of shallow soil layer around the pile and

Fig. 17. Comparison of pore pressure contour between before and after the application of cyclic loading.

9
Y. Sun et al. Applied Ocean Research 133 (2023) 103481

Fig. 18. History curves of transient excess pore pressure for shallow zone around the pile.

Fig. 19. History curves of transient excess pore pressure ratio for different depth around the pile.

causes progressive subsidence of the soil in the immediate vicinity of the 4.1. Pile diameter
pile. Thus, the subsidence in saturated sand is more significant than that
of no pore pressure case. Finally, increasing pore pressure and soil The effect of pile diameter on the cumulative pile-top displacement is
subsidence due to transient pore pressure both lead to reduced resis­ investigated in this section by varying the pile diameter D, from 2.0 m to
tance of saturated sand and hence increases the pile lateral deformation. 10.0 m. In all cases, the soil is medium-dense saturated sand with Dr =
Therefore, it is necessary to consider the soil saturated state and 55%, and the pile embedded depth and its part above the seabed are 4
development of transient excess pore pressure in the analysis of offshore times its diameter.
monopiles deformation. Ahmed et al. (2016) and Luo et al. (2018) suggested that the ultimate
lateral capacity criterion for offshore monopile to be based on the
4. Parametric studies pile-top displacement or rotation. In this study, the pile ultimate lateral
capacity Fu, is taken as the force applied at the pile-top that produces pile
The above indicated the effect of transient excess pore pressure on lateral displacement of 0.1D at mudline. The variation of the calculated
the cumulative deformation is remarkable. Based on the numerical ultimate lateral load capacity is plotted versus the monopile diameter in
model of saturated sand, in this section a series of numerical simulations Fig. 20. Using the ultimate lateral capacity Fu, the lateral cyclic load
is conducted to calculate the pile-top cumulative displacement of amplitude can be normalized as:
offshore monopiles considering some important parameters including:
Fmax
pile diameter D, soil relative density Dr, soil permeability, loading fre­ ξL = (3)
Fu
quency and amplitude of cyclic loading. Correspondingly, the effect of
these important parameters on the correlation parameters of the existing where ξL is the normalized cyclic load amplitude; Fmax is the cyclic load
analytical model is explored. amplitude.

10
Y. Sun et al. Applied Ocean Research 133 (2023) 103481

The parameter Tb is given as a function of monopile diameter, i.e.:


D
Tb = 0.005 × + 0.06 (4)
Dref

4.2. Soil relative density

The soil relative density is a crucial parameter in simulating its


behavior in this numerical model. In order to investigate its effect on the
cumulative pile-top displacement, carious analyses are conducted
considering Dr = 50%, 55%, 70%, 80% and 90%. Pile diameter is 7.5 m
and the normalized cyclic load amplitude ξL is 0.17 (4.1MN, 5.0MN,
5.8MN, 6.5MN and 7.2MN). The calculated normalized cumulative pile-
top displacements are presented in Fig. 23 vs the number of load cycles.
From Fig. 23 it can also be seen that the rate of cumulative deformation
decreases with increasing the number of cyclic loadings, and the cu­
mulative pile-top deformation decreases with increasing the relative
density. The reason is that the capacity of soil increases as increasing the
soil relative density. In addition, the results are curve fitted into Eq. (2)
Fig. 20. Ultimate lateral capacity of offshore monopile under different to establish its correlation parameters, and the predictions of Eq. (2) are
pile diameter.
then presented in Fig. 24 for comparison. The curve fitting showed that
the influence parameter Tb is always about 0.066 for all Dr values
Based on the foregoing numerical results, the normalized cyclic load considered in this study. However, the influence parameter c is found to
amplitude ξL is taken as 0.25 in this section (0.2MN, 1.4MN, 4.1MN, closely correlated with Dr as shown in Fig. 24. The relationship between
7.5MN, 9.1MN and 16.2MN). This cyclic load is then used to determine c and Dr is obtained through curve fitting the data in Fig. 20 and is given
the normalized cumulative pile-top displacement for different piles and by:
the variation of normalized cumulative displacement is plotted versus
number of load cycles for different pile diameter as shown in Fig. 21. As c = − 0.0015 × Dr + 0.6 (5)
observed in Fig. 21, the rates of cumulative deformation for different This linear inverse correlation between c and Dr is because the soil
diameter monopile all also decreases with increasing the number of resistance increases as its relative density increases, which reduces the
cyclic loading, and in the case of the same normalized cyclic load cumulative pile-top displacement, hence c decreases.
amplitude ξL the cumulative pile-top deformation increases as the pile The subsidence depth of the different soil relative density is shown in
diameter increases. It is due to that the amplitude of cyclic loading for Fig. 25. It can be seen that the subsidence depth increases with
the large-diameter monopile is more than the small-diameter monopile. increasing the time of cyclic loading, but the subsidence depth decreases
It is indicated that pile diameter has a remarkable effect on the cumu­ with increasing the time of cyclic loading. It is due to the capacity of the
lative deformation of laterally loaded monopile. soil increases and the soil void ratio decreases, with increasing the soil
To consider the effect of pile diameter on the cumulative deforma­ relative density. It is indicated that the subsidence of the soil around the
tion in the analytical model, the normalized pile-top displacement pile is one of the reasons that cause the cumulative deformation greater.
curves shown in Fig. 21 are fitted into Eq. (2) to back-calculate the pa­
rameters Tb and c for different pile diameters. It is found that c is almost
4.3. Sand permeability
around 0.517 for all D values considered herein (which covers the range
of pile diameter values used in offshore wind turbines). The parameter
In this section, the effect of soil permeability on the pore pressure and
Tb increases linearly with increasing the monopile diameter as shown in
the cumulative deformation of pile is investigated. This study is mainly
Fig. 22 (reference pile diameter Dref = 1 m). This is also because the
in medium dense and dense sand. So the variation of soil permeability
applied loading amplitude increases with increasing the pile diameter.
ranges between 2 × 10− 5 m/s and 8 × 10− 5 m/s. It is assumed that the

Fig. 21. Normalized pile-top lateral displacement for different pile diameter. Fig. 22. Relation of influence parameter Tb and normalized pile diameter.

11
Y. Sun et al. Applied Ocean Research 133 (2023) 103481

permeability is 1 × 10− 1 m/s.


The transient pore pressure and pile-top displacement is shown in
Figs. 26 and 27. It is presented from Fig. 26 that when the soil perme­
ability varies from 2 × 10− 5 m/s to 8 × 10− 5 m/s, the development law
of transient pore pressure has no noticeable change. When the soil
permeability is 1 × 10− 1 m/s, the peak value of transient pore pressure
is about 20% of other peak for transient pore pressure. The same pattern
is found from Fig. 27. when the soil permeability varies from 2 × 10− 5
m/s to 8 × 10− 5 m/s, the variation trend of cumulative deformation is
almost the same.
To investigate the reason, soil settlement of soil in the vicinity of the
pile is compared, when the soil permeability is from 2 × 10− 5 m/s to 8
× 10− 5 m/s and 1 × 10− 1 m/s. It is found from Fig. 28 that the soil
settlement is minimum in the limit drainage case, but in other cases the
soil settlement is about the same. It is due to the transient excess pore
pressure of limit drainage (1 × 10− 1 m/s) accumulates less than others.
Thus, when the sand permeability is 1 × 10− 1 m/s, soil settlement due to
transient excess pore pressure is less than others. When the soil
Fig. 23. Curves of normalized pile-top lateral displacement for different soil permeability increases, the accumulative deformation of pile and soil
relative density.
deformation decreases. It is also indicated that the cumulative and
dissipation of transient pore pressure decreases the capacity of sand
around the pile. But for medium dense sand and dense sand (50%− 90%)
of this study, the effect of soil permeability on the cumulative defor­
mation for offshore monopile is basically unchanged. The reason is that
in this study the amplitude annd frequency of cyclic loading is small.

4.4. Frequency of cyclic loading

The effect of frequency of cyclic loading on the cumulative defor­


mation for monopile is investigated in the section. Wave and wind is the
cyclic loading of low frequency, so the frequency of cyclic loading
applied to the pile ranges from 0.01 Hz to 0.2 Hz. The normalized cyclic
load amplitude ξL remains 0.16 (5MN).
The effect of frequency of cyclic loading on transient pore pressure
and pile-top displacement is shown in Figs. 29 and 30. It can be seen
from Fig. 30 that the displacement of monopile increases with increasing
the loading frequency. Fig. 31 shows that the subsidence depth increases
with increasing the loading frequency, and the law supported the results
of Fig. 30. This can be explained by this fact, the soil at the shallow
Fig. 24. Relation of influence parameter c and soil relative density Dr. depeths has the chance to drain partially some of the excess pore pres­
sure, so obtaining some compaction and subsidence (Cuéllar, 2011). But
the deformation of monopile and subsidence depth has minor de­
viations. The maximum difference due to the frequency of cyclic loading

Fig. 25. Comparison of the subsidence depth soil of soil around pile.

soil relative density and the diameter is fixed in 55% and 7.5 m. The
normalized cyclic load amplitude ξL is 0.16 (5MN). The limit drainage
case is calculated to indicate the effect of transient pore pressure on the Fig. 26. History curves of transient pore pressure at 1 m depth for the different
cumulative deformation of monopile. At the limit drainage case, the soil soil permeability.

12
Y. Sun et al. Applied Ocean Research 133 (2023) 103481

Fig. 30. Curves of normalized pile-top lateral displacement for different


Fig. 27. Curves of normalized pile-top lateral displacement for different soil loading frequency.
permeability.

Fig. 31. Soil settlement in the vicinity of the pile under different
Fig. 28. Soil settlement in the vicinity of the pile under different soil loading frequency.
permeability.

is less than 5%. Under the different frequency of cyclic loading, the
transient pore pressure and the normalized pile-top displacement is
basically unchanged. The loading frequency has no effect on the varia­
tion of transient excess pore pressure and the cumulative deformation,
when the loading frequency ranges from 0.01 Hz to 0.20 Hz. So, in this
study the effect of loading frequency on the cumulative deformation is
neglectful. The same phenomena were found by Cuéllar (2011) and Li
et al. (2019) and Zhu et al. (2021). In this case, since this parametric
study only one parameter at a time has been investigated, the mutual
inter-dependence between the different factors such as permeability and
frequency of cyclic loading are not being captured. So for other cyclic
loading such as storm and other soil site, this effect of transient excess
pore pressure remains unclear and should be addressed in forthcoming
investigations.

4.5. Cyclic loading levels

The effect of one-way lateral cyclic loading is more than the two-way
cyclic loading for piles installed in sand (Richards et al., 2021; Cui et al.,
2016). Meanwhile to clearly present the effect of relevant parameters on
Fig. 29. History curves of transient pore pressure at 1 m depth for the different the cumulative deformation, Tc is taken as 1.0 (the cyclic lading is
loading frequency. one-way). Therefore, the pile cumulative lateral displacement is calcu­
lated considering one-way cyclic loading. In addition, the pile

13
Y. Sun et al. Applied Ocean Research 133 (2023) 103481

cumulative deformation under long-term cyclic loading belongs to the


fatigue loading condition; therefore, the normalized cyclic load ampli­
tude ξL is taken less than 0.3 (DNV, 2013; Zhu et al., 2013).
In this section, the pile diameter is 7.5 m and the relative density is
55%. The normalized parameter ξL is varied from 0.05 to 0.30. Fig. 31
presents the calculated variation of normalized lateral displacement at
the pile head with the number of load cycles for different cyclic load
levels. It is shown from Fig. 32 that the cumulative deformation of
monopile increases as the cyclic loading level increases. The results
presented in Fig. 32 are then curve fitted into Eq. (2) to back-calculate
the correlation parameters. The curve fitting showed that the param­
eter c is almost the same for different cyclic loading levels. On the other
hand, the parameter Tb increases linearly with the increase in normal­
ized cyclic load amplitude, ξL, as shown in Fig. 33. This is attributed to
the increase of pile head displacement as the amplitude of cyclic load
increases. Correspondingly, the parameter Tb can be correlated to the
cyclic loading parameter ξL.
Tb = 0.4ξL (6)
Fig. 33. Relation of influence parameter Tb and normalized loading param­
5. Proposed model of calculating pile-top cumulative eter εL.
displacement

Based on the above parametric study of cumulative pile deformation, yN = y0 [1 + ξL (0.02


D
+ 0.24)N ( − 0.0015 × Dr +0.6)] (9)
an analytical model is proposed herein, which accounts for the effect of Dref
monopile diameter, soil relative density and cyclic loading level. The
proposed model for calculating the cumulative deformation of the pile 5.2. Validation of proposed analytical model
head is then validated by analyzing the centrifugal load tests of model
offshore monopile. The proposed analytical model is validated by comparing its pre­
dictions with the measured responses of model offshore monopiles in the
centrifuge tests conducted at 50 g by Wang et al. 0 and model pile tests
5.1. Modified calculation model for accumulated deformation of
performed by Li et al. (2021).
monopile
5.2.1. Centrifuge tests
The parameters Tb and c in Eq. (2) are determined for offshore
The prototype cyclic loading frequency was 0.02 Hz and its ampli­
monopile from the parametric study described above. The parameter Tb
tude was 425 kN. The transient pore pressure is basically unchanged by
is determined as a function of the pile diameter, D, and cyclic loading
the variation of loading frequency when the frequency of cyclic loading
level ξL from Eqs. (4) and (6), and may be given by:
is low frequency (0.01 Hz-0.2 Hz). Therefore, this test can be used to
Tb = f (ξL , D) (7) evaluate this analytical method.
The numerical model developed in this paper is employed to calcu­
Tb = ξL (0.02
D
+ 0.24) (8) late the load at the pile head when the lateral pile displacement of 0.25
Dref m (0.1×2.5 m) is induced at the mudline. This is the ultimate lateral
force and is found to be 1540 kN based on the calculated results. This
Combining Eqs. (5) and (8), the analytical model for the cumulative
lateral load is considered the pile ultimate lateral capacity. The proposed
displacement of the pile head is obtained as:
analytical model is used to calculate the cumulative lateral displacement
at the pile head in the centrifuge model test. The normalized cyclic
loading amplitude ξL is 0.276 (i.e., cyclic load amplitude of 425 kN
divided by ultimate load of 1540 kN). The parameter c is calculated to be
0.51 from Eq. (3) using Dr = 60%. Substituting Dr = 60%, ξL = 0.276 and
c = 0.51 into Eq. (9), the cumulative displacement of the pile head can
be calculated for any number of load cycles. The calculated cumulative
displacements at the pile head obtained from the developed analytical
model and the LeBlanc analytical model (Leblanc et al., 2010) are
compared in Fig. 34. As can be observed from Fig. 34, the LeBlanc model
significantly underestimates the cumulative displacement of monopiles
installed in saturated soil. This is owning to LeBlanc calculated method
is developed based on lateral loaded pile tests in dry sand, and the effect
of pore water pressure on soil deformation in saturated sand is not
included. Meanwhile, the calculated cumulative monopile displacement
employing the proposed analytical model are in good agreement with
the experimental measurements, which validates the developed model.

5.2.2. Model pile tests


Li et al. (2021) conducted model pile tests in a steel box 80 cm long,
Fig. 32. Normalized pile-top lateral displacement under different 50 cm wide and 50 cm deep. The size of T2 model pile was 550 mm long
loading levels. aluminum pipe with 40 mm diameter and 2 mm wall thickness. The

14
Y. Sun et al. Applied Ocean Research 133 (2023) 103481

prototype case when prepared at the same void ratio (Hellmigk, 2021).
The difference in the results may also be attributed to the small diameter
of the test pile diameter, which renders the diameter effect to be
negligible (Sun et al., 2021).

6. Conclusions

A numerical model was developed employing the FLAC3D finite


difference platform and incorporating the SANISAND constitutive model
and was used to investigate and compare the lateral response of
monopiles installed in no pore pressure and saturated cases. The effects
of soil relative density, pile diameter and amplitude of cyclic loading on
the accumulative displacement of the pile head were evaluated. Finally,
an analytical model was developed to calculate the cumulative
displacement at the head of offshore monopiles under lateral cyclic
loading considering the effects of pile diameter. The main conclusions
drawn from the results of this study are as follows.

(1). Under lateral cyclic loading, the larger subsidence and the less
Fig. 34. Comparison of calculated and measured pile-top lateral displacement. capacity of soil around the pile is caused by accumulation and
dissipation of transient excess pore pressure in saturated sand. Thus
model pile tests were conducted in the laboratory (i.e., 1 g condition). the monopile in saturated case would experience the larger cumu­
The soil was commercial silica sand with relative density, Dr 90%. The lative deformation than no pore pressure case.
sand physical and mechanical properties are reported by ref (Li et al., (2). In this analytical model, the influence parameter c in the power
2021). The sand sample was saturated and the water level was 1–2 cm function for evaluating cumulative pile deformation under cyclic
above soil surface. The applied lateral one-way cyclic load was a sinu­ loading is independent of the pile diameter and cyclic load level, but
soidal load with an amplitude of 15 N acting at 100 mm above the soil is inversely linear with the soil relative density, Dr.
surface with frequency of 0.5 Hz. The pile displacement was measured at (3). In this analytical model, the influence parameter Tb in the power
a point 200 mm above the soil surface (Li et al., 2021). function for evaluating cumulative pile deformation under cyclic
To evaluate the versatility of the proposed analytical solution, it is loading is independent of Dr, but is linearly correlated with both the
used to calculate the cumulative displacement of the model pile tests. pile diameter, D, and amplitude of cyclic loading.
The ultimate lateral capacity is 191 N (the lateral displacement of pile at (4). In this study, because the load level is the frequent load case such
mudline is 55 mm). Based on the sand properties, the parameter c is 0.46 as common wave load, the amplitude and frequency of cyclic loading
and ξL is 0.08. Fig. 35 compares the calculated and measured pile is very low. So, sand permeability and frequency of cyclic loading
displacements. have a little effect on the cumulative displacement. For the large
It can be seen from Fig. 35 that at N = 5000, the difference between cyclic loading such as storm and other load levels, it remains unclear
the measured and calculated pile displacements is 15%. As the number and will be addressed in the forthcoming investigations.
of load cycles increases, the difference between the measured and (5). An analytical model is proposed to calculate the cumulative
calculated responses decreases to 4% at N = 20,000. Most importantly, displacement of offshore monopiles under cyclic loading considering
the evolution trend of the cumulative displacement is similar between the monopile diameter effect. The proposed model is validated by
the calculated and measured results. The observed differences are comparing its predictions with the results of centrifuge tests in
attributed to the variation of the soil constitutive behavior at the saturated sand and a 1 g model pile tests in saturated sand. The
different stress levels. This is because the model will exhibit a more proposed model can be used in preliminarily evaluation of the cu­
dilatant response and a lower excess pore pressure compared to the mulative lateral deformation of offshore wind turbine monopiles.

This developed analytical model is based on limited numerical


studies and validated with limited experimental studies. Therefore, the
applicability of the developed analytical model should be further
investigated by lateral cyclic load tests on full scale piles.

Author contributions statement

Sun Yilong: Conceptualization, Methodology, Writing-Reviewing,


Data curation, Writing-Original draft preparation, Software.
Xu Chengshun (corresponding author): Conceptualization, Method­
ology, Investigation, Supervision, Visualization.
M. Hesham El Naggar: Methodology, Validation, Review and Editing.
Du Xiuli: Investigation, Supervision, Visualization and Editing.
Dou Pengfei: Review and Editing.

CRediT authorship contribution statement

Yilong Sun: Conceptualization, Methodology, Writing – review &


editing, Data curation, Writing – original draft, Software. Chengshun
Xu: Conceptualization, Methodology, Investigation, Supervision, Visu­
Fig. 35. Comparison of calculated and measured pile-top lateral displacement. alization. M. Hesham El Naggar: Methodology, Validation, Writing –

15
Y. Sun et al. Applied Ocean Research 133 (2023) 103481

review & editing. Xiuli Du: Investigation, Supervision, Visualization, Chong, S.H., 2017. Numerical simulation of offshore foundations subjected to repetitive
loads. Ocean Eng. 142, 470–477. https://doi.org/10.1016/j.oceaneng.2017.07.031.
Writing – review & editing. Pengfei Dou: Writing – review & editing.
Cuéllar, P., Mira, P., Pastor, M., et al., 2014. A numerical model for the transient analysis
of offshore foundations under cyclic loading. Comput. Geotech. 59, 75–86. https://
doi.org/10.1016/j.compgeo.2014.02.005.
Declaration of Competing Interest Cuéllar, P., 2011. Pile Foundations For Offshore Wind Turbines: Numerical and
Experimental Investigations on the Behaviour Under Short-Term and Long-Term
Cyclic Loading. Von der Fakultt VI-Planen Bauen Umwelt der Technischen,
The authors declare that they have no known competing financial Universitt Berlin.
interests or personal relationships that could have appeared to influence Cui L., S. Bhattacharya. 2016. Soil–monopile interactions for offshore wind turbines. In:
Proceedings of the Institution of Civil Engineers—Engineering and Computational
the work reported in this paper. Mechanics. 0.1680/jencm.16.00006.
Dafalias, Y.F., Papadimitriou, A.G., Li, X.S., 2004. Sand plasticity model accounting for
Data availability inherent fabric anisotropy. J. Eng. Mech. 130 (1), 1319–1333. https://doi.org/
10.1061/(ASCE)0733-9399(2004)130:11(1319).
Damgaard, M., Bayat, M., Andersen, L.V., et al., 2014. Assessment of the dynamic
No data was used for the research described in the article. behaviour of saturated soil subjected to cyclic loading from offshore monopile wind
turbine foundations. Comput. Geotech. 61, 116–126. https://doi.org/10.1016/j.
compgeo.2014.05.008.
Det Norske Veritas. 2013. Offshore standard: Design of Offshore Wind Turbine
Acknowledgments Structures. DNV-OS-J101, Hovek, Norway.
Wang, Fuqiang, Rong, bing, Zhang, Ga, Zhang, jianmin, 2011. Centrifugal model test of
pile foundation for wind power unit under cyclic lateral loading. Yantu Lixue 32
This present study was supported by the National Outstanding Youth (07), 1926–1930. https://doi.org/10.16285/j.rsm.2011.07.042 (In Chinese).
Science Fund Project of National Natural Science Foundation of China Georgiadis, M., Anagnostopoulos, C., Saflekou, S., 1992. Centrifugal testing of laterally
(Grant No. 51722801). loaded piles in sand. Can. Geotech. J. 29 (2), 208–216. https://doi.org/10.1139/t92-
024.
Gerolymos, N., Escoffier, S., Gazetas, G., et al., 2009. Numerical modeling of centrifuge
References cyclic lateral pile load experiments. Earthquake Engineering and Engineering
Vibration 8 (1), 61–76. https://doi.org/10.1007/s11803-009-9005-8.
Giannakos, S., Gerolymos, N., 2012. Cyclic lateral response of piles in dry sand: finite
Achmus, M., Kuo, Y.S., Abdel-Rahman, K., 2009. Behavior of monopile foundations
element modeling and validation. Computers and Geotechnics. Gazetas G. 44,
under cyclic lateral load. Comput. Geotech. 36 (5), 725–735. https://doi.org/
116–131. https://doi.org/10.1016/j.compgeo.2012.03.013.
10.1016/j.compgeo.2008.12.003.
Itasca, 2012. Fast Lagrangian analysis of Continua in 3-dimensions, Version 5.0, Manual.
Ahmed, S.S., Hawlader, B., 2016. Numerical analysis of large-diameter monopiles in
Itasca, Minneapolis, MN.
dense sand supporting offshore wind turbines. Int. J. Geomech. 16 (5), 04016018.
Klinkvort, R.T., Hededal, O., 2013. Lateral response of monopile supporting an offshore
Albiker, J., Achmus, M., Frick, D., Flindt, F., 2017. 1G model tests on the displacement
wind turbine. Proc. Geotech. Eng. 166 (2), 147–158. https://doi.org/10.1680/
accumulation of large-diameter piles under cyclic lateral loading. Geotech. Test. J.
geng.12.00033.
40 (2), 20160102 https://doi.org/10.1520/GTJ20160102.
Klinkvort, R.T., Leth, C.T., Hededal, O., 2012. Centrifuge modelling of monopiles in
American Petroleum Institute, 2014. Recommended Practice for planning, Designing and
dense sand at The Technical University of Denmark. In: Proceedings of the 2nd
Constructing Fixed Offshore Platforms. American Petroleum Institute 2A-WSD, U.S.
European conference on Physical Modelling in Geotechnics. Eurofuge.
A.
Kuo, Y.S., Achmus, M., Abdel-Rahman, K., 2011. Minimum embedded length of cyclic
Arany, L., Bhattacharya, S., Macdonald, J., et al., 2017. Design of monopiles for offshore
horizontally loaded monopiles. J. Geotech. Geoenviron. Eng. 138 (3), 357–363.
wind turbines in 10 steps. Soil Dyn. Earthquake Eng. 92, 126–152. https://doi.org/
https://doi.org/10.1061/(ASCE)GT.1943-5606.0000602.
10.1016/j.soildyn.2016.09.024.
Lai Y., Wang L., Hong Y., et al. 2020. Centrifuge modeling of the cyclic lateral behavior of
Ashour, M., 2014. Contribution of vertical skin friction to the lateral resistance of large-
large-diameter monopiles in soft clay: effects of episodic cycling and reconsolidation.
diameter shafts. J. Bridge Eng. 19 (2), 289–302. https://doi.org/10.1061/(ASCE)
Ocean Eng.. 200, 107048. https://doi.org/10.1016/j.oceaneng.2020.107048.
BE.1943-5592.0000505.
Leblanc, C.B., Byrne, W., Houlsby, G.T., 2010. Response of stiff piles in sand to long-term
Barari, A., Bagheri, M., Rouainia, M., et al., 2017. Deformation mechanisms for offshore
cyclic lateral loading. Géotechnique 60 (2), 79–90. https://doi.org/10.1680/
monopile foundations accounting for cyclic mobility effects. Soil Dyn. Earthquake
geot.7.00196.
Eng. 97, 439–453. https://doi.org/10.1016/j.soildyn.2017.03.008.
Li, W., Igoe, D., Gavin, K., 2015. Field tests to investigate the cyclic response of
Bayat, M., Andersen, L.V., 2016. Ibsen. p-y-ẏ curves for dynamic analysis of offshore
monopiles in sand. Proc. Inst. Civil Eng.-Geotech. Eng. 168 (5), 407–421. https://
wind turbine monopile foundations. Soil Dyn. Earthquake Eng. 90, 38–51. https://
doi.org/10.1680/geng.14.00104.
doi.org/10.1016/j.soildyn.2016.08.015.
Li, S., Zhang, Y., Jostad, H.P., 2019. Drainage conditions around Monopiles in Sand.
Bayat, M., Andersen, L.V., Ibsen, L.B., et al., 2017. Influence of pore water in the seabed
Appl. Ocean Res. 86, 111–116. https://doi.org/10.1016/j.apor.2019.01.024.
on dynamic response of offshore wind turbines on monopiles. Soil Dyn. Earthquake
Li J.L., Guan D.W., Chiew Y.M., Zhang J.S., Zhao J.L. 2020. Temporal evolution of soil
Eng. 100, 233–248. https://doi.org/10.1016/j.soildyn.2017.06.001.
deformations around monopile foundations subjected to cyclic lateral loading.
Bhattacharya, S., Nikitas, G., Arany, L., Nikitas, N., 2017. Soil-structure interactions (SSI)
Ocean Eng.. 217, 107893. https://doi.org/10.1016/j.oceaneng.2020.107893.
for offshore wind turbines. Eng. Technol. Ref. https://doi.org/10.1049/
Li, L.C., Liu, H., Wu, Wb, Wen, M.J., El Naggar, M.H., Yang, Y.Z., 2021a. Investigation on
etr.2016.0019.
the behavior of hybrid pile foundation and its surrounding soil during cyclic lateral
Bienen, B., Dührkop, J., Grabe, J., et al., 2011. Response of piles with wings to monotonic
loading. Ocean Eng. 240 (15), 110006 https://doi.org/10.1016/j.
and cyclic lateral loading in sand. J. Geotech. Geoenviron. Eng. 138 (3), 364–375.
oceaneng.2021.110006.
https://doi.org/10.1061/(ASCE)GT.1943-5606.0000592.
Li, L., Liu, H., Wu, W., Wen, M., El Naggar, M.H., Yang, Y., 2021b. Investigation on the
Bouzid, D.A., 2018. Numerical investigation of large-diameter monopiles in Sands:
behavior of hybrid pile foundation and its surrounding soil during cyclic lateral
critical review and evaluation of both API and newly proposed p-y curves. Int. J.
loading. Ocean Eng. 240, 110006 https://doi.org/10.1016/j.
Geomech. 18 (11), 4018141 https://doi.org/10.1061/(ASCE)GM.1943-
oceaneng.2021.110006.
5622.0001204.
Lin, S.S., Liao, J.C., 1999. Permanent strains of piles in sand due to cyclic lateral loads.
Burd, H.J., Abadie, C.N., Byrne, B.W., Houlsby, G.T., Martin, C.M., McAdam, R.A.,
J. Geotech. Geoenviron. Eng. 125 (9), 798–802. https://doi.org/10.1061/(ASCE)
Jardine, R.J., Pedro, A.M.G., Potts, D.M., Taborda, D.M.G., Zdravković, L.,
1090-0241(1999)125:9(798).
Andrade, M.P., 2020a. Application of the PISA design model to monopiles embedded
Liu, H., Kaynia, A.M., 2022. Monopile Responses to Monotonic and Cyclic Loading in
in layered soils. Géotechnique 70 (11), 1067–1082. https://doi.org/10.1680/
Undrained Sand Using 3D FE with SANISAND-MSu. Water Sci. Eng. 15 (1), 69–77.
jgeot.20.PISA.009.
https://doi.org/10.1016/j.wse.2021.12.001.
Burd, H.J., Taborda, D.M.G., Zdravković, L., Abadie, C.N., Byrne, B.W., Houlsby, G.T.,
Long, J.H., Vanneste, G., 1994. Effects of cyclic lateral loads on piles in sand. J. Geotech.
Gavin, K.G., Igoe, D.J.P., Jardine, R.J., Martin, C.M., McAdam, R.A., Pedro, A.M.G.,
Eng. 120 (1), 225–244.
Potts, D.M., 2020b. PISA design model for Monopiles for offshore wind turbines:
Luo, R., Yang, M., Li, W, 2018. Numerical study of diameter effect on accumulated
application to a marine sand. Géotechnique 70 (11), 1048–1066. https://doi.org/
deformation of laterally loaded monopiles in sand. Eur. J. Environ. Civil Eng. 1–13.
10.1680/jgeot.18.P.277.
https://doi.org/10.1080/19648189.2018.1506828.
Cheng, Z., Dafalias, Y.F., Manzari, M.T., 2013. Application of SANISAND Dafalias-
Ong, D.E.L., Leung, C.E., Chow, Y.K., 2006. Pile behavior due to excavation-induced soil
Manzari model in FLAC3D. Continuum and Distinct Element Numerical Modeling in
movement in clay. I: stable wall. J. Geotech. Geoenviron. Eng. 132, 36–44. https://
Geomechanics. Itasca International, Inc, Minneapolis. H. Zhu, C. Detournay, R. Hart,
doi.org/10.1061/(ASCE)1090-0241(2006)132:1(36).
and M. Nelson, Eds.
Ong, D.E.L., Sim, Y.S., Leung, C.F., 2018. Performance of field and numerical back-
Chong, S.H., Pasten, C., 2018. Numerical study on long-term monopile foundation
analysis of floating stone columns in soft clay considering the influence of dilatancy.
response. Mar. Georesour. Geotechnol. 36 (2), 190–196. https://doi.org/10.1080/
Int. J. Geomech. 18 (10), 4018131–4018135. https://doi.org/10.1061/(ASCE)
1064119X.2017.1293200.
GM.1943-5622.0001261.
Chong, S.H., Shin, H.S., Cho, G.C, 2019. Numerical analysis of offshore monopile during
Prasad, Y.V.S.N., Chari, R., 1999. lateral capacity of model rigid piles in cohesionless soil.
repetitive lateral loading. Geomech. Geoeng. 19 (1), 79–91. https://doi.org/
Soils Found. 39 (2), 21–29. https://doi.org/10.3208/sandf.39.2_21.
10.12989/gae.2019.19.1.079.

16
Y. Sun et al. Applied Ocean Research 133 (2023) 103481

Ramirez, J., Barrero, A.R., Chen, L., Dashti, S., Ghofrani, A., Taiebat, M., Arduino, P., Sun Y., Xu C., Du X., et al. 2020. Nonlinear lateral response of offshore large-diameter
2018. Site response in a layered liquefable deposit: evaluation of different numerical monopile in sand. Ocean Eng.. 216, 108013. https://doi.org/10.1016/j.
tools and methodologies with centrifuge experimental results. J. Geotech. oceaneng.2020.108013.
Geoenviron. Eng. 144 (10), 04018073 https://doi.org/10.1061/(ASCE)GT.1943- Thieken, K., Achmus, M., Lemke, K., 2015. A new static p-y approach for piles with
5606.0001947. arbitrary dimensions in sand. Geotechnik 38 (4), 267–288. https://doi.org/10.1002/
Reese, L.C., Cox, W.R., Koop, F.D., 1974. Analysis of laterally loaded piles in sand. gete.201400036.
Offshore Technol. Conf., Richardson, TX. Verdugo, R., Ishihara, K., 1996. The steady state of sand soils. Soils Found. 36 (2), 81–92.
Richards, I.A., Bransby, M.F., Byrne, B.W., Gaudin, C., Houlsby, G.T., 2021. Effect of Yang, M., Taiebat, M., Dafalias, Y.F., 2020. Sanisand-msf: a memory surface and
stress level on response of model monopile to cyclic lateral loading in sand. semifluidized state enhanced sand plasticity model for undrained cyclic shearing.
J. Geotech. Geoenviron. Eng. 147 (3), 4021002 https://doi.org/10.1061/(ASCE) Géotechnique. https://doi.org/10.1680/jgeot.19.P.363.
GT.1943-5606.0002447. Yang, Min, et al., 2017. Numerical study on accumulated deformation of laterally loaded
Rosquoët, F., Garnier, J., Thorel, L., Canepa, Y., 2004. Horizontal cyclic loading of piles monopiles used by offshore wind turbine. Bull. Eng. Geol. Environ. 77 (3), 911–921.
installed in sand: study of the pile head displacement and maximum bending https://doi.org/10.1007/s10064-017-1138-9.
moment. editor. In: Triantafyllidis, T (Ed.), Proceedings of the international Zhu, B., Byrne, B.W., Houlsby, G.T., 2013. Long-term lateral cyclic response of suction
conference on cyclic behaviour of soils and liquefaction phenomena. Taylor & caisson foundations in sand. J. Geotech. Geoenviron. Eng., ASCE 139 (1), 73–83.
Francis, Bochum. https://doi.org/10.1061/(ASCE)GT.1943-5606.0000738.
Sørensen S P H L, 2010. Effects of diameter on initial stiffness of p-y curve for large- Zhu, Z., Zhang, F., Peng, Q., et al., 2021. Effect of the loading frequency on the sand
diameter piles in sand: In Proc., 7th Conf. on Numerical Methods in Geotechnical liquefaction behaviour in cyclic triaxial tests. Soil Dyn. Earthquake Eng. 147, 106779
Engineering, 907–912. Norway: Trondheim. https://doi.org/10.1016/j.soildyn.2021.106779.
Zienkiewicz, O.C., Chang, C.T., Bettess, P., 1980. Drained, undrained, consolidating
dynamic behaviour assumptions in soils. Géotechnique 30, 385–395.

17

View publication stats

You might also like