A Unified Lateral Soil Reaction Model For Monopiles in Soft Clay Considering Various Length-To-Diameter Ratios

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 15

Ocean Engineering 212 (2020) 107492

Contents lists available at ScienceDirect

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

A unified lateral soil reaction model for monopiles in soft clay considering
various length-to-diameter (L/D) ratios
Lizhong Wang a, Yongqing Lai a, Yi Hong a, *, David Ma�sín b
a
Key Laboratory of Offshore Geotechnics and Material of Zhejiang Province, College of Civil Engineering and Architecture, Zhejiang University, China
b
Faculty of Science, Charles University in Prague, Czech Republic

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

Keywords: Large-diameter monopiles are the most commonly used foundation to support offshore wind turbines. Early
Monopile designs usually adopted pile diameters (D) between 4 and 6 m, which is recently extended to 8 m and will target
Soft clay 10 m in the future. It is increasingly evident that the existing design method (i.e., API’s p-y model) can signif­
p-y reaction model
icantly under-predict the lateral stiffness and capacity of large-diameter monopiles in soft clay, due to ignoring
Base reaction model
Length-to-diameter ratio
the soil resistances from base shear and base moment which become more pronounces as L/D reduces. In this
Finite element analysis study, a two-spring approach is proposed, aiming to predict the lateral behaviour of monopiles with varied L/D
ratios in a unified manner. In light of the soil flow mechanisms around monopiles, the pure lateral soil resistance
above the rotation point (RP) is quantified using a p-y model, while the resistances below the RP including the
base shear and base moment are integrated into a moment-rotation spring (characterized by a MR-θR model) at
the RP. It can naturally recover to a p-y model while analyzing flexible piles, where θR ¼ 0 at RP. Formulations of
the ‘p-y þ MR-θR’ model (including diameter-related p-y and MR-θR models, and the depth of the RP) are proposed
based on the results of a series of well-calibrated 3D numerical models. The proposed model has satisfactorily
reproduced a number of field and centrifuge test results on laterally loaded monopiles with a wide range of L/D
ratios (including flexible, semi-rigid and rigid piles), using a unified set of parameters. Compared to the standard
p-y model, the adoption of the proposed ‘p-y þ MR-θR’ model is shown to substantially reduce design
conservatism.

1. Introduction based on the Winkler approach (1867) in conjunction with p-y models
(API, 2014; DNVGL, 2016). The original p-y curves were deduced from
Monopiles are the most commonly adopted foundation for support­ the field tests on long slender piles with D ¼ 0.324 m and L/D ¼ 39
ing offshore wind turbines (OWTs) (Zhang et al., 2016; Wang et al., (Matlock, 1970), and have been widely used for the design of slender
2017, 2019), accounting for more than 80% of the readily installed piles supporting offshore oil and gas platforms for decades. In recent
OWTs (EWEA, 2016). Due to the extreme environmental loads and years, attempts have been made to apply these p-y models for the design
increasing turbine sizes, early designs usually adopted pile diameters (D) of monopiles supporting offshore wind turbines, which have larger D
between 4 and 6 m (Leblanc et al., 2010; Negro et al., 2017; Stone et al., and much smaller L/D ratios than the slender piles for offshore oil and
2018), with recent designs extending to 8 m (Byrne et al., 2017; Zhang gas structures. It becomes increasingly evident that the original p-y
and Andersen, 2019; Achmus et al., 2019) and future designs exceeding curves are insufficient to capture large-diameter monopiles with varied
10 m (Shadlou and Bhattacharya, 2016; Byrne et al., 2017). The L/D ratios, due to the significant deviation of the geometries of the
embedded pile length to diameter ratio (L/D) of a monopile is typically monopiles from those adopted in the early calibration field tests (Ste­
in the range of 4–8 (Doherty and Gavin, 2012; Qi et al., 2016; Murphy vens and Audibert, 1979; Lam and Martin, 1986; Jeanjean, 2009;
et al., 2018), with future designs anticipating L/D ratios of 3 or smaller Madabhushi and Haiderali, 2013; Achmus and Thieken, 2016; Finn and
for super-large monopiles (Murphy et al., 2018). Dowling, 2016; Wang et al., 2015; Lau, 2015; He, 2016; Hong et al.,
The design for a laterally loaded monopile, which is mainly con­ 2017a; Byrne et al., 2017; Zhang and Andersen, 2019). For this reason,
cerned with the initial stiffness and the ultimate capacity, is typically the recent edition of the DNVGL (2016) guideline has been updated to

* Corresponding author.
E-mail addresses: wanglz@zju.edu.cn (L. Wang), yongqing_lai@zju.edu.cn (Y. Lai), yi_hong@zju.edu.cn (Y. Hong), david.masin@natur.cuni.cz (D. Ma�sín).

https://doi.org/10.1016/j.oceaneng.2020.107492
Received 22 December 2019; Received in revised form 25 April 2020; Accepted 4 May 2020
Available online 7 July 2020
0029-8018/© 2020 Elsevier Ltd. All rights reserved.
L. Wang et al. Ocean Engineering 212 (2020) 107492

Fig. 1. Soil flow mechanism and lateral displacement profile of (a) flexible pile (L/D ¼ 30), (b) semi-rigid pile (L/D ¼ 8) and (c) rigid pile (L/D ¼ 4) (Lai et al., 2019).

2
L. Wang et al. Ocean Engineering 212 (2020) 107492

2. Soil flow mechanisms around monopiles and the existing


mechanism-based soil-pile interaction models: a review

The previous centrifuge and numerical analyses have revealed that


the L/D ratio of a monopile (controlling the relative soil-pile stiffness)
plays an important role in determining its nearby soil flow mechanisms,
and thus the modes of soil-pile interaction (Hong et al., 2017a; Lai et al.,
2019). Fig. 1(a) and (b) and 1(c) show the soil flow mechanism and
lateral displacement profile of a typical flexible pile (L/D ¼ 30), a
semi-rigid pile (L/D ¼ 8), and a rigid pile (L/D ¼ 4), respectively (Lai
et al., 2019), which were computed by well-calibrated 3D numerical
models. While Fig. 2 illustrates the measured soil flow mechanisms
around a semi-rigid pile in soft clay, as quantified through the particle
image velocimetry technique (PIV) in half model centrifuge tests (Hong
et al., 2017a).
For the flexible pile (L/D ¼ 30, see Fig. 1(a)), the soil flow mecha­
nisms are composed of wedge-type flow near the ground surface and
full-flow below the wedge, with zero pile deflection and rotation at the
pile base. Under this circumstance, the lateral load imposed on the pile
Fig. 2. Measured soil flow mechanism around a semi-rigid pile in soft clay head is mostly resisted by mobilizing the lateral soil resistance in the
(Hong et al., 2017a). wedge and full-flow zones. As the L/D ratio of the pile reduces, the piles
behave as semi-rigid or rigid piles (Fig. 1(a) and (b) and Fig. 2), which do
suggest that any method used for monopile design should be validated not only bend but also rotates with a distinct pile toe-kick. An additional
by other means, such as finite element analyses. It is also worth noting flow mechanism, namely rotational flow near the lower half of the piles
that the track records of success using the existing design guidelines (see Fig. 1(a) and (b) and Fig. 2), is thus involved in the semi-rigid or
(API, 2014; DNVGL, 2016) are mainly proven in the design of monopiles rigid piles. In other words, the lateral load imposed to the semi-rigid or
in European wind farms, where overconsolidated stiff clay or dense sand rigid piles is not only resisted by the lateral soil resistance near its upper
are usually encountered (Lau, 2015; Wang et al., 2018). These guide­ half, but also by a rotational moment near its lower half with contri­
lines do not appear to be very suitable for designing monopiles in nor­ butions from base shear, base moment and rotational soil flow. This
mally consolidated clay (Zhu et al., 2017), as widely deposited in explains the inadequacy of the API (2014)’s approach to capture the
offshore China (Cai et al., 2010; Di et al., 2013; Nan et al., 2018). lateral behaviour of monopiles with small L/D ratios, due to the igno­
Given the research gaps identified above, this study aims to develop a rance of the latter mechanism (Gerolymos and Gazetas, 2006; Lam,
unified lateral soil reaction model for monopiles in soft clay, with 2009; Gao et al., 2015; Byrne et al., 2017, 2019a; 2019b; McAdam et al.,
consideration of various length-to-diameter (L/D) ratios. In the 2019; Taborda et al., 2019; Zdravkovi�c et al., 2019).
following, the soil flow mechanisms around monopiles with varied L/D To better predict the lateral behaviour of monopiles with relatively
ratios in conjunction with some existing mechanism-based conceptual/ small L/D ratios (i.e., semi-rigid or rigid piles), a number of improved
explicit models are reviewed, with their merits and potential demerits soil-pile interaction models have been proposed over the last decades.
highlighted. The lessons learned from the review has led to the devel­ Most of these models were proposed based on semi-empirical ap­
opment of a new hybrid ‘p-y þ MR-θR’ approach, that is capable of proaches by adding some lump-sum factors into the p-y curves (Stevens
predicting lateral behaviour of monopiles with varied L/D ratios using a and Audibert, 1979; O’Neill and Gazioglu, 1984; Lau, 2015; Hong et al.,
unified set of parameters. The validity of the unified model is justified 2017a), without explicitly consider the contributions from the base
against a number of field and centrifuge model tests in soft clay shear, base moment and vertical shaft shear stress. These semi-empirical
embedded with monopiles, which cover a broad range of L/D ratios. approaches have worked well case-specifically, but would probably not
be able to predict lateral behaviour of piles with a wide range of L/D in a
unified manner.
In view of the limitation stated above, mechanism-based soil-pile

Fig. 3. Existing models for predicting lateral response of monopiles: (a) one-spring p-y model; (b) four-spring model (Byrne et al., 2017, 2019a); (c) two-spring model
(Zhang and Andersen, 2019).

3
L. Wang et al. Ocean Engineering 212 (2020) 107492

are intended to be linked to the elemental stress-strain response of the


soil (Zhang et al., 2016; Zhang and Andersen, 2017). The resistance due
to the base moment, which could also be an important resistance
component (as quantified in sub-section 4.4), is ignored in the
two-spring model. Due to this omission, the two-spring model is likely to
under-predict the lateral pile stiffness and capacity.

3. A new conceptual framework for unified modelling of


monopiles with varied L/D

Based on the merits and potential demerits of the existing models,


this study proposes a new two-spring model, i.e., ‘p-y þ MR-θR’ model, as
shown in Fig. 4. It considers the three most important components
resisting a laterally loaded monopile, namely (i) lateral soil resistance;
(ii) base shear and (iii) base moment. Above the rotation point, the
lateral soil resistance is modelled by distributed lateral translational
springs (p-y springs), as routinely practiced. While the overall re­
sistances at and below the rotation point, including the contributions
from the rotational soil flow (as evident numerically and experimen­
tally, see Figs. 1 and 2), base shear and base moment, are integrated into
a concentrated rotational spring (MR-θR spring) at the rotation point.
The rotation point, which involves a pure rotation with little lateral
displacement (Randolph and Gourvenec, 2011; Zhang and Ahmari,
2011; Lau, 2015), is usually located within a narrow range of depth z ¼
0.75–0.80L (Achmus et al., 2009; Zhang et al., 2013; Ahmed and
Hawlader, 2016; He, 2016; Truong and Lehane, 2017; Murphy et al.,
2018). The well-defined boundary conditions and location of the

Fig. 4. A new model proposed in this study for monopiles with varied L/D.

interaction models were recently reported by Byrne et al. (2017, 2019a)


and Zhang and Andersen (2019), which explicitly consider the re­
sistances from the pile base. Compared to the traditional one kind spring
p-y model (see Fig. 3(a)), Byrne et al. (2017, 2019a)’s four-spring con­
ceptual framework (as elaborated in Fig. 3(b)) comprehensively con­
siders all the four components resisting a laterally loaded monopile,
namely (i) lateral soil resistance (p-y springs); (ii) base shear (sb-yb
spring); (iii) base moment (Mb-θb spring) and (iv) vertical shaft shear
stress (Mp-θp springs). However, the four-spring model should be cali­
brated using a set of bespoke 3D finite element analyses of monopile
performance, for pile characteristics and loading conditions that span a
predefined design space (Byrne et al., 2019a). On the other hand, Zhang
and Andersen (2019)’s two-spring model (as shown in Fig. 3(c)) ac­
counts for two components resisting a laterally loaded monopile,
Fig. 5. Typical three-dimensional finite element mesh for a pile with D ¼ 6 m
including (i) lateral soil resistance (p-y springs) and (ii) base shear (sb-yb
and L/D ¼ 5.
spring). For the ease of practical design, the equations of these springs

Table 1
Program for the numerical parametric study.
Pile diameter D (m) Pile wall thickness t (m) Embedded depth L (m) L/D ratio Load eccentricity h Relative pile-soil stiffness Pile rigidity
Ep I p
Es L4
1 0.016 30 30 2D 0.0002 flexible
2 0.026 15 0.0022 flexible
3 0.036 10 0.0104 semi- rigid
4 0.046 7.5 0.0316 semi- rigid
5 0.056 6 2D, 4D 0.0750 semi- rigid
6 0.066 5 2D 0.1528 semi- rigid
7 0.076 4.3 0.2792 rigid
8 0.086 3.8 2D, 7D 0.4716 rigid
9 0.096 3.3 2D 0.7494 rigid
10 0.106 3 1.1349 rigid

Ep Ip
Note: According to the criterion proposed by Poulos and Hull (1989), the upper bound and lower bound of for flexible and rigid piles are 0.0025 and 0.208,
Es L4
E p Ip
respectively. Piles whose lies somewhere between those of flexible and rigid piles are defined as semi-rigid piles (Hong et al., 2017a).
Es L4

4
L. Wang et al. Ocean Engineering 212 (2020) 107492

Table 2 4. Quantifying effect of L/D ratio on lateral response of


Summary of model parameters adopted in the numerical analyses. monopiles in soft clay: 3D numerical parametric study
Parameter Value Remark
4.1. Objective and program
Monotonic Slope of the λ* 0.095 Calibrated against
response at isotropic NCL in the Hong et al. (2017b)’s
medium to ln(1 þ e) - lnp’ isotropic This section aims to quantify the lateral behaviour of piles with
large strain space consolidation data varied length-to-diameter ratios (i.e., L/D) in soft clay, through a series
levels Slope of the κ* 0.024 of numerical parametric study using a well-calibrated hypoplastic clay
isotropic unloading
line in the ln(1 þ e)
model. The computed results are to form the bases for formulating the
- lnp’ space proposed ‘p-y þ MR-θR’ model, which is mainly concerned with the p-y
Position of the N 1.44 relation above the rotation point, MR-θR relation at the rotation point
isotropic NCL in the and the depth of the rotation point. A wide range of L/D (between 3 and
ln(1 þ e) - lnp’
30) is considered in the numerical parametric study, where L is
space
Critical state φc 25.4� Calibrated against constantly taken as 30 m with the pile diameter D varying between 1 and
friction angle Hong et al. (2017b, 10 m. The pile wall thickness (t) was determined based on the following
Parameter ν 0.1 2019)’s triaxial test equation recommended by the API (2014) code:
controlling the data
proportion of bulk D
t ¼ 0:00635 þ (1)
and shear stiffness 100
Small-strain Strain range of soil R 10–4 Ma�sín (2005)
stiffness upon elasticity Table 1 summarizes the program for the numerical parametric study.
various strain Path-dependent mrat 0.7 Three different values of load eccentricity above the mudline (h ¼ 2D,
reversal parameter 4D and 7D) are considered. The two-spring model and model parameters
Strain-dependent 0.12 Hong et al. (2017a)
βr
were derived based on the analyses with h ¼ 2D, and used to predict the
parameter 1
Strain-dependent χ 5 lateral behavior of the piles with h ¼ 4D and 7D, aiming to check the
parameter 2
Stress-dependent Ag 650
parameter 1
Stress-dependent ng 0.65
parameter 2

rotation point bring convenience for constructing the ‘p-y þ MR-θR’


model.
For unified modelling on flexible, semi-rigid and rigid piles, the
proposed ‘p-y þ MR-θR’ model is anticipated to remain the flexibility of
adjusting the relative contribution of p-y and MR-θR depending on L/D
ratios. For example, the ‘p-y þ MR-θR’ model should naturally recover to
the conventional p-y model while analysing a flexible pile.
The detailed formulations for the p-y and the MR-θR springs
(including the depth of the rotation point) are proposed based on a series
of well-calibrated finite element analyses, which simulate piles with a
broad range of L/D in soft clay. The numerical analysis and the formu­
lation of the ‘p-y þ MR-θR’ model are presented in the following two
sections.

Fig. 7. Contribution of the pile base shear, base moment and vertical shaft
shear stress to the ultimate lateral pile capacity of the typical piles.

Fig. 6. Comparison between computed and predicted pile-head load-displacement response of typical piles: (a) D ¼ 2 m and L/D ¼ 15 (flexible pile), (b) D ¼ 4 m and
L/D ¼ 7.5 (semi-rigid pile) and (c) D ¼ 10 m and L/D ¼ 3 (rigid pile).

5
L. Wang et al. Ocean Engineering 212 (2020) 107492

Fig. 8. Comparison between (a) p-y curves extracted from FEA and (b) p-y curves fitted by the hyperbolic tangent function.

Fig. 9. Relationship between fitting parameter and pile diameter: (a) fitting parameter a; (b) fitting parameter b.

applicability of the proposed ‘p-y þ MR-θR’ model for piles with varying with the two mesh sizes differs by no more than 3%, confirming the
load eccentricity. validity of the mesh size adopted in the present study.
The parameters of Malaysia kaolin, which has been well- The clay and the pile are modelled using Eight-node brick with pore
characterized and is widely used in geotechnical model tests (Ilyas pressure (C3D8P) elements and Eight-node brick (C3D8) elements,
et al., 2004; Xie et al., 2012; Hong et al., 2017b, 2019; Zhu et al., 2018) respectively. The interface shear behaviour between the pile and the soil
was adopted in each analysis. Table 1 also includes the relative pile-soil is modelled based on the Coulomb friction law, where the tangential
stiffness, EpIp/EsL4 (Poulos and Hull, 1989). Es denotes the soil modulus frictional stress is linearly proportional to the normal stress. The inter­
at middle pile depth, which is equal to 400 times of su for Malaysia face frictional coefficient μ ¼ 0.31 was adopted in this study following
kaolin (Lai et al., 2019, 2020). As can be seen in the table, the piles with Randolph and Wroth (1981).
D between 1 and 2, 3–6 and 7–10 m tend to behave as flexible, semi-rigid
and rigid piles, respectively. 4.3. Constitutive model and model parameters

An advanced hypoplastic clay model with consideration of small-


4.2. Finite element mesh and boundary conditions strain stiffness (Ma�sín, 2005), which has been calibrated against
centrifuge tests on a lateral pile in kaolin (Hong et al., 2017a; He et al.,
Each numerical analysis is performed in the finite element program 2019), is adopted for this numerical parametric study. The hypoplastic
ABAQUS. Fig. 5 shows an isometric view of a typical finite element model formulates the nonlinear stress-strain behaviour of soil based on
mesh, which contains a pile with D ¼ 6 m and L/D ¼ 5. By taking the combination of a linear and a non-linear term, as follows (Wu and
advantage of symmetry, only half of the pile model was modelled for Kolymbas, 1990; Wu et al., 1996; Gudehus, 1996):
computational efficiency. To minimize the boundary effect, the diam­
eter of the cylindrical computation domain and the distance from the �
T ¼ fs ðL : D þ fa NkDkÞ (2)
pile base to the bottom boundary were chosen to be 20D and 5D,
respectively (Chen and Poulos, 1993). Each lateral boundary and the where � T and D represent the objective stress rate and the Euler
bottom boundary of the finite element mesh is constrained by roller and stretching tensor, respectively; L and N are fourth- and second-order
pinned supports, respectively. The suitability of the mesh density is constitutive tensors; and fs and fd are two scale factors. More details of
justified by halving the current size of the mesh and running one more each term in Eq. (2) are given in Ma�sín (2005, 2014). The strength
analysis. The computed load-displacement response from the analyses anisotropy is considered in the model, which follows Matsuoka and

6
L. Wang et al. Ocean Engineering 212 (2020) 107492

Fig. 11. Change of the depth of the rotation point with pile-head displacement.

In addition to the five basic model parameters, there are six pa­
rameters controlling the non-linearity of small strain stiffness, i.e., R,
mrat, βr, χ , Ag, and ng. The parameter R denotes the size of the elastic
range. The parameter mrat controls the very small strain behaviour upon
strain path reversals. The parameters βr and χ control the rate of stiffness
degradation. The remaining parameters Ag, and ng influence the initial
shear modulus through the equation (Wroth and Houlsby, 1985):
� ’ �ng
p
G 0 ¼ p r Ag (3a)
pr

where pr is a reference pressure of 1 kPa.


Fig. 10. Comparison between computed and calculated distribution of lateral These parameters for kaolin clay have been calibrated against
bearing capacity factor. triaxial tests (Hong et al., 2017b, 2019) and centrifuge model tests on
laterally loaded piles in kaolin (Hong et al., 2017a; He et al., 2019), as
summarized in Table 2. Normally consolidated clay is modelled in each
Table 3
analysis of this study, where the undrained shear strength (being a
Summary of equations for Np distribution.
function of void ratio) increases with depth.
Reference Equation

API (2014) z
Np ¼ 3 þ J �9 4.4. Computed lateral load-displacement response of piles with varying L/
D
where J ¼ 0.5 for linearly increasing shear strength profiles D
Stevens and Audibert z
Np ¼ 5 þ 2:5 � 12
(1979) D
Fig. 6(a) and (b) and 6(c) show the computed (by 3D FEA) load-
Jeanjean (2009)
� z�
Np ¼ 12 4 exp ξ displacement response at the head of a flexible (D ¼ 2 m, L/D ¼ 15),
D
ξ ¼ 0:25 þ 0:05λ � 0:55 semi-rigid (D ¼ 4 m, L/D ¼ 7.5) and rigid pile (D ¼ 10 m, L/D ¼ 3),
λ ¼ su0 =ðsu1 DÞ
respectively. Each figure also includes the calculated load-displacement
where su0 and su1 denote shear strength intercept at mudline
and rate of increase of shear strength with depth,
response at the pile head via beam on elastic foundation analyses using
respectively. API (2014)’s p-y curves and using the p-y relations extracted from the
Truong and Lehane
h � z �i three piles in 3D FEA analysis. As shown in each figure, the lateral load
Np ¼ 10:5 1 0:75 exp 0:6
(2017) D of each pile increases with head displacement but at a decreasing rate,
with the load-displacement curve eventually reaching a plateau. The
lateral load at the plateau is defined as the ultimate lateral pile capacity
Nakai (1974) failure criterion. The non-linearity of soil stiffness at small
(Fu) in this study (Lau, 2015; Truong and Lehane, 2017).
strains, which governs the initial stiffness of laterally loaded pile, is
It can be seen from Fig. 6(a) that for the flexible pile (L/D ¼ 15), the
considered in the hypoplastic model with the aid of the intergranular
adoption of API (2014)’s p-y curves underestimates the lateral stiffness
strain concept (Niemunis and Herle, 1997).
and the ultimate lateral pile capacity compared to the computed lateral
The model consists of eleven parameters. Five of these parameters, i.
pile load-displacement curves from 3D FEA. Similar observations were
e., ϕ0 c, N,λ*,κ* and ν, are identical or similar to those defined in the Cam-
also made by Jeanjean (2009), Zakeri et al. (2016) and Truong and
clay based models (Wang et al., 2016; Hong et al., 2020). Parameters N,
Lehane. (2017). With the increase of pile diameter (i.e., reducing L/D),
λ*, and κ* define the position of the isotropic virgin compression, slope
the underestimation of lateral pile stiffness and capacity by API (2014)
of the isotropic virgin compression, and the slope of the unloading line
becomes more pronounced, as shown in Fig. 6(b) and (c). In particular,
in the ln(1 þ e)-ln (p’) plane (e ¼ void ratio and p’ ¼ mean effective
the analysis based on API (2014) has underestimated the lateral pile
stress), respectively. The parameter ν controls the proportion of bulk and
capacity of the monopile with L/D ¼ 3 (Fig. 6(c)) is by 152%, implying
shear stiffness.
the current design practice may be over-conservative for rigid piles.

7
L. Wang et al. Ocean Engineering 212 (2020) 107492

Fig. 12. Relationship between the bending moment and the rotation angle at the rotation point.

appears to be marginal, and may be ignored. These observations suggest


that the two-spring model proposed by Zhang and Andersen (2019),
which considers the lateral soil resistance and base shear, may be further
improved by accounting for the contribution from the base moment.
In this study, the combined effects of the base shear and base moment
are integrated into one single spring of MR-θR, while the lateral soil
resistance is still quantified using the p-y curves. The proposed ‘p-y þ
MR-θR’ model is formulated in the following section, by generalizing the
computed lateral responses of piles with varying pile geometries.

5. Proposed ‘p-y þ MR-θR’ approach

5.1. Generalizing p-y curves

This section aims to generalize a unified p-y formulation that is able


to capture the lateral behaviour of piles with varied pile diameters using
a single set of parameters, based on the computed p-y curves via the 3D
FEA. For this purpose, the p-y curves of all the ten piles (i.e., D ¼ 1–10 m)
were extracted. Fig. 8(a) shows the typical computed p-y curves at the
shallow depths (which cover the region of wedge or full-flow types of
Fig. 13. Computed ultimate bending moment at the rotation point. soil flow mechanisms) of the typical piles with D ¼ 2, 4, 5, 6, 8 and 10 m,
as illustrated in different colours. The lateral soil resistance (p) was
normalised by the ultimate lateral soil resistance (pu), while the lateral
Based on the p-y curves extracted from the piles in the 3D FEA, the
pile displacement (y) was normalised by the pile diameter (D). The
beam-on-elastic-foundation analysis calculates almost the same load-
normalised API p-y curve is also plotted in the figure for comparison.
displacement curve for the flexible pile (L/D ¼ 15). This is anticipated
It can be seen that for each pile, the p-y curves at different depths of
because the lateral load applied to the head of a flexible pile is almost
pile embedded within the region of wedge or full-flow types of soil flow
resisted by the lateral resistance of soil (characterized by p-y curves). For
mechanisms are largely similar. While for piles with varied pile di­
the semi-rigid pile (L/D ¼ 7.5) and the rigid pile (L/D ¼ 3), however, the
ameters, the increase of the pile diameter has led to a higher initial
calculated load-displacement relationship using the p-y curves extracted
stiffness and smaller lateral displacement for reaching pu. These trends
from the 3D FEA underestimates the computed response. The degree of
are consistent with the conclusions drawn from the previous experi­
the underestimation increases with the pile diameter (i.e., reducing L/
mental and numerical investigations (Carter, 1984; Ling, 1988; Jeong
D). This is because the lateral load imposed to the semi-rigid or rigid
et al., 2011; He, 2016; Futai et al., 2018). The figure also shows the
piles is not only taken by the lateral soil resistance but also the base
normalised API (2014) p-y curve, which does not consider the effect of
shear, base moment and vertical shaft stress, which cannot be captured
pile diameter, predicts a softer behaviour than all the piles (ranging from
with sole consideration of p-y curves. Quantitatively, it can be seen from
flexible to rigid piles) simulated in this study. This has necessitated the
Fig. 6(c) that the base shear, base moment and vertical shaft shear stress
development of new p-y formulation which considers the effect of pile
contribute to approximately 25% of the ultimate lateral capacity for the
diameter.
monopile with L/D ¼ 3.
Attempts were made to fit the computed p-y curves (see Fig. 8(a))
The contribution of the base shear, base moment and vertical shaft
with three functions that typically for this purpose, i.e., parabolic
shear stress to the afore-mentioned flexible, semi-rigid and rigid piles
function (Matlock, 1970), hyperbolic function (Georgiadis et al., 1992;
are quantified based on the 3D FEA results, as shown in Fig. 7. As
Dewaikar and Patil, 2006) and hyperbolic tangent function (Gabr et al.,
anticipated, the contribution of the three resistance components in­
1994; Jeanjean, 2009; Truong and Lehane, 2017). Among the three
creases with a decrease in L/D. Among these resistance components, the
choices, the hyperbolic tangent function (shown in Eq. (3)) best captures
base shear shows the greatest contribution to the ultimate lateral pile
all the computed p-y curves (see Fig. 8(b)), and is thus adopted for the
capacity (up to 13%), followed by the base moment (up to 8%). While
p-y formulation in this study:
the contribution of the vertical shaft shear stress (smaller than 5%)

8
L. Wang et al. Ocean Engineering 212 (2020) 107492

Table 4
Summary of the proposed ‘p-y þ MR-θR’ model for monopile analysis.
Model Formulation

p-y curves:
� � � �
p y b
¼ tanh a
pu D
�� � � � �
D 2 D
a ¼ 0:25 þ0:23 þ7:15
Dref Dref
� �
D
b ¼ 0:017 þ 0:53
Dref
p u ¼ Np s u D
� z�
Np ¼ 12 4 exp ξ
D
ξ ¼ 0:25 þ 0:05λ � 0:55
λ ¼ su0 =ðsu1 DÞ
Depth of the MR-θR spring (rotation point): z ¼ 0.8L
MR-θR curve:
MR
¼ tanh½33:5ðθR Þ0:73 �
Mu
� �
Mu D
¼ 0:0124 þ 0:20
su 0:8L DL2 Dref

Note: The lateral pile displacement at the rotation point is constrained to be 0.

�� �2 � � �
Table 5 a ¼ 0:25
D
þ 0:23
D
þ 7:15 (4)
Contributions of the vertical shaft shear stress to the lateral pile capacity. Dref Dref
L/D ratio Pile rigidity Percentage contribution � �
D
30 flexible 0.93% b ¼ 0:017 þ 0:53 (5)
Dref
15 flexible 0.99%
10 Semi- rigid 1.36%
7.5 Semi- rigid 2.00% where Dref is a reference pile diameter of 1 m.
6 Semi- rigid 3.06% After determining the functional form of the p-y curves with
5 Semi- rigid 4.02% consideration of pile geometry (see Eqs. (3)–(5)), the remaining issue is
4.3 Rigid 4.75%
concerned with the value of ultimate soil resistance pu which can be
3.8 Rigid 5.14%
3.3 Rigid 5.39% calculated by:
3 Rigid 5.60%
pu ¼ Np su D (6)

� � �� Fig. 10 shows the computed distributions of the lateral bearing ca­


p
¼ tanh a
y b
(3b) pacity factor (Np) of each pile within shallow depth, where the wedge
pu D and full-flow types of mechanisms are involved (i.e., the depth relevant
to the p-y curves). In the figure, the Np distributions recommended by
where a and b are two fitting parameters. It is found that the values of a
API (2014), Stevens and Audibert (1979), Jeanjean (2009) and Truong
and b for all the piles considered in this study are well correlated to the
and Lehane (2017) are also included for comparison. The equations for
pile diameter D through a quadratic polynomial (as shown in Fig. 9(a)
these Np distributions are summarized in Table 3. It can be seen that the
and formulated in Eq. (4)) and a linear relationship (as shown in Fig. 9
computed distribution of Np for piles with varied diameters broadly
(b) and formulated in Eq. (5)), respectively:
merge into a single trend, because of identical soil flow mechanisms

Table 6
Parameter of pile test adopted in this study for model validation.
Pile parameter Field tests (Zhu et al., 2017) Centrifuge tests (Lai et al., 2019, 2020) Centrifuge tests (Murali et al., 2015, 2019)

Pile diameter, D (m) 2.2 4 6 3.47


Pile embedded length, L (m) 57.4 52.5 60 7.1
Pile wall thickness, t (m) 0.03 0.02 0.042
Load eccentricity, h (m) 12.64 13.53 8 4.16 12.15 5.21 8.68
Pile elastic modulus, Ep (GPa) 210 72 72
L/D ratio 26 24 15 10 2
Pile rigidity Flexible flexible semi-rigid rigid

Note: All the dimensions of the centrifuge tests are in prototype.

9
L. Wang et al. Ocean Engineering 212 (2020) 107492

Fig. 11 shows the change of the depth of the rotation point (normalised
by pile embedded depth L) with an increase of lateral pile-head
displacement. As the pile-head is laterally displaced, the depth of the
rotation point shifts downwards, tending to stabilize at a depth of about
z ¼ 0.8L regardless of D and L/D ratio. This depth (i.e., z ¼ 0.8L) falls
within the reported range of depth for the rotation point of laterally
loaded piles with varied L/D, i.e., z ¼ 0.75–0.80L (Achmus et al., 2009;
Klinkvort and Hededal, 2011; Zhang et al., 2013; Ahmed and Hawlader,
2016; He, 2016; Truong and Lehane, 2017; Murphy et al., 2018).
Fig. 11 also suggests that, despite the wide range of load eccentricity
considered in this study (h ¼ 2D to 7D), the resulted depths of rotation
point in all the analyses are almost identical (approximately 0.8L below
the mudline). It is thus decided to adopt a fixed depth of z ¼ 0.8L for the
rotation points.

5.3. Generalizing MR-θR relation

To generalize the MR-θR relation at the rotation point, the computed


relationship between the pile bending moment (MR) and the rotation
angle (θR) at the rotation point (z ¼ 0.8L) for piles with varied geome­
tries (i.e., D ¼ 3–10 m) are shown in Fig. 12(a). It is worth noting that the
Fig. 14. Validation against field tests performed by Zhu et al. (2017) on computed MR-θR curves for the piles with D ¼ 1 and 2 m (i.e., L/D ¼ 30
pile-head load-displacement relationship of the flexible pile GK04 (D ¼ 2.2 m, and 15) are not included in the figure, because of zero rotation at z ¼
L/D ¼ 26) and GK08 (D ¼ 2.2 m, L/D ¼ 24). 0.8L for these two flexible piles. As illustrated, for a constant pile
embedded depth in this study, i.e., L ¼ 30 m, the initial stiffness and the
involved in the piles and depths of interests, i.e., wedge or full-flow ultimate bending moment (Mu) both increase with pile diameter. This is
mechanism). Among the exiting approaches, which all appear to because the lateral soil resistance and the contribution of the base shear
broadly capture the computed Np distribution with depth, Jeanjean and base moment to the bending becomes more pronounced for piles
(2009)’s equation offers the best prediction, and is thus adopted in the with a larger diameter, as evident in Fig. 7 of this study and reported
proposed p-y formulation (see Eq. (3)). elsewhere (Byrne et al., 2017; Murphy et al., 2018).
It can also be seen from Figs. 9 and 10 that a single set of parameters By normalizing each MR-θR curve in Fig. 12(a) with its ultimate Mu
including a and b (featuring the shape of p-y) and Np (describing the value, all the results for piles with varied D or L/D ratios converge to a
ultimate soil resistance) can satisfactorily describe the p-y curves of piles single trend, as shown in Fig. 12(b). This trend can be best captured by
with varying load eccentricity (h ¼ 2D to 7D) in a unified manner. It is the following hyperbolic tangent function:
also found by Zhu et al. (2018) through a numerical study that the p-y
MR � �
curves of a pile in soft clay were merely altered by a significant variation ¼ tanh 33:5ðθR Þ0:73 (8)
Mu
of h between 20 and 35D.
The remaining issue is to formulate Mu as a function of pile geome­
5.2. Determination of rotation point tries (i.e., D and L). A careful inspection of the computed results in
Fig. 13, has led to the following linear correlation between normalised
In addition to the p-y formulation, another key element of the pro­ Mu (by su-0.8LDL2, where su-0.8L is undrained shear strength at z ¼ 0.8L)
posed ‘p-y þ MR-θR’ model is the depth of the rotation point (or the point and pile diameter D:
with minimum lateral displacement), where the MR-θR spring is located.

Fig. 15. Validation against field tests performed by Zhu et al. (2017) on (a) pile bending moment profiles and (b) pile displacement profiles of the flexible pile GK04
(D ¼ 2.2 m, L/D ¼ 26).

10
L. Wang et al. Ocean Engineering 212 (2020) 107492

Fig. 16. Validation against field tests performed by Zhu et al. (2017) on (a) pile bending moment profiles and (b) pile displacement profiles of the flexible pile GK08
(D ¼ 2.2 m, L/D ¼ 24).

naturally recovers to the conventional p-y model. In other words, the


proposed ‘p-y þ MR-θR’ model has the flexibility of adjusting the relative
contribution of p-y (i.e., lateral soil resistance) and MR-θR (resistance by
base shear and base moment) to the lateral behaviour of piles, according
to pile rigidity.

5.4. Discussion

It is worth noting that the two-spring model developed in this study


includes 3 of the 4 resistance components that are proposed in the PISA’s
model, namely (i) lateral soil resistance; (ii) base shear and (iii) base
moment. While the resistance component of vertical shaft shear stress is
ignored for simplicity. The impact of this simplification is further
examined in Fig. 6, by comparing the computed load-displacement re­
lations at the pile head with those predicted by the newly proposed two-
spring model. The difference between the computed and the predicted
lateral pile capacity is resulted from the ignorance of the vertical shaft
shear stress in the two-spring model.
It can be seen from Fig. 6(a) that the proposed model well repro­
duced the load-displacement response of the flexible pile (L/D ¼ 15),
suggesting the lateral behavior of flexible pile is merely affected by the
vertical shaft shear stress. Similar conclusion was also made by Zhang
Fig. 17. Validation against centrifuge tests performed by Lai et al. (2019, 2020) and Andersen (2019). When L/D decreases, as shown in Fig. 6(b) and (c),
on pile-head load-displacement relationship of the flexible monopile (D ¼ 4 m,
the two-spring model underestimates the pile capacity due to the
L/D ¼ 15) and the semi-rigid monopile (D ¼ 6 m, L/D ¼ 10).
omission of the contribution from pile vertical shaft shear stress. The
� � percentage underestimation for lateral pile capacity due to the igno­
Mu
¼ 0:015
D
þ 0:25 (9) rance of vertical shaft shear stress increases as the L/D ratio decreases.
This has also been observed by Byrne et al. (2017) through field pile tests
su 0:8L DL 2 Dref

In other words, the value of Mu for piles can be readily calculated and Murphy et al. (2018) through 3D FEA.
from Eq. (9), simply based on pile geometries (i.e., D and L) and shear To quantify when the simplification in the two-spring model might
strength profile. impact the response, the percentage contributions of the vertical shaft
It can also be seen from Figs. 12 and 13 that the choice of the shear stress to the lateral capacity for all the analyses reported herein
different load eccentricity considered in this study (h ¼ 2D to 7D) has (where L/D ¼ 3 to 30, see Table 1) are extracted, as summarized in
little influence on the MR-θR formulations. Table 5.
By combining the formulations in the above sub-sections, the pro­ It can be seen that for flexible piles in this study (10 < L/D � 30), the
posed ‘p-y þ MR-θR’ model for piles with varied L/D has been readily effect of vertical shaft shear stress is likely to be negligible, with its
available. Table 4 summarizes the equations for each component of the percentage contribution to the lateral capacity smaller than 1%. On the
model, including the p-y formulation, location of the ration point and the other hand, the ignorance of the vertical shaft shear stress might impact
MR-θR formulation at the rotation point. the response for semi-rigid and rigid piles (3 ¼ L/D � 10), where its
It is worth noting that for a flexible pile, which experiences zero percentage contribution to the lateral capacity can be up to 6%.
rotation at the ‘rotation point’, the term MR-θR vanishes (according to The above quantitative conclusions were drawn from analyses of
Eq. (8)). Under this circumstance, the proposed ‘p-y þ MR-θR’ model laterally loaded piles in normally consolidated soft clay. Comparatively,
a less pronounced contribution of vertical shaft shear stress is

11
L. Wang et al. Ocean Engineering 212 (2020) 107492

Fig. 18. Validation against centrifuge tests performed by Lai et al. (2019, 2020) on (a) pile bending moment profiles and (b) pile displacement profiles of the flexible
monopile (D ¼ 4 m, L/D ¼ 15).

Fig. 19. Validation against centrifuge tests performed by Lai et al. (2019, 2020) on (a) pile bending moment and (b) pile displacement of the semi-rigid monopile (D
¼ 6 m, L/D ¼ 10).

anticipated for piles in over-consolidated clay (stiff clay), due to the 6.1. Field tests on flexible piles in soft clay (Zhu et al., 2017)
developed tension gap at the rear of the pile (Randolph and Gourvenec,
2011; Zhang et al., 2016) which reduces the contact area between the Zhu et al. (2017) performed field tests on two laterally loaded piles,
soil and the pile shaft. which were driven offshore in soft clay at Guishan Offshore Wind Farms
in Guangdong Province, China. The two piles, referred as pile GK04 and
6. Validation of the proposed ‘p-y þ MR-θR’ model GK08, were composed of Q345B steel, with a pile diameter of 2.2 m and
a wall thickness of 0.03 m. The embedded length L of pile GK04 and
This section aims to verify the predictive capability of the proposed GK08 are 57.4 and 52.5 m, respectively. The load eccentricities for pile
‘p-y þ MR-θR’ model, against published test results on monopiles with a GK04 and GK08 are 12.6 and 13.5 m, respectively. The dominating soil
broad range of length-to-diameter ratios (L/D) in soft clay. These have strata supporting pile GK04 and GK08 are soft clay. The in-situ un­
included field tests on flexible pile with L/D ¼ 24 and 26 in soft clay drained shear strength (su) of the soft clay, which was characterized
(Zhu et al., 2017), centrifuge tests in soft clay involving semi-rigid pile through cone penetration tests (CPTs), increases almost linearly with
with L/D ¼ 10 (Lai et al., 2019, 2020) and rigid pile with L/D ¼ 2 depth z, i.e., su ¼ 5 þ 0.75z (Zhu et al., 2017). The measured pile
(Murali et al., 2015, 2019). The geometric and stiffness parameters of displacement profiles under lateral loading indicate that pile GK04 and
the piles in the field and centrifuge experiments are summarized in GK08 behave as flexible piles.
Table 6. The performance of the proposed ‘p-y þ MR-θR’ model is eval­ Fig. 14 compares the measured lateral head displacements of the two
uated in the following sub-sections, and compared to that of the API piles, and the predicted results using the proposed ‘p-y þ MR-θR’ model
(2014)’s p-y model. as well as API (2014)’s p-y model. The proposed ‘p-y þ MR-θR’ model is

12
L. Wang et al. Ocean Engineering 212 (2020) 107492

Fig. 20. Validation against centrifuge tests performed by Murali et al. (2015, 2019) on load-displacement relationship of the rigid monopile (D ¼ 3.47 m, L/D ¼ 2) in
(a) clay bed 1 and (b) clay bed 2.

shown to reasonably capture the measured pile head responses, while 6.3. Centrifuge tests on rigid piles in soft clay (Murali et al., 2015, 2019)
the lateral pile responses predicted by API (2014) appear to be
over-conservative. In addition to the pile head response, the adoption of Murali et al. (2015, 2019) presented four centrifuge tests on lateral
the proposed ‘p-y þ MR-θR’ model has also led to reasonable predictions behaviour of a rigid monopile (L/D ¼ 2, where L and D are 7.1 and 3.47
for bending moment and lateral pile displacement profiles of the flexible m in prototype) in normally consolidated soft clay. Four eccentricities of
piles GK 04 (see Fig. 15) and GK08 (see Fig. 16). On the other hand, the lateral load above the ground surface, i.e., h ¼ 1.2, 1.5, 2.5, and 3.5D
API (2014) method overestimates the maximum pile displacement (up were considered in their tests. The pile load tests with the eccentricity of
to 120%) and bending moment (up to 23%) of the piles. 1.2 and 3.5D were carried out in clay bed 1, while the remaining were
conducted in clay bed 2. The undrained shear strength profiles of the
6.2. Centrifuge tests on flexible and semi-rigid piles in soft clay (Lai et al., clay beds 1 and 2 are approximately su ¼ 1 þ 1.1z and su ¼ 1 þ 1.3z,
2019, 2020) respectively.
Fig. 20 compares the measured load-displacement relationships at
Lai et al. (2019, 2020) reported centrifuge tests on two laterally the head of the four rigid piles, and the predicted results using the
loaded large-diameter long monopiles in normally consolidated soft proposed ‘p-y þ MR-θR’ model as well as API (2014)’s model. The pro­
clay. The two piles have an identical embedded depth (L ¼ 60 m in posed ‘p-y þ MR-θR’ model has broadly captured the lateral behaviour of
prototype) but different diameters (D ¼ 4 and 6 m in prototype), leading the four extremely short and rigid piles (L/D ¼ 2), due to proper
to L/D ratios of 10 and 15. The model pile was made of type ‘7075-T6’ consideration of the contributions from base shear and base moment
aluminium alloy pipe (Young’s modulus ¼ 72 GPa) with a thickness of 2 which are essential for piles with small L/D. With ignorance of the
mm (t ¼ 0.2 m in prototype). Each pile was instrumented with 27 levels contributions from the pile base, API (2014)’s p-y model has predicted a
of full Wheatstone bridge strain gauges to obtain the bending moment much softer lateral response of the four rigid piles.
profile. The soil used in the centrifuge tests was Malaysia kaolin, which
is the same as the soil simulated in the numerical parametric analyses 7. Summary and conclusions
reported herein. The measured strength profile exhibits an approxi­
mately linear increase of su with depth. The increasing rate of su per unit This study proposed a mechanism-based, unified ‘p-y þ MR-θR’ model
depth is 1.54 kPa/m (in prototype scale). According to the criterion for predicting lateral behaviour of monopiles in soft clay with varied
proposed by Poulos and Hull (1989), the pile with D of 4 and 6 m are length-to-diameter (L/D) ratios. It considers all the important compo­
characterized as a flexible pile and a semi-rigid pile, respectively. More nents contributing to the lateral resistance of a monopile, namely lateral
details are given in Lai et al. (2019, 2020). soil resistance, base shear and base moment. In the proposed approach,
Fig. 17 compares the measured and predicted pile-head load-dis­ the lateral soil resistance above the rotation point is described by a p-y
placements of the flexible (L/D ¼ 15) and the semi-rigid monopile (L/D model. While the resistance below the rotation point including the pile
¼ 10). The predicted results are based on the proposed ‘p-y þ MR-θR’ base shear and base moment, which is important for piles with relatively
model and API (2014)’s p-y model. It can be seen that the API (2014)’s small L/D, are integrated into a moment-rotation spring (characterized
p-y model underestimates the initial stiffness and capacity for both piles, by a MR-θR model) at the rotational point.
with the level of underestimation increasing with pile diameter. While The formulation of the proposed ‘p-y þ MR-θR’ model, including the
the computed load-displacement relations by the proposed ‘p-y þ MR-θR’ p-y model, the MR-θR model and the location of the rotation point, are
model show satisfactory agreements with the measured data of both proposed based on a series of well-verified 3D finite element (FE) ana­
flexible and semi-rigid piles. In addition, as can be seen in Fig. 18 and lyses considering a wide range of L/D. The formulated ‘p-y þ MR-θR’
Fig. 19, the bending moment and lateral displacement profiles of the model can then be applied to a 1D Winkler model for efficient compu­
flexible and semi-rigid piles at different loading stages have been tation with comparable accuracy to the underlying 3D FE model.
reasonably reproduced by the proposed ‘p-y þ MR-θR’ model, with a The proposed ‘p-y þ MR-θR’ model is capable of capturing the
maximum percentage error of 15%. However, using API (2014) method behaviour of monopiles with varied L/D ratios in a unified manner,
results in an overestimation in maximum pile bending moment (up to because of its ability to adjust the relative contribution of p-y and MR-θR
35%) and pile displacement (up to 120%). to the lateral behaviour of piles considering pile diameter effects. It
naturally recovers to a p-y model while analyzing flexible piles, where θR

13
L. Wang et al. Ocean Engineering 212 (2020) 107492

¼ 0 at the rotation point. DNVGL, 2016. DNVGL-ST-0126-Support Structure for Wind Turbines. Det Norske
Veritas, Oslo.
The validity of the model has been well justified against a number of
Doherty, P., Gavin, K., 2012. Laterally loaded monopile design for offshore wind farms.
field and centrifuge model tests in soft clay embedded with laterally Proc. Instit. Civil Eng Energy 165 (1), 7–17.
loaded monopiles, which cover a broad range of L/D ratios (including EWEA, 2016. The European offshore wind industry. Key Trends Stat 2017, 33, 2016.
flexible, semi-rigid and rigid piles). On the other hand, the standard p-y Finn, W.D.L., Dowling, J., 2016. Modelling effects of pile diameter. Can. Geotech. J. 53
(1), 173–178.
method (e.g. API, 2014) significantly under-predict the lateral stiffness Futai, M.M., Dong, J., Haigh, S.T., Madabhushi, S.P.G., 2018. Dynamic response of
and capacity of the piles in these field and centrifuge model tests, monopiles in sand using centrifuge modelling. Soil Dynam. Earthq. Eng. 115,
particularly for piles with low L/D ratios. Adoption of the proposed 90–103.
Gabr, M.A., Lunne, T., Powell, J.J., 1994. p–y analysis of laterally loaded piles in clay
model, therefore, has shown the potential in reduced design conserva­ using DMT. J. Geotech. Eng. 120 (5), 816–837.
tism and better economies for future wind farms. Gao, F.P., Li, J.H., Qi, W.G., Cun, H., 2015. On the instability of offshore foundations:
theory and mechanism. Sci. China Phys. Mech. Astron. 58 (12), 124701.
Georgiadis, M., Anagnostopoulos, C., Saflekou, S., 1992. Cyclic lateral loading of piles in
Declaration of competing interest soft clay. Geotech. Eng. 23 (1), 47–60.
Gerolymos, N., Gazetas, G., 2006. Development of Winkler model for static and dynamic
The authors declare that they have no known competing financial response of caisson foundations with soil and interface nonlinearities. Soil Dynam.
Earthq. Eng. 26 (5), 363–376.
interests or personal relationships that could have appeared to influence Gudehus, G.A., 1996. Comprehensive constitutive equation for granular materials. Soils
the work reported in this paper. Found. 36, 1–12.
He, B., 2016. Lateral Behaviour of Single Pile and Composite Pile in Soft Clay. Ph.D.
dissertation. Zhejiang University.
CRediT authorship contribution statement He, B., Lai, Y.Q., Wang, L.Z., Hong, Y., Zhu, R.H., 2019. Scour effects on the lateral
behavior of a large-diameter monopile in soft clay: role of stress history. J. Mar. Sci.
Lizhong Wang: Conceptualization, Methodology, Investigation, Eng. 7 (170), 1–23.
Hong, Y., He, B., Wang, L.Z., Wang, Z., Ng, C.W.W., Ma�sín, D., 2017a. Cyclic lateral
Supervision, Writing - review & editing. Yongqing Lai: Methodology, response and failure mechanisms of semi-rigid pile in soft clay: centrifuge tests and
Investigation, Formal analysis, Validation, Writing - original draft. Yi numerical modelling. Can. Geotech. J. 54 (6), 86–824.
Hong: Methodology, Investigation, Formal analysis, Writing - review & Hong, Y., Wang, L.Z., Ng, C.W.W., Yang, B., 2017b. Effect of initial pore pressure on
undrained shear behaviour of fine-grained gassy soil. Can. Geotech. J. 54 (11),
editing. David Ma�sín: Software, Writing - review & editing.
1592–1600.
Hong, Wang, L.Z., Yang, B., Zhang, J.F., 2019. Stress-dilatancy of bubbled fine-grained
Acknowledgements sediments. Eng. Geol. 260, 1–7.
Hong, Y., Wang, L.Z., Zhang, J.F., Gao, Z.W., 2020. 3D elastoplastic model for fine-
grained gassy soil considering the gas-dependent yield surface shape and stress-
The authors gratefully acknowledge the financial supports provided dilatancy. J. Eng. Mech. 146 (5), 04020037.
by National Key Research and Development Program Ilyas, T., Leung, C.F., Chow, Y.K., Budi, S.S., 2004. Centrifuge model study of laterally
loaded pile groups in clay. J. Geotech. Geoenviron. Eng. 130 (3), 274–283.
(2018YFE0109500), National Natural Science Foundation of China Jeanjean, P., 2009. Re-assessment of p-y curves for soft clays from centrifuge testing and
(51939010 and 51779221), the Key Research and Development Pro­ finite element modeling. In: Proc. Offshore Technology Conf. Paper OTC20158,
gram of Zhejiang Province (2018C03031) and Joint Fund of Ministry of Houston.
Jeong, S., Kim, Y., Kim, J., 2011. Influence on lateral rigidity of offshore piles using
Education for Pre-research of Equipment (6141A02022137).
proposed p–y curves. Ocean Eng. 38, 397–408.
Klinkvort, R.T., Hededal, O., 2011. Centrifuge modelling of offshore monopile
References foundation. In: Frontiers in Offshore Geotechnics II, 1 ed. Taylor & Francis,
pp. 581–586.
Lai, Y.Q., Wang, H., Wang, L.Z., Hong, Y., 2019. Experimental Investigation on
Achmus, M., Kuo, Y.S., Abdel-Rahman, K., 2009. Behavior of monopile foundations
Monotonic and Cyclic Lateral Responses of Large-Diameter Monopiles in Sand and
under cyclic lateral load. Comput. Geotech. 36 (5), 725–735.
Soft Clay. A Report to Power China Huadong Engineering Limited Corporation.
Achmus, M., Thieken, K., 2016. Evaluation of p-y approaches for large diameter
Zhejiang University.
monopiles in soft clay. Proc. 26th Int. Ocean Polar Eng. Conf(ISOPE) 805–816.
Lai, Y.Q., Wang, L.Z., Hong, Y., He, B., 2020. Centrifuge modeling of cyclic lateral
Achmus, M., Thieken, K., Saathoff, J.E., Terceros, M., Alboker, J., 2019. Un- and
behavior of large-diameter monopiles in soft clay: effects of episodic cycling and
reloading stiffness of monopile foundations in sand. Appl. Ocean Res. 84, 62–73.
reconsolidation. Ocean Eng. 200, 107048.
Ahmed, S.S., Hawlader, B., 2016. Numerical analysis of large-diameter monopiles in
Lam, I.P.O., 2009. Diameter Effects on P–Y Curves. Deep Foundations Institute,
dense sand supporting offshore wind turbines. Int. J. GeoMech. 16 (5), 04016018.
Hawthorne, N.J.
API, 2014. Recommended Practice Planning, Designing, and Constructing Fixed Offshore
Lam, I.P.O., Martin, G.R., 1986. Seismic Design of High-Way Bridge Foundations. US
Platforms-Working Stress Design. API 2A-WSD, twenty-second ed. (Washington,
Department of Transportation Report No. FHWA/RD-86/102.
DC).
Lau, B.H., 2015. Cyclic Behaviour of Monopile Foundations for Offshore Wind Turbines
Byrne, B.W., McAdam, R.A., Burd, H., Houlsby, G.T., Martin, C.M., Beuckelaers, W.J.A.
in Clay. Ph.D. dissertation. University of Cambridge.
P., Zdravkovic, L., Taborda, D.M.G., Potts, D.M., Jardine, R.J., Ushev, E., Liu, T.,
Leblanc, C., Houlsby, G.T., Byrne, B.W., 2010. Response of stiff piles in sand to long-term
Abadias, G.D., Gavin, K., Igoe, D., Doherty, P., Skov, G.J., Pacheco, A.M., Muir, W.A.,
cyclic lateral loading. Geotechnique 60 (2), 79–90.
Schroeder, F.C., Turner, S., Plummer, M.A.L., 2017. PISA: new design methods for
Ling, L.F., 1988. Back Analysis of Lateral Load Tests on Piles. Civil Engineering Dept.,
offshore wind turbine monopiles. In: Proceedings of the Society for Underwater
Univ. of Auckland, New Zealand. Rep. No. 460.
Technology Offshore Site Investigation and Geotechnics 8th International
Madabhushi, G.S.P., Haiderali, A.E., 2013. Evaluation of the p-y method in the design of
Conference, London.
monopiles for offshore wind turbines. In: Offshore Technology Conference OTC,
Byrne, B.W., Houlsby, G.T., Burd, H.J., Gavin, K., Igoe, D., Jardine, R.J., Martin, C.M.,
pp. 1824–1844.
McAdam, R.A., Potts, D.M., Taborda, D.M.G., Zdravkovi�c, L., 2019a. PISA Design
Ma�sín, D., 2005. A hypoplastic constitutive model for clays. Int. J. Numer. Anal. Methods
Model for Monopiles for Offshore Wind Turbines: Application to a Stiff Glacial Clay
GeoMech. 29, 311–336.
till. G�eotechnique, p. 255. https://doi.org/10.1680/jgeot.18.
Ma�sín, D., 2014. Clay hypoplasticity model including stiffness anisotropy. Geotechnique
Byrne, B.W., McAdam, R.A., Burd, H.J., Beuckelaers, W.J.A.P., Gavin, K., Houlsby, G.T.,
64 (3), 232–238.
Igoe, D., Jardine, R.J., Martin, C.M., Muir Wood, A., Potts, D.M., Skov Gretlund, J.,
Matlock, H., 1970. Correlations for design of laterally loaded piles in clay. In:
Taborda, D.M.G., Zdravkovi�c, L., 2019b. Monotonic lateral loaded pile testing in a
Proceedings of the Offshore Technology Conference, Houston, pp. 577–588. Paper
stiff glacial clay till at Cowden. Geotechnique. https://doi.org/10.1680/jgeot.18.
OTC1204.
pisa.003.
Matsuoka, H., Nakai, T., 1974. Stress-deformation and strength characteristics of soil
Cai, G.J., Liu, S.Y., Tong, L.Y., et al., 2010. Field evaluation of undrained shear strength
under three different principal stresses. Proc. Japanese Soc. Civil Eng. 232, 59–70.
from piezocone penetration tests in soft marine clay. Mar. Georesour. Geotechnol.
McAdam, R.A., Byrne, B.W., Houlsby, G.T., Beuckelaers, W.J.A.P., Burd, H.J., Gavin, K.,
28, 143–153.
Igoe, D., Jardine, R.J., Martin, C.M., Muir Wood, A., Potts, D.M., Skov Gretlund, J.,
Carter, D.P., 1984. A Non-linear Soil Model for Predicting Lateral Pile Response. M.Eng.
Taborda, D.M.G., Zdravkovi�c, L., 2019. Monotonic lateral loaded pile testing in a
Thesis. University of Auckland.
dense marine sand at Dunkirk. Geotechnique. https://doi.org/10.1680/jgeot.18.
Chen, L., Poulos, H.G., 1993. Analysis of pile-soil interaction under lateral loading using
pisa.004.
infinite and finite elements. Comput. Geotech. 15 (4), 189–220.
Murali, M., Grajales, F., Beemer, R.D., Biscontin, G., Aubeny, C., 2015. Centrifuge and
Dewaikar, D.M., Patil, P.A., 2006. A New Hyperbolic P–Y Curve Model for Laterally
numerical modeling of monopiles for offshore wind towers installed in clay. In:
Loaded Piles in Soft Clay. GeoShanghai International Conference, Shanghai.
ASME 2015 34th International Conference on Ocean, Offshore and Arctic
Di, S.J., Shan, Z.G., Wang, M.Y., 2013. Property analysis and applications of in-situ
Engineering. American Society of Mechanical Engineers.
testing shear wave velocity at intertidal zone. Chin. J. Rock Mech. Eng. 32 (10),
2053–2060 (in Chinese).

14
L. Wang et al. Ocean Engineering 212 (2020) 107492

Murali, M., Grajales, F., Beemer, R.D., Aubeny, C.P., Biscontin, G., 2019. Capacity of Wang, X.F., Zeng, X.W., Li, J.L., 2019. Vertical performance of suction bucket foundation
short piles and caissons in soft clay from geotechnical centrifuge tests. J. Geotech. for offshore wind turbines in sand. Ocean Eng. 180, 40–48.
Geoenviron. Eng. 145 (10), 04019079. Winkler, E., 1867. Die lehre von elasticitat und festigkeit, (H. Dominic us), pp. 182–184.
Murphy, G., Igoe, D., Doherty, P., Gavin, K., 2018. 3D FEM approach for laterally loaded Prague.
monopile design. Comput. Geotech. 100, 76–83. Wang, L.Z., Wang, K.J., Hong, Y., 2016. Modeling temperature-dependent behavior of
Nan, J., Zhang, P.Y., Liu, Y.G., Ding, H.Y., 2018. Bearing capacity of composite bucket soft clay. J. Eng. Mech. 142 (8), 04016054.
foundations for offshore wind turbines in silty sand. Ocean Eng. 151, 1–11. Wang, L.Z., Wang, H., Zhu, B., Hong, Y., 2018. Comparison of monotonic and cyclic
Negro, V., L�opez-Guti�errez, J.-S., Esteban, M.D., Alberdi, P., Imaz, M., Serraclara, J.-M., lateral response between monopod and tripod bucket foundations in medium dense
2017. Monopiles in offshore wind: preliminary estimate of main dimensions. Ocean sand. Ocean Eng. 155, 88–105.
Eng. 133, 253–261. Wroth, C., Houlsby, G., 1985. Soil mechanics-property characterization, and analysis
Niemunis, A., Herle, I., 1997. Hypoplastic model for cohesionless soils with elastic strain procedures. In: Proceedings of the 11th International Conference on Soil Mechanics
range. Mech. Cohesive-Frict. Mater. 2, 279–299. and Foundation Engineering, San Francisco, CA, USA, vol. 1, pp. 1–55.
O’Neill, M.W., Gazioglu, S.M., 1984. Evaluation of p–y relationships in cohesive soils. In: Wu, W., Kolymbas, D., 1990. Numerical testing of the stability criterion for hypoplastic
Proceedings of the Symposium on Analysis and Design of Pile Foundations. ASCE, constitutive equations. Mech. Mater. 9, 245–253.
pp. 192–213, 1–5 October. Wu, W., Bauer, E., Kolymbas, D., 1996. Hypoplastic constitutive model with critical state
Poulos, H.G., Hull, T.S., 1989. The role of analytical geomechanics in foundation for granular materials. Mech. Mater. 23, 45–69.
engineering. In: Foundation Engineering: Current Principles and Practices. ASCE, Xie, Y., Leung, C.F., Chow, Y.K., 2012. Centrifuge modelling of spudcan-pile interaction
pp. 1578–1606. in soft clay. Geotechnique 62 (9), 799–810.
Qi, W.G., Gao, F.P., Randolph, M.F., Lehane, B.M., 2016. Scour effects on p-y curves for Zakeri, A., Clukey, E.C., Kebadze, E.B., Jeanjean, P., 2016. Fatigue analysis of offshore
shallowly embedded piles in sand. Geotechnique 66 (8), 648–660. well conductors: Part I-Study overview and evaluation of Series 1 centrifuge tests in
Randolph, M.F., Gourvenec, S., 2011. Offshore Geotechnical Engineering. Taylor & normally consolidated to lightly over-consolidated kaolin clay. Appl. Ocean Res. 57,
Francis. 78–95.
Randolph, M.F., Wroth, C.P., 1981. Application of the failure state in undrained simple Zdravkovi�c, L., Taborda, D.M.G., Potts, D.M., Abadias, D., Burd, H.J., Byrne, B.W.,
shear to the shaft capacity of driven piles. Geotechnique 31, 143–157. Gavin, K., Houlsby, G.T., Jardine, R.J., Martin, C.M., McAdam, R.A., Ushev, E., 2019.
Shadlou, M., Bhattacharya, S., 2016. Dynamic stiffness of monopiles supporting offshore Finite element modelling of laterally loaded piles in a stiff glacial clay till at Cowden.
wind turbine generators. Soil Dynam. Earthq. Eng. 83, 15–32. Geotechnique. https://doi.org/10.1680/jgeot.18.pisa.005.
Stevens, J.B., Audibert, J.M.E., 1979. Re-examination of p-y curve formulations. In: Zhang, G., Rong, B., Fu, P., 2013. Centrifuge model test study of static and cyclic
Proceedings of the 11th Annual Offshore Technology Conference, Houston, TX, USA, behavior of a pile foundation for an offshore wind generator. J. Test. Eval. 41 (5),
22–24 April 1970, pp. 397–403. 801-712.
Stone, K.J.L., Arshi, H.S., Zdravkovic, L., 2018. Use of a bearing plate to enhance the Zhang, P.Y., Guo, Y., Liu, Y., Ding, H., 2016. Experimental study on installation of hybrid
lateral capacity of monopiles in sand. J. Geotech. Geoenviron. Eng. 144 (8), bucket foundations for offshore wind turbines in silty clay. Ocean Eng. 114, 87–100.
04018051. Zhang, L.Y., Ahmari, S., 2011. Nonlinear analysis of laterally loaded rigid piles in
Taborda, D.M.G., Zdravkovi�c, L., Potts, D.M., Burd, H.J., Byrne, B.W., Gavin, K., cohesive soil. Int. J. Numer. Anal. Methods GeoMech. 37 (2), 201–222.
Houlsby, G.T., Jardine, R.J., Liu, T., Martin, C.M., McAdam, R.A., 2019. Finite Zhang, Y.H., Andersen, K.H., 2019. Soil reaction curves for monopiles in clay. Mar.
element modelling of laterally loaded piles in a dense marine sand at Dunkirk. Struct. 65, 94–113.
Geotechnique. https://doi.org/10.1680/jgeot.18.pisa.006. Zhang, Y.H., Andersen, K.H., 2017. Scaling of lateral pile p-y response in clay from
Truong, P., Lehane, B.M., 2017. Effects of pile shape and pile end condition on the lateral laboratory stress-strain curves. Mar. Struct. 53, 124–135.
response of displacement piles in soft clay. Geotechnique 68 (9), 794–804. Zhang, Y.H., Andersen, K.H., Tedesco, G., 2016. Ultimate bearing capacity of laterally
Wang, L.Z., He, B., Hong, Y., Guo, Z., Li, L.L., 2015. Field tests of the lateral monotonic loaded piles in clay-some practical considerations. Mar. Struct. 50, 260–275.
and cyclic performance of jet-grouting reinforced cast-in-place piles. J. Geotech. Zhu, B., Wen, K., Kong, D., Zhu, Z., Wang, L., 2018. A numerical study on the lateral
Geoenviron. Eng. 141 (5), 06015001. loading behaviour of offshore tetrapod piled jacket foundations in clay. Appl. Ocean
Wang, X.F., Yang, X., Zeng, X.W., 2017. Lateral response of improved suction bucket Res. 75, 165–177.
foundation for offshore wind turbine in centrifuge modelling. Ocean Eng. 141, Zhu, B., Zhu, Z.J., Li, T., Liu, J.C., Liu, Y.F., 2017. Field tests of offshore driven piles
295–307. subjected to lateral monotonic and cyclic loads in soft clay. J. Waterw. Port, Coast.
Ocean Eng. 143 (5), 05017003.

15

You might also like