0% found this document useful (0 votes)
130 views17 pages

A Novel Hybrid Monopile Foundation For Offshore Wind Turbines PDF

Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
130 views17 pages

A Novel Hybrid Monopile Foundation For Offshore Wind Turbines PDF

Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 17

Ocean Engineering 198 (2020) 106963

Contents lists available at ScienceDirect

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

A novel hybrid monopile foundation for offshore wind turbines


Hongwang Ma a, *, Jun Yang a, b
a
School of Naval Architecture, Ocean and Civil Engineering, Shanghai Jiao Tong University, Shanghai, China
b
Department of Civil Engineering, The University of Hong Kong, Hong Kong, China

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

Keywords: This paper presents a hybrid monopile foundation for offshore wind turbines. It is an ultra-high performance
Offshore wind turbine concrete-filled double skin steel tubular structure (CFDST) used as a replacement of the conventional steel tube
Monopile between the water level and the mudline so as to reduce the monopile diameter and thereby reducing the wave
Concrete-filled double skin steel tubular
loads on the pile. To study the feasibility of this monopile, the NREL 5 MW wind turbine supported by a con­
structure
Accumulated rotation
ventional monopile is selected as a reference and a three-dimensional (3D) finite element model is developed.
Natural frequency The natural frequency, the various responses under the serviceability limit state (SLS) and the ultimate limit state
(ULS) of the hybrid monopile are presented. Particularly, the effect of varying outer diameter of the CFDST on
the structural performance is investigated. By applying a consistent accumulated rotation at the mudline under
the SLS, the natural frequency is found to be within a desired range, and an optimized embedded length of the
hybrid monopile is determined. The results indicate that the proposed hybrid monopile is able to meet the design
requirements for both SLS and ULS and the optimization of pile embedded length leads to an efficient and
economic monopile foundation for offshore wind turbines.

1. Introduction turbines, the foundation cost amounts to over 20% of the total capital
cost (Gentils et al., 2017; Kim and Kim, 2018). Therefore, there is an
Offshore wind energy has become one of the fastest growing sus­ urgent need to develop a cost effective foundation for offshore wind
tainable energy sources. At the end of 2017, the global capacity of turbines.
offshore wind energy reached over 18.8 GW (GW), nearly 84% of which At present, different types of foundations exist for offshore wind
is located in European countries and the remaining 16% is located turbines, such as monopiles, tripods, suction caissons, jackets and
mainly in China (GWEC, 2017). The United Kingdom maintains the gravity foundations (O’Kelly and Arshad, 2016). Among them, monopile
largest offshore wind market, accounting for over 36% of installed ca­ is the dominant foundation type due to several advantages (O’Kelly and
pacity, followed by Germany (28.5%), while China ranks third in the Arshad, 2016; Alamo
� et al., 2018). Generally, a monopile comprises of a
global offshore rankings, accounting for 15% (GWEC, 2017; REN21, transition piece and a single large diameter open-ended tubular steel
2017). WindEurope expects that by the year 2020, the total European pipe that is driven, drilled or vibrated into the seabed. Typically,
offshore wind capacity will reach 25 GW (GWEC, 2017), while China’s monopile diameters vary from 4 to 6 m, with the slenderness ratio (ratio
offshore wind target will reach 5 GW (Ou et al., 2018). Although between the embedded length and the diameter) between 4 and 8. Most
offshore wind energy has undergone a rapid development in recent of monopiles are installed in water depths not exceeding 25 m (Murphy
years, the Levelized Cost of Energy (LCOE) for offshore wind turbines is et al., 2018; Scharff and Siems, 2013; Negro et al., 2017). As wind farms
still much higher than that of onshore wind turbines, and it is also move further offshore, wind turbines will need to move into much
significantly higher than the LCOE of conventional power plants, such as deeper waters and withstand greater loads from waves and winds. As a
energy from coal or gas (Andres et al., 2017; Gentils et al., 2017). If result, the cost can increase dramatically. Therefore, many studies have
offshore wind turbines are installed in deeper waters and on larger been performed with the aim of reducing the cost of monopiles through
scales, the LCOE will increase further (Schwanitz and Wierling, 2016). structural optimization (Gentils et al., 2017; Muskulus and Schafhirt,
This makes it difficult for offshore wind energy to have a competitive 2014; Schmoor and Achmus, 2015; Gjersøe et al., 2015; Kallehave et al.,
price in the energy market (Bocher et al., 2018). For offshore wind 2015; Rad et al., 2014). While the structural optimization may

* Corresponding author.
E-mail address: [email protected] (H. Ma).

https://doi.org/10.1016/j.oceaneng.2020.106963
Received 23 May 2019; Received in revised form 14 November 2019; Accepted 15 January 2020
Available online 23 January 2020
0029-8018/© 2020 Elsevier Ltd. All rights reserved.
H. Ma and J. Yang Ocean Engineering 198 (2020) 106963

Tower

Tower Bottom flange


Wind
c Transition piece
Ltp L
turbine
tp
Steel tube
A A A-A H gap
Tower
Steel tube
B B tc tc
B-B MSL
(b)
Tower bottom
Transition Inner steel tube
piece MSL
Inner steel tube
Outer steel tube Concrete
Dinner
Mudline A Concrete
A A A
Hw Outer steel tube
Pile
Steel tube
B B D out
B B

(a) (c)

Fig. 1. Schematic illustration of the hybrid and conventional monopile:(a) a Ltp


monopile supported wind turbine system; (b) conventional monopile; (c)
Mudline
hybrid monopile.

D
contribute to finding better and more economic solutions, the tendency
of rapid increase in monopile cost with the increase in water depth Pile
cannot change.
In the design of monopiles, the wave and wind loads are the critical
concerns (Arany et al., 2017; Morato � et al., 2017; Stansby et al., 2013). Fig. 2. Details of the CFDST
Generally, the proportions of wave and wind loads are related to the
water depth, and the wave loads become significant for greater water has a high load-bearing capacity, high rigidity, good energy absorption,
depths and wave heights, which results in a larger area applying wave high fire resistance, and construction cost-effectiveness (Wang et al.,
loads and a larger moment arm for those wave loads, leading to larger 2018; Li et al., 2018). Therefore, the use of CFDST instead of steel
mudline stresses in the support structure. Moreover, wave load depends tubular can fulfill the capacity and stiffness requirements with a reduced
not only on the marine environment, but also on the size of the pile diameter. If ultra-high performance concrete (UHPC) is used instead
monopile. According to DNV⋅GL (2016), the total horizontal force on a of normal strength concrete in CFDST, the sectional size of the monopile
monopile due to waves consists of drag force and inertia force. The drag can be further reduced (Zhang, 2017; Chen et al., 2018).
force is directly proportional to the diameter of the monopile whereas In this paper, we present a feasibility study using CFDST as part of a
the inertia force is directly proportional to the square of the diameter of monopile foundation for offshore wind turbines. The National Renew­
the monopile. In this respect, a reduction in the diameter of the able Energy Laboratory (NREL) 5 MW offshore wind turbine (Jonkman
monopile can reduce the wave load on the monopile. et al., 2009) is adopted as a reference model. We investigated the
The main design considerations for a monopile-supported wind structural responses using the finite element method (FEM) for two
turbine include natural frequency, stability, structural strength and fa­ foundation types: one is the conventional monopile and the other is the
tigue, as well as allowable deformations of the system during operation proposed hybrid monopile. Detailed comparisons are made for both the
(DNV.GL, 2016; Velarde, 2016), typically governed by permanent de­ SLS and the ULS. Furthermore, attempt is made to optimize the
formations under the SLS and natural frequency (Velarde, 2016; embedded length of the hybrid monopile using the accumulated rotation
Schmoor and Achmus, 2015; Senanayake, 2016). In order to reduce the at mudline and the natural frequency as design-driving parameters.
wave loads by reducing the diameter of the monopile while still satis­
fying the design requirements, two measures have been developed: one 2. Concept description for hybrid monopile
is to reduce the diameter of the monopile only for the portion between
the mudline and the water level, and the other is to use concrete-filled The key idea here is to reduce the diameter of the portion of the
double skin steel tubular structure (CFDST) for that portion of the monopile between the mudline and the water level by replacing the pure
monopile. CFDST can be regarded as a composite member that is con­ steel tubular structure with the CFDST such that the wave load can be
structed by filling concrete between two concentrically placed steel reduced. Fig. 1 compares the geometric shapes of conventional and
tubes. It utilizes advantages of both steel and concrete (Han, 2016; Wang hybrid monopiles. The conventional monopile foundation consists of the
et al., 2018; Zhang, 2017), and has been considered as loading bearing pile and the transition piece that connects the pile to the tower. The
elements in some structural engineering applications, such as high piers proposed hybrid foundation consists of the pile and the CFDST. The
for bridges, columns for high-rise buildings, support columns for outer steel tube of CFDST can be fabricated together with the pile
offshore platforms, and transmissions towers (Han et al., 2011; Hassa­ through a conical steel tube. In practice, the mechanical shear connec­
nein et al., 2018; Uenaka, 2016). Compared to pure steel tubular, CFDST tors (e. g. Eom et al., 2019; Shimizu et al., 2013; Thang et al., 2016; Yan

2
H. Ma and J. Yang Ocean Engineering 198 (2020) 106963

et al., 2016) can be used to achieve the composite action in Table 1


concrete-filled double skin steel tubes, providing shear resistance and Main parameters of the NREL 5 MW wind turbine.
minimizing slip at the steel-concrete interface. A transition piece is set in Item Value
the hybrid monopile, and grouted to the CFDST. The tower is connected
Rated Power (MW) 5
to the transition piece by means of bolts on the top flange. The details of Number of blades 3
the CFDST are shown in Fig. 2. In this paper, we focused on a feasibility Rotor diameter (m) 126
study using CFDSF as part of a monopile foundation for offshore wind Rated wind speed (m/s) 11.4
turbines. The mechanical shear connector and transition piece are not Cut-in, Rated rotor speed (rpm) 6.9, 12.1
Rotor-nacelle assembly mass (kg) 350,000
modeled in detail, and the tower is connected directly to the outer tube Tower base diameter (m) 6
of the CFDST. The length of the conical steel tube Ltp is set to 7 m (Gentils Tower base thickness (m) 0.027
et al., 2017). The outer diameter of the outer steel tube (Douter ) is smaller Tower top diameter (m) 3.87
than the diameter of the pile, and is related to the wall thickness (touter ) Tower top thickness (m) 0.019
Coordinate location of RNA (x,y,z) (m) (0.41, 0.00, 1.97)
as (DNV.GL, 2016):
Moment of inertia of RNA (x, y, z) (kg.m2) (4.37, 2.35, 2.54) � 107
Douter Blade mass (kg) 17,740
touter ¼ 6:35 þ ðmmÞ (1)
100

where Douter is the outer diameter of the outer steel tube of the CFDST. 3. Reference model: NREL 5 MW turbine supported on OC3
The inner steel tube is like a concrete formwork with a constant wall monopile
thickness (tinner ¼0.02 m), and the outer diameter of the inner steel tube
is calculated as follows: We use the NREL 5 MW wind turbine (Jonkman et al., 2009) as a
reference model. This wind turbine has been used as a reference by
Dinner ¼ Douter 2touter 2tc (2) research teams around the world to quantify the benefits of advanced
offshore wind energy technologies (Gentils et al., 2017; Morato � et al.,
where Dinner is the outer diameter of inner steel tube of the CFDST; tc is 2017). Its main characteristics are described in Table 1. The NREL 5 MW
the concrete thickness of CFDST. is considered to be supported by a monopile foundation with a water
The total second moment of inertia for a CFDST structure can depth of 20 m, designed during the Offshore Code Comparison Collab­
approximately be estimated by the sum of the second moment of inertia oration (OC3) project for the International Energy Agency (IEA). The
for the outer steel tube, the concrete and the inner steel tube (Huang, height of the tower is 77.6 m, with a base diameter of 6 m and a top
2005). Though CFDST structures have been shown to perform well diameter of 3.87 m. The monopile has a length of 56 m and a constant
under a variety of loading conditions (Liang et al., 2019; Li et al., 2019; section, with an outer diameter of 6 m and a thickness of 60 mm. The
Wang et al., 2016), there is still limited knowledge regarding the 36-meter monopile was driven into layered sandy soils, and the
interaction between the concrete and the outer and inner steel tubes, remaining 20 m was from the seabed up to the sea level. The dimensions
especially for large diameter monopile structures. We used a simple of the pile, transition piece and tower and the soil profile are depicted in
method to approximately estimate the concrete thickness of the CFDST Fig. 3. The pile is connected to the tower through a transition piece
structures. The concrete thickness of the CFDST structure is approxi­ attached by grout at 10 m above the mean sea level (MSL). The tower is
mately calculated by an equivalent stiffness between the cross section of bolted together with a transition pie through the internal flange-bolt.
the CFDST and the pile of the reference monopile, which is an input for The material properties of the pile, tower and transition piece are
our FEM model. The performance of the CFDST structure under different based on S355 steel, whose density is increased by 8% to account for
loading conditions is then examined by the comprehensive FEM analysis secondary steel appurtenances, coatings and welds. In this study, a
not by the simplified method. commercial grout material Ducorit D4 is selected; it is commonly used in
Es Ipile Es Iouter Es Iinner grouted connections on offshore wind turbines. The mechanical prop­
Ic ¼ (3) erties of structural steel and grout are summarized in Table 2.
Ec
In order to compare the hybrid and conventional monopile founda­
� �14 tions, similar wind turbines, towers, site and met-ocean conditions used
Douter 2touter ðDouter 2touter Þ4 64Ic
π in the OC3 project are selected for the hybrid monopile study. The dis­
(4) tance between the mudline and the tower bottom Ltop mp is 30 m, and the
tc ¼
2
water depth Hw is 20 m. The diameter ratio λ is defined as the ratio
where Es is the Young’s modulus of steel; Ipile is the moment of inertia of between the outer tube of the CFDST and the outer diameter of the
the pile; Iouter is the moment of inertia of the outer steel tube of the conventional monopile, which is shown as follows:
CFDST; Iinner is the moment of inertia of the outer steel tube of the
CFDST; Ec is the Young’s modulus of concrete; Ic is the moment of inertia Douter
λ¼ (5)
of CFDST concrete. D
Like the conventional monopile (Hermans and Peeringa, 2016), the where D is the outer diameter of the conventional monopile;
hybrid monopile foundation can be fabricated onshore and transported It is assumed that a kind of ultra-high performance concrete (UHPC)
to designated location using floating method, after which it will be is used to fill into the gap between the outer and inner tubes of the
upended and vibro-driving into the seabed to the required depth (Mar­ hybrid monopile. According to some specifications (UHPC, 2017;
yruth, 2014). From the viewpoint of fabrication cost, filling the void Jammes et al., 2013), the design parameters for UHPC are listed in
between the tubes with concrete will lead to thinner shells and smaller Table 2. The thickness of the concrete ring, given in Table 3, is calcu­
diameters of tubes and therefore less steel fabrication costs. Although lated by Eq. (4).
use of concrete will add extra fabrication cost compared with the
traditional pure steel monopile, the tubes themselves act as the concrete 4. Load determination
formwork and thereby the cost for concrete fabrication is minimal.
4.1. Site-specific met-ocean conditions

Offshore wind turbine foundations are designed based on the met-

3
H. Ma and J. Yang Ocean Engineering 198 (2020) 106963

RNA
+87.6 m
+83 m ttower 19.4mm
ttower 20.3mm
+75 m
ttower 21.3mm
+65 m
ttower 22.3mm
+55 m Tower
ttower 23.4mm
+45 m
ttower 24.5mm
+35 m 6m× 30mm
10 m
ttower 24.5mm
+25 m
7.5 m
ttower 25.4mm
+15 m
+10 m ttower 27.0mm

0m MSL Transition
piece

0m
-20 m Mudline
Loose sand 5m

Medium sand 9m
Pile driven
6.31m× 30mm
in sand

Dense sand
22 m
-56 m -7 m
12.5mm concrete
6m× 60mm

(a) (b)
Fig. 3. Geometry: (a) support structure of NREL 5 MW; (b) details of the transition piece.

Table 2 Table 3
Material properties of steel and grout used in analysis. Parameters of the CFDST monopile used in analysis.
Properties Steel Grout UPHC λ Outer tube (mm) Concrete ring (mm) Inner tube (mm)

Young’s modulus (GPa) 210 70 50 Douter touter tc Dinner tinner


Poisson’s ratio 0.38 0.19 0.2
Density (kg/m3) 8500 2740 2500 0.85 5100 57.35 110 4765.3 20
Compressive strength (MPa) – 200 150 0.80 4800 54.35 250 4191.3 20
Tensile strength (MPa) – 10 10 0.75 4500 51.35 510 3377.3 20
Yield strength (MPa) 355 – – 0.70 4200 48.35 1000 2103.3 20

ocean conditions which highly depend on the location of the project. The 4.2. Permanent loads
site considered is located at the Dutch sector of the North Sea, 8 km away
from the shore of IJmuiden City and is referred to as the NL-1 location in Permanent load refers to the weight of the complete structure, which
Gentils et al. (2017). The general characteristics of the site met-ocean includes the weights of the tower, monopile, Rotor-Nacelle Assembly
conditions are summarized in Table 4. (RNA) and blades.

4
H. Ma and J. Yang Ocean Engineering 198 (2020) 106963

Table 4 turbines, the normal wind speed profile is given by the power law (ABS,
General information of Site NL-1 (Gentils et al., 2017). 2010):
Site NL-1 characteristics � �α
z
Water depth (m) 20
Vz ¼ Vhub (8)
zhub
Reference wind speed Vref (m/s) 50
50-year extreme 3 s gust wind speed at hub height Vg50 (m/s) 60
where zhub is the height of the hub; α is the power law exponent which is
Annual average wind speed Vave (m/s) 10
assumed to be 0.2.
50-year significate wave height Hs50 (m) 6.9
50-year pear spectral period Ts50 (s) 7.7
4.4. Wave and current load
50-year extreme current speed Vc50 (m/s) 0.8

Wave loads on slender structural members, such as a cylinder sub­


4.3. Aerodynamic wind loads merged in water, can be predicted by Morison’s equation (DNV GL,
2016; Nie and Liu, 2002).
The wind load acting on the turbine rotors is estimated to be (Arany Z ηðtÞ
D2
Z ηðtÞ
D
et al., 2017): Fwave ¼ FM þ FD ¼ Cm ρπ €xdz þ CD ρ xj
_ xjdz
_ (9)
dw 4 dw 2
Fvh ¼ 0:5πρa R2T V 2hub CT (6)
where FM is the inertia force in N; FD is the drag force in N; dw is the
where Fvh is the wind load acting on the hub in N, RT is the radius of the water depth in m; CM is the mass coefficient (2 for a smooth tubular
rotor in m, Vhub is the wind speed at the hub height in m/s, ρa is the air section); CD is the drag coefficient (1.2) for a smooth tubular section; ρ is
density with the value of 1.23 kg/m3, CT is the thrust coefficient which is the mass density of sea water (1030 kg/m3); D is the outer diameter of
a function of the tip speed ratio and is approximately assumed as 0.5 for the monopile foundation in m; x_ and x € are the wave-induced velocity
Vhub ¼ 11:4 m=s in the present study. and the acceleration of water, respectively in the horizontal direction,
The wind load acting on the turbine tower depends on the wind and ηðtÞ is the surface wave profile. The surface wave profile according
velocity along the tower. The tower is divided into different segments to linear wave theory is given by (Nie and Liu, 2002):
and the wind load is treated as a concentrated load in each segment. The ηðtÞ ¼ 0:5hw cosðkx ww tÞ (10)
wind load is calculated as follows (ABS, 2010):
hw πcoshðkðZ2 þ dw ÞÞ
F Ztower ¼ 0:5ρa Cs Aztower V 2z (7) x_ ¼ cosðkx ww tÞ (11)
Tw sinhðkdw Þ
where FZtower is the wind load acting on the tower of height z in N; Aztower is
2hw π2 coshðkðz2 þ dw ÞÞ
the wind pressure area on the tower of height z in m2; Cs is shape co­ €x ¼ sinðkx ww tÞ (12)
T 2w sinhðkdw Þ
efficient which equals 0.5 for the tubular steel tower; z is the height
above the sea water level. The wind profile, Vz , denotes the average
where hw is the wave height in m; k is the wave number in m 1, ww is the
wind speed as a function of the height z; in the case of standard wind
wave frequency in rad/s, Tw is the wave period in s; z2 is the depth below

Outer tube
Wind turbine

Infill concrete
Tower
Conventional
monopile

Inner tube

Finite element
zone
36 m

Infinite boundary element zone


36 m

Soil in
Soil in
pile
pile

30 m 102 m 30 m

(a) (b) (c)


Fig. 4. 3D Finite element model constructed: (a) model as a whole; (b) new monopile; (c) conventional monopile.

5
H. Ma and J. Yang Ocean Engineering 198 (2020) 106963

Table 5 5.2. Material model and properties


Properties for different soil layers used in analysis.
Type of Effective Young’s Angle of Cohesion Friction The soil is simply represented by a Mohr-Coulomb constitutive
sand unit weight Modulus friction (kPa) coefficient model with an elastic-perfectly plastic behavior. The elastic-plastic
(kN.m 3) (MPa) (deg.) behavior is mainly defined by cohesion, internal friction angle, dila­
Loose 10 30 33 2 0.40 tion angle, modulus of elasticity and Poisson’s ratio as shown in Table 5.
Medium 10 50 35 2 0.43 The piles and tower are assumed to be made of steel material having the
Dense 10 80 38.5 2 0.48 typical properties shown in Table 2. Yielding of the steel is not consid­
ered in this study. The Concrete Damaged Plasticity model available in
the sea surface in m. ABAQUS is used to model the infilled concrete. This model is a contin­
Current structural design models usually involve a simple current uum, plasticity-based, damage model for concrete. It is assumed that the
profile over depth, using the known current velocity at the water surface two main failure mechanisms are tensile cracking and compressive
as an input parameter. A commonly used method; the power law profile, crushing of the concrete material (ABAQUS, 2013). Four parameters are
is adopted here (ABS, 2010). The horizontal current load is estimated as: required to fully describe the yield surface and the flow rule in the
three-dimensional stress space, including the dilation angle (ψ), the
D
Fcurrent ¼ CD ρ U 2current (13) ratio of compressive strength under biaxial loading to uniaxial
2 compressive strength (fb0 =fb Þ, plastic flow potential eccentricity (ε), and
the ratio of the second stress invariant on the tensile meridian to that on
where Fcurrent is the horizontal current drag force per unit length in N;
the compress meridian (Kc ). The values of these parameters are set as
Ucurrent is the local current velocity in m/s.
55� , 1.1, 0.1 and 0.67 (Krahl et al., 2018; Shafieifar et al., 2017). The
This study did not consider other sources of dynamic loads on
concrete between the outer and inner tubes is considered to be confined
monopile foundations due to break waves, ice and earthquakes.
and an equivalent concrete model is adopted (Ye et al., 2017). The
confined concrete is then introduced into ABAQUS. A uniaxial
4.5. Design load combinations compressive stress (σÞ versus strain (ε) relationship is given as follows:
8
< 2χ χ 2 ðχ � 1Þ
Two sets of load combinations are used to assess the feasibility of the
y¼ χ (14)
new monopile foundation. One corresponds to the SLS, one corresponds : ðχ > 1Þ
to the SLS, which includes a rated wind speed of 11.4 m/s under a β0 ðχ 1Þ2 þ χ
combined effect of wave and current loads whose return period is 50
years with a load factor of 1.0 for all load categories. The other corre­ where χ ¼ ε=ε0 ; y ¼ σ=f ’c ; f ’c is the cylindrical compressive strength of
sponds to the ULS, and under characteristic load effects similar to the infill concrete in N/mm2 and is equal to 0:85fcu (fcu is the cube strength
SLS, the load factors are equal to 1.0 and 1.35 for permanent and of infill concrete); ε0 and β0 are determined as
environmental loads, respectively (DNV.GL, 2016). �
ε0 ¼ 1300 þ 12:5f ’c þ 800ξ0:2 � 10 6 (15)

5. Numerical modelling and verification �½0:25þðξ 0:5Þ7 � �0:5


β0 ¼ 1:18 � 10 5
f ’c � 0:12 (16)
In order to comprehensively analyze the wind turbine-pile-soil sys­
where ξ is the confinement factor that can be calculated as:
tem, a three-dimensional (3D) finite element model (FEM) is constructed
using the software ABAQUS, as shown in Fig. 4. The key geometric pa­ As fys
ξ¼ (17)
rameters are summarized in Figs. 1 and 2 and Tables 1 and 3 The FEM Ac fck
model consists of seven main components, namely turbine, tower, steel
pile, concrete, the soil around the pile, the soil inside the pile (Ma et al., where As and Ac are the cross-section areas of the outer steel tube and
2017), and infinite boundary elements (Stro €mblad, 2014). The wind the infill concrete; fys is the yield stress of the steel; fck is the charac­
turbine load is applied to the finite element model at tower top. The teristic strength of the infill concrete, equivalent to 0:67fcu .
tower is divided into nine segments and the wind load acts as a The infilled concrete between the outer and inner tubes of CFDST
concentrated load at each segment. The wave load and current load per under wave and wind loads may be subjected to tension. The simplified
unit length are calculated using Eq. (9) and Eq. (13), respectively. They stress-strain relationship of infill concrete under tension is assumed as
are assumed to act as uniform load on the surface of the monopile follows. The tensile stress increases linearly with the increase in tensile
foundation in the FE model. Two cases of monopile are investigated for strain; after concrete cracking it decreases linearly with further increase
comparison purposes: one is the conventional monopile and the other in strain and approaches zero. The ultimate tensile strain of infill con­
case is the hybrid monopile. crete is specified as 10 times the cracking strain, and the cracking strain
is set as 0.0001 (Mirzazadeh and Green, 2017).

5.1. Geometric configuration


5.3. Interaction properties
The size of the soil domain in the FEM is 162 � 162 � 72 m [40, 46].
The lower boundary is fixed against movements in all directions, and the The soil-pile and concrete-pile interactions are modeled using a
30 m thick layer of infinite soil creates vertical boundaries that do not surface-to-surface contact formulation in ABAQUS/Standard. In this
reflect shear waves in the soil medium. The continuum element type approach, the master surface is defined as a surface belonging to the
(C3D8R) is used to model the soil. The outer layer of the soil is modeled material that is relatively stiff or has finer mesh geometry, and the slave
using a single layer of solid infinite element (CIN3D8). The tower and surface corresponds to the softer material or material with a coarser
the piles are modeled using shell elements (S4R). The concrete is mesh (Johnson et al., 2001). The pile surface is defined as the master
modeled using the continuum element type (C3D8R). The nacelle is surface, and the soil and concrete surfaces in contact with the pile are
modeled as a lumped mass at the top node of the tower with a rotational defined as the slave surface. In the normal direction, the interface con­
inertia as specified in Table 1, with contribution from the mass of the tact is assumed to be a “hard” contact, and no separation was allowed.
rotor, nacelle and blades. When the surfaces are in contact, any contact pressure can be

6
H. Ma and J. Yang Ocean Engineering 198 (2020) 106963

Table 6 experiments. In this study, the friction coefficient (μ) between the pile
Computed and measured natural frequencies. and the concrete is selected as 0.3 (Hassanein et al., 2018).
FEM Measurements (Shirzadeh et al., 2013)

0.3657 Hz Over speed stop Ambient excitation 5.4. Model validation


0.3529 Hz 0.3565 Hz

For the reference wind turbine model compatible with the conven­
transmitted between them. The surfaces are separated if the contact tional monopile foundation, our FEM is validated by comparing the
pressure is reduced to zero. In the tangential direction, the classical computed results with the field measurements at natural frequencies
isotropic Coulomb friction model is used to simulate the shear-resistance (Shirzadeh et al., 2013) and with full-scale lateral loading test data
interaction (Johnson et al., 2001; Stro €mblad, 2014; Sheng et al., 2005). (Hokmabadi et al., 2012).
The friction coefficient (μ) for the interface between the pile and the soil Shirzadeh et al. (2013) reported the field measurements of the first
� � fore-aft mode frequency for a 3 MW offshore wind turbine supported by
is simply taken as μ ¼ tan 23 φ , where φ is the friction angle of the soil. a monopile in the Belgian North Sea. The overspeed stop test and
For the interface between the pile and the concrete, values of 0.2–0.3 are ambient excitation were used to estimate the first for-aft mode fre­
typically used as they provide acceptable results compared to quency. Based on the available soil, monopile, tower and turbine data
from Shirzadeh et al. (2013), a 3D FEM is developed using ABAQUS. The

100
Test results Hokmabadi et al., 2012

Finite element analysis results


Pile head lateral load (kN)

80

60
Loading ±0
point
2m
40
-15 m
Seabed

20
-29 m
Pile diameter: 1.778m
0
0 50 100 150 200 250 300 350

Pile head horizontal displacement (mm)


Fig. 5. Comparison of field test results with finite element analysis results.

30 =1.00
Tower bottom
=0.85
26
Monopile height above the mudline (m)

=0.80
22 =0.75
=0.70
18 MSL

14

10

2 Mudline

-2
400 900 1400 1900 2400 2900
Horizontal force on monopile (kN)

Fig. 6. Computed horizontal forces on monopile under the SLS.

7
H. Ma and J. Yang Ocean Engineering 198 (2020) 106963

30 =1.00
Tower bottom
=0.85
26

Monopile height above the mudline (m)


=0.80
22 =0.75
=0.70
18 MSL

14

10

2 Mudline

-2
38 43 48 53 58 63 68 73 78 83 88
Bending moment on the monopile (MN·m)

Fig. 7. Computed bending moments on monopile under the SLS.

comparative results of the finite element analysis with the reported re­
Table 7
sults are shown in Table 6. It can be seen that the result of the finite
Natural frequencies.
element analysis is in good agreement with that of the reported results.
Hokmabadi et al. (2012) reported full-scale load tests to investigate Natural frequency λ ¼ 1:0 λ ¼ 0:85 λ ¼ 0:80 λ ¼ 0:75 λ ¼ 0:70
the behavior of offshore monopiles in marine sandy soils. The length and f1 0.2332 0.2256 0.2231 0.2218 0.2167
diameter of the monopile were approximately 34 m and 1.78 m, f2 1.4471 1.4485 1.4446 1.4439 1.3949
respectively. A 3D FEM similar to that developed in the present study is
established using ABAQUS for this monopile-supported wind turbine,
with soil and pile properties taken from Hokmabadi et al. (2012). The monopile.
result of pile head load versus pile head deflection obtained from the
finite element analysis is presented in Fig. 5, showing a reasonably good
6.2. Natural frequency
agreement.
The natural frequency of an offshore wind turbine is a major design
6. Results and discussion driver for the monopile support structure, since it defines the dynamic
behavior of the structure (Velarde, 2016). The overall natural frequency
This section presents the numerical results for hybrid and conven­ must not coincide with the excitation frequencies due to wind and
tional monopile foundations. The results include the horizontal loads, waves. In design practice, the first natural frequency (f1 Þ lies between
the natural frequency, the lateral displacement and the rotation of the the frequencies 1P (rotor speed frequency) and 3P (blade passing fre­
monopile under the SLS, the maximum von Mises stress of the monopile quency), which is called the “soft-stiff” design (O’Kelly and Arshad,
under the ULS, the optimum of the embedded length and materials cost 2016). Based on the rotor speeds as shown in Table 1, the first natural
of the hybrid monopile. The outer diameter ratio λ is a key design frequency (f1 Þ is required to be in a range between 0.202 Hz (1P) and
parameter that plays an important role in the performance of the hybrid 0.345 Hz (3P) to avoid resonance. If the calculation uncertainties �5%
monopile. To investigate its influence, λ was chosen to be 1.0, 0.85, 0.8, are included (DNV.GL, 2016), the above range goes from 0.212 Hz to
0.75 and 0.7. 0.328 Hz. The linear perturbation method is used to determine f1 and f2 .
As shown in Table 7, all f1 fall within the allowable “soft-stiff” region
(1P–3P), and all f2 fall well above the maximum limit of the 3P opera­
6.1. Horizontal load and bending moment along monopile
tional (0.636 Hz). Consequently, the hybrid monopile can be considered
safe from resonance or resonance-related effects. In Table 7, as λ is
Using the load calculation methods described above, the wind, wave
reduced from 1.0 to 0.7, f1 is reduced from 0.233 Hz to 0.217 Hz, and f2
and current loads are determined based on the met-ocean conditions at
is reduced from 1.447 Hz to 1.395 Hz, by about 6.9% and 3.6%,
site NL-1. The total horizontal load and the bending moment along the
respectively. These changes are due to the fact that the stiffness remains
monopile with different λ values under the SLS are plotted in Fig. 6 and
constant as λ decreases from 1.0 to 0.7 but the mass of the monopile
Fig. 7, where the horizontal axis represent the cross-sectional load and
increases due to the use of CFDST.
bending moment at different levels of the monopile, respectively. It can
be seen that the change in λ directly influences the horizontal load and
bending moment profiles along the monopile, and they all have a 6.3. Pile stress distribution under the ULS
decreasing tendency as λ decreases. At mudline, as the λ decreases from
1.0 to 0.7, the horizontal load and the bending moment are reduced by The stress distribution of the outer steel tube is shown in Fig. 8. The
almost 41.8 and 19.5%, respectively. It is indicated that the reduction of maximum von Mises stresses of the outer steel tube of the hybrid
the pile diameter can effectively reduce the total external load for the monopile are 142.7, 127.8, 124.3 and 105.5 MPa for λ ¼ 0.7, 0.75, 0.8

8
H. Ma and J. Yang Ocean Engineering 198 (2020) 106963

Fig. 8. Von Mises stress distribution of the outer steel tube: (a) λ ¼ 0:7; (b) λ ¼ 0:75; (c) λ ¼ 0:8; (d) λ ¼ 0:85; (e) λ ¼ 1:0

Fig. 9. Tensile and compressive stress distribution of the concrete between outer and inner steel tube: (a) λ ¼ 0:7; (b) λ ¼ 0:75; (c) λ ¼ 0:8; (d) λ ¼ 0:85

and 0.85, respectively, which are located at the cross-section above the monopile are depicted in Fig. 10. The maximum cross-section von Mises
top surface of the concrete. These values are below the permissible value stresses are 15.8, 27.2, 35.6 and 49.4 MPa for λ ¼ 0.7, 0.75, 0.8 and
of 355 MPa. The maximum von Mises stress of the pile for λ ¼ 1:0 is 82.3 0.85, respectively, which are far less than the permissible value of 355
MPa, which is located near the mudline. Fig. 9 shows the tensile and MPa. Therefore, the CFDST structure is safe under the ULS.
compressive stress distribution of the concrete. The maximum
compressive stresses are far below the compressive strength of the UHPC
(150 MPa). The maximum tensile stresses are 7.7, 7.9, 9.4 and 9.5 MPa 6.4. Global buckling check under the ULS
for λ ¼ 0.7, 0.75, 0.8 and 0.85, respectively, which are less than the
tensile strength of UHPC (10 MPa), and therefore the concrete will not For the ultimate strength of the monopile, the buckling stability
crack. The stress distributions of the inner steel tube of the hybrid should be checked using a rational and justifiable engineering approach
(DNV.GL, 2016). Buckling refers to sudden collapse of the structure

9
H. Ma and J. Yang Ocean Engineering 198 (2020) 106963

Fig. 10. Von Mises stress distribution of the inner steel tube: (a) λ ¼ 0:7; (b) 0.75; (c) λ ¼ 0:8; (d) λ ¼ 0:85

mudline are reduced from 19.3 to 12.8 mm (about 33.7%) and from
0.104 to 0.072� (about 30.8%), respectively. The critical depth as a
Tower bottom =1.00
28 transition point (from positive to negative displacement) is also reduced
Monopile height above mudline (m)

=0.85 from 19 to 17 m. In addition, the lateral deflection and rotation at the


Variable section
23 =0.80 pile toe is also reduced from 1.52 to 1.07 mm (about 19.6%) and
=0.75
from 0.0047 to 0.0021� (about 55%). These are attributed to the
18 reduction of horizontal wave loads by reducing λ.
=0.70

13 6.6. Permanent accumulated rotation at mudline under the SLS

8 Variable section In most cases, the permanent deformation tolerance of the pile under
the SLS is design-driving for monopiles in sandy soils (Velarde, 2016;
3
Mudline Schmoor and Achmus, 2015; Senanayake, 2016). The deformation tol­
erances are usually given in the design basis and are often specified in
-2 terms of maximum tolerance rotations of the pile at the mudline in a
0 0.2 0.4 0.6 0.8 1 vertical plane. Typically, the tolerance for the total rotation is 0.5� and
Global buckling utilization ratio the installation tolerance is 0.25� , then the limit for permanent accu­
mulated rotation becomes 0.25� for SLS loads throughout the design life
Fig. 11. Global buckling check for monopiles under the ULS. (DNV.GL, 2016).
According to Schmoor and Achmus (2015), the accumulated rotation
either globally or locally. For the hybrid monopile, the mechanical shear of the pile at mudline can be estimated with the empirical exponential
connectors should be used to achieve the composite action in approach as follows:
concrete-filled double skin steel tubes, and can be beneficial for local
θN ¼ θ1 ζSDM (18)
buckling prevention (Eom et al., 2019); the local buckling of both steel
tubes can be effectively constrained due to the presence of infilled UHPC where θN is the pile accumulated rotation by N cycles of lateral load; θ1
concrete (Wang et al., 2019). Therefore, the local buckling will be is the rotation obtained in the first loading cycle, which can be calcu­
complicated and delayed, such detailed analysis does not included in lated in a static analysis under the SLS loads; ζSDM is derived from the
this paper. The global buckling is checked according to Standard for Stiffness Degradation Method (SDM) presented in Kuo (2008) and
design of steel structures (GB 50017–2017) and Technical specification Achmus et al. (2009), which is
for concrete-filled double skin steel tubular structures (CCES, 2018). As !
shown in Fig. 11, the global buckling utilization ratios of the conven­ H ðhþLpile Þ

tional and novel monopiles are far less than 1.0. This indicates the ðA 1:208Þln
γ ’ DL3
þB 0:588

(19)
pile

present hybrid monopile is not likely to suffer from global buckling ζSDM ¼ e �1
failure.
where Lpile is the pile embedded length in m; H is the horizontal force in
kN; h is the moment arm in m (31.6 m, 36.5 m, 38.6 m, 40.9 m and 43.6
6.5. Pile deformation under the SLS m for λ ¼ 1.0, 0.85, 0.8, 0.75 and 0.7, respectively); γ ’ � 10 kN=m3 is the
soil effective unit weight; A and B are regression parameters and take
Figs. 12 and 13 show the comparison of lateral deflection and rota­ values of 1.361 and 1.331 for the SLS (number of cycles N ¼ 100) ac­
tion along the pile for different λ values under the SLS, respectively. It is cording to Schmoor and Achmus (2015) and Barari et al. (2017).
obvious that its lateral deformation decreases when λ decreases. As λ The pile permanent accumulated rotation θN at mudline under the
decreases from 1.0 to 0.7, the lateral deflection and the rotation in the SLS loads is presented in Table 8. It can be clearly seen that the overall

10
H. Ma and J. Yang Ocean Engineering 198 (2020) 106963

Lateral displacement (mm)


-2 0 2 4 6 8 10 12 14 16 18 20
0

-4

Pile penetration depth (m) -8

-12

-16

-20

-24
Undeformed pile
-28

=1.00 =0.85 =0.80 =0.75 =0.70


-32

-36
Fig. 12. Distribution of pile deflection under the SLS.

Pile rotation ( º)
0 0.02 0.04 0.06 0.08 0.1
0

-4

-8
Pile penetration depth (m)

-12

-16

-20

-24

-28

=1.00 =0.85 =0.80 =0.75 =0.70


-32

-36
Fig. 13. Distribution of pile rotation under the SLS.

trend of θN decreases with λ. As λ decreases from 1.0 to 0.7, ​ θN reduces


Table 8
from 0.145� to 0.095� (about 34.5%). Therefore, it is possible to opti­
Pile rotation at mudline under the SLS.
mize the geometry of the monopile while maintaining a similar accu­
Pile rotation λ ¼ 1:0 λ ¼ 0:85 λ ¼ 0:80 λ ¼ 0:75 λ ¼ 0:70 mulated rotation at mudline with that of the conventional monopile
θ1 ( )

0.104 0.090 0.085 0.078 0.072 (λ¼1.0). This can lead to a more economical design.
ζSDM 1.395 1.356 1.343 1.330 1.317
θN ( )

0.145 0.122 0.114 0.104 0.095 7. Embedded length optimization for hybrid monopile

To compare the embedded length between the hybrid monopile and


the conventional monopile (OC3), two major design drivers are selected
for monopile foundation: one is the pile permanent accumulated

11
H. Ma and J. Yang Ocean Engineering 198 (2020) 106963

Table 9 thickness of the embedded pile of the hybrid monopile remain constant.
Monotonic rotation ( ) at the mudline of the monopile with different.β Several embedded pile length ratio β ¼ 0:9; ​ 0:8; ​ 0:7 ​ and ​ 0:6 are

β λ ¼ λ ¼ 0:80 λ ¼ 0:75 λ ¼ 0:70 selected.


0:85 .
β ¼ Lpile LOC3
pile (20)
0.6 0.1095 0.1074 0.0995 0.0911
0.7 0.0938 0.0924 0.0866 0.0786
0.8 0.0930 0.0883 0.0817 0.0744 where LOC3
pile ¼ 36 m is the pile embedded length for the reference model.
0.9 0.0912 0.0861 0.0798 0.0731 For different cases of β, a total of 20 (4 � 5) finite element models are
1.0 0.0900 0.0851 0.0783 0.0722
established for calculating the pile rotation in the first loading cycle (θ1 ),
as shown in Table 9. When β is reduced from 1.0 to 0.6, the value of θ1 is
rotation (θN ) at mudline under the SLS; the other is the natural frequency increased by nearly 22, 26, 27 and 26.2% for λ values of 0.85, 0.8, 0.75
(f1 ; f2 ) of the overall support structure. During the comparison process, and 0.7, respectively. The pile accumulated rotations with different β are
θN remains similar between the hybrid and the conventional monopile; calculated with Eqs. (18) and (19). As shown in Fig. 14, θN at mudline
f1 lies in the desired range, I.e. between 0.212 Hz and 0.328 Hz; ​ f2 falls increases with decreasing β. Specifically, when λ ¼ 0:85, 0.80, 0.75 and
above the maximum limit of the 3P operational; the diameter and wall 0.70, β decreases from 1.0 to 0.6, and θN increases by 48, 54, 55.7 and

0.19
Accumulated permanent rotation of pile at mudline ( º)

=0.85
0.18
=0.80
0.17
=0.75
0.16 =0.60
0.15 = 1.0, = 1.0

0.14

0.13

0.12

0.11

0.1
0.61 0.655 0.69 0.75
0.09
0.6 0.7 0.8 0.9 1
3 3
= / , = 36
Fig. 14. Pile accumulated rotation at mudline under the SLS.

3 = 0.328 =0.85
0.245
=0.80
= 0.233
The first natural frequency (Hz)

=0.75
= 1.0, = 1.0 =0.70
0.235

0.225

0.215

1 = 0.212
0.205
0.6 0.7 0.8 0.9 1
OC 3 OC 3
L pile / L pile ,L pile 36m
Fig. 15. Values of the first natural frequency under different β and λ values.

12
H. Ma and J. Yang Ocean Engineering 198 (2020) 106963

2
= 1.447 = 1.0, = 1.0
1.45

The second natural frequency (Hz)


1.43

1.41

1.39
=0.85
=0.80
1.37 =0.75
The upper limit value of 3P: 0.636 Hz
=0.70
1.35
0.6 0.7 0.8 0.9 1
OC 3 OC 3
L pile / L pile ,L
pile 36m
Fig. 16. Values of the second natural frequency under different β and λ values.

55%, respectively. This means that the embedded pile length has a Fig. 15 shows the first natural frequency of the NREL 5 MW offshore
significant impact on the pile accumulated rotation. The conventional wind turbine supported by the hybrid monopile foundation. The
monopile is taken with λ ¼ 1:0 and β ¼ 1:0, and in this case the accu­ reduction in β from 1.0 to 0.6 has little effect on the first natural fre­
mulated permanent rotation at mudline is 0.145� . For the same accu­ quency under different λ conditions. The first natural frequency is less
mulated permanent rotation at mudline, values of λ and β are varied for than that of the comparative model (OC3, λ ¼ 1:0, β ¼ 1:0Þ, on
hybrid monopile and an optimum condition is achieved, as shown in decreasing by 3.7, 4.4, 5.6 and 7.7% for λ ¼ 0:85, 0.80, 0.75 and 0.70,
Fig. 14. The results show that the equal accumulated permanent rotation respectively, which are within the desired range of 0.212 Hz and 0.328
is achieved for λ ¼ 0:85 at 75% of the embedded pile length. Similarly, Hz. As shown in Fig. 16, the second natural frequencies are all above the
for λ ¼ 0.80, 0.75, and 0.6, the same accumulated permanent rotation is above the maximum limit of the 3P operational (0.636 Hz), and are less
achieved at 69, 65.5, and 61% of the embedded pile length, respectively. than that of the comparative model.

1.8

1.6

1.4 Total weight ratio of the


hybrid and OC3 monopile
Weight ratio

1.2

0.8

Steel weight ratio of the


0.6 hybrid and OC3 monopile

0.4 Concrete weight to


steel weight of OC3
0.2

0
0.7 0.75 0.8 0.85 0.9 0.95 1

Fig. 17. Weight ratio of the hybrid and OC3 monopile.

13
H. Ma and J. Yang Ocean Engineering 198 (2020) 106963

750

700 Total material cost of


OC3 monopile

)
Material cost of monopile (×104 650

600

550

¥ 1500/ton
500
Total material cost of ¥ 2000/ton
the hybrid monopiles ¥ 2500/ton
450 ¥ 3000/ton
¥ 3500/ton

400
0.7 0.75 0.8 0.85 0.9 0.95 1

Fig. 18. Total material cost of the hybrid and OC3 monopile.

8. Cost comparison for hybrid monopile and OC3 monopile result, the total material cost (Ctot ) of the hybrid monopiles is less than
that of the convention monopile in most situations. Fig. 18 shows that
In general, costs for monopile foundation include those for materials, Ctot increases with the increase of the unit price of UHPC (Cc ) and the
fabrication, installation, transport and maintenance. Among them, the magnitude of increase becomes larger as λ decreases. For example in the
material weights and costs are relatively straightforward to estimate, but case of Cc ¼ ¥3500=ton, the total material cost for the hybrid monopile
costs related to installation, manufacturing and transport (Muskulus and is reduced approximately 13.6%, 18.9%, 20.7% and 18.3% for λ ¼ 0:7;
Schafhirt, 2014) are complicated and depend on many factors. In the ​ 0:75; ​ 0:8 ​ and ​ 0:85, respectively. This suggests that the hybrid type
present study we focus mainly on the comparison of costs of materials of monopile is a cost-effective alternative to the conventional type of
for the hybrid type and the conventional type of monopile. In doing this, monopile for offshore wind turbines.
a steel weight ratio ws is defined as the ratio between the weight of the
hybrid monopile (wnovelsteel ) and the weight of the conventional monopile
9. Conclusions
(wOC3
steel ):
This paper presents a hybrid type of monopile foundation for
W novel offshore wind turbines and investigates its feasibility by a comprehen­
ws ¼ steel
(21)
W OC3
steel sive numerical analysis. The NREL 5 MW wind turbine supported by a
conventional monopile is used as a reference model for comparison. The
For steel, a price of 10000 ¥/ton including the cost for raw material
natural frequency and the structural responses under the SLS and ULS
and labor (MIIC, 2018) is supposed. Ultra-high performance concrete is
conditions are computed and compared. The main results are summa­
much more expensive than normal concrete and the price depends on
rized as follows:
the formulation of products for various applications (Christopher et al.,
2017); the price also varies in different markets and regions around the
(a) Compared to the conventional monopile (OC3λ ¼ 1:0), the hor­
world (NPCA, 2015). According to Xu (2015), the unit price of UHPC is
izontal load at mudline for the hybrid monopile can be reduced
from ¥1000/ton to ¥1200/ton; the unit price of Nanodur® Compound
by 22.8, 29.5, 35.9 and 41.8%, respectively, at λ values of 0.85,
5941 available in the market is about ¥2800/ton to ¥3420/ton
0.8, 0.75 and 0.7, and the bending moment can be reduced by
(~€366/ton to €446/ton) (UHPC, 2017). Therefore, a range of values of
10.7, 13.7, 16.7 and 19.5%, respectively. This indicates that the
unit price for UHPC (¥1500/ton to ¥3500/ton) is used in the cost esti­
hybrid monopile can effectively reduce the total external load on
mation. The total material cost Ctot is calculated as:
the pile.
Ctot ¼ Cc wc þ Cs ws (22) (b) The first natural frequency of the system decreases by approxi­
mately 7% as λ decreases from 1.0 to 0.7 and the natural fre­
where Cc and wc are the unit price and weight of UHPC, respectively; Cs quencies at various λ values remain between the frequencies 1P
and ws are the unit price and weight of steel, respectively. (rotor speed frequency) and 3P (blade passing frequency). This is
Fig. 17 illustrates the effects of λ on the material weight of the hybrid due to the fact that the cross-sectional stiffness remains un­
monopile. Compared to the conventional monopile, the total weight of changed for the hybrid type and the conventional type of
the hybrid monopile increases nearly by 1.85 times while the steel monopile but the mass of the hybrid monopile increases.
weight ratio ws decreases by 55.7% as λ decreases from 1.0 to 0.7. As a

14
H. Ma and J. Yang Ocean Engineering 198 (2020) 106963

(c) Under the ultimate limit state, the maximum stresses in the steel conditions.
tube pile and in the filled concrete do not exceed the permissible
stresses. Author contributions
(d) Under the serviceability limit state, the permanent accumulated
rotation at mudline is reduced from 0.145 to 0.095� (about Hongwang Ma: Concept of the new hybrid monopile, Methodology,
34.5%) as λ decreases from 1.0 to 0.7. This is mainly due to the Finite element model and analysis, Data curation, Writing- Original
reduction of wave loads on the hybrid monopile. Considering that draft preparation. Jun Yang: Methodology, Writing- Reviewing and
the permanent accumulated rotation under the SLS is a major Editing.
design driver for monopile foundation, the embedded pile length
can possibly be optimized to achieve a more economical design.
(e) Using the proposed optimization criteria, the embedded length of Declaration of competing interest
the hybrid monopile can be reduced by 25, 31, 34.5 and 39% for λ
being 0.85, 0.8, 0.75 and 0.7, respectively. This indicates that the The authors declare that they have no known competing financial
steel weight and the relevant cost can be significantly reduced. interests or personal relationships that could have appeared to influence
The hybrid monopile therefore is a cost-effective alternative to the work reported in this paper.
the conventional monopile for offshore wind turbines.
Acknowledgments
Last but not least, it should be noted that a number of issues need to
be studied for practical use of the new type of monopile, such as the The authors wish to acknowledge the financial support provided by
detailed connection between the concrete and the steel tube, the fatigue the National Basic Research Program of China (973 Program, No.
performance of the concrete and the nonlinear wave theory used for 2014CB046200) and the Visiting Professorship awarded by Shanghai
prediction of wave loads on the hybrid monopile for extreme wave Jiao Tong University, China.

Notation

The following symbols are used in this paper


Ac cross-section area of the infill concrete
As cross-section area of the outer steel tube
Aztower wind pressure area on the tower of height z
CD drag coefficient
CM mass coefficient
Cs shape coefficient
CT thrust coefficient
CFDST concrete-filled double skin steel tubular structure
dw water depth
D​ outer diameter of the conventional monopile
Dinner inner diameter of the inner steel tube of the innovative monopile
Douter outer diameter of the outer steel tube of the innovative monopile
Ec Young’s modulus of concrete
Es Young’s modulus of steel
f1 First natural frequency
f2 Second natural frequency
f ’c compressive strength of infill concrete
fck characteristic strength of the infill concrete
fcu cube strength of the infill concrete
fys the yield stress of steel
Fcurrent horizontal current drag force per unit length
FD drag force
FM inertia force
FZtower wind load acting on the tower of height z
Fvh wind load acting on the hub
g acceleration of gravitation
hw wave height
Ic concrete section moment of inertia
Iinner inner steel tube section moment of inertia of the innovative monopile
Iouter outer steel tube section moment of inertia of the innovative monopile
Ipile pile section moment of inertia of conventional monopile
k wave number
Lpile pile embedded length
RT rotor radius
tc concrete thickness of CFDST
tinner wall thickness of the inner steel tube of the innovative monopile
touter wall thickness of the outer steel tube of the innovative monopile
Ucurrent local current velocity

15
H. Ma and J. Yang Ocean Engineering 198 (2020) 106963

UHPC Ultra-high performance concrete


Vhub wind speed at the hub height
Vz wind profile
ww wave frequency
x_ wave induced velocity of water
x
€ wave induced acceleration of water
z height above the sea water level
z2 depth below sea surface
α power law exponent
β ratio of pile embedded length between the new and conventional monopile
λ ratio of outer diameter between the new and conventional monopile
ρ mass density of the sea water
ρa air density
σ concrete stress
ε concrete strain
θN pile accumulated rotation at mudline
θ1 rotation obtained in the first loading cycle
ζSDM cyclic increase factor
γ’ soil effective unit weight
ξ confinement factor
ηðtÞ surface wave profile
MSL mean sea level
SLS serviceability limit state
ULS ultimate limit state

References Hassanein, M.F., Elchalakani, M., Karrech, A., Patel, V.I., Daher, E., 2018. Finite element
modelling of concrete-filled double skin short compression members with CHS outer
and SHS inner tubes. Mar. Struct. 61, 85–99.
ABAQUS. User’s Manual, Version 6.13. Karlsson & Sorenson, Inc., Pawtucket, Rhode
Hermans, K.W., Peeringa, J.M., 2016. Future XL Monopile Foundation Design for a 10
Island.
MW Wind Turbine in Deep Water. ECN-E-16-069.
ABS, 2010. Guide for Building and Classing Offshore Wind Turbine Installations.
Hokmabadi, A.S., Fakher, A., Fatahi, B., 2012. Full scale lateral behaviour of monopile in
American Bureau of Shipping, ABS Plaza, USA.
granular marine soils. Mar. Struct. 29, 198–210.
Achmus, M., Kuo, Y.S., Abdel-Rahman, K., 2009. Behavior of monopile foundation under
Huang, H., 2005. Behavior of Concrete Filled Double-Skin Steel Tubular Beam-Columns,
cyclic lateral load. Comput. Geotech. 36, 725–735.
PhD Thesis. Fuzhou University, China.
Alamo,
� G.M., Azn�arez, J.J., Padr�on, L.A., Martínez-Castro, A.E., Gallego, R., Maeso, O.,
Jammes, F.X., Cespedes, X., Resplendino, J., 2013. Design of Offshore Wind Turbines
2018. Dynamic soil-structure interaction in offshore wind turbines on monopiles in
with UHPC, Symposium on Ultra-high Performance Fibre-Reinforced Concrete.
layered seabed based on real data. Ocean Eng. 156, 14–24.
Marseille, France.
Andres, A.D., MacGillivray, A., Roberts, O., Guanche, R., Jeffrey, H., 2017. Beyond
Johnson, K., Karunasena, W., Sivakugan, N., Guazzo, A., 2001. Modeling pile-soil
LCOE: a study of ocean energy technology development and deployment
interaction using contact surfaces. Comput. Mech.-N. Front. New Millenn. 375–380.
attractiveness. Sustain. Energy Technol. Assess. 19, 1–16.
Jonkman, J., Butterfield, S., Musial, W., Scott, G., 2009. Definition of a 5-MW Reference
Arany, L., Bhattacharya, S., Macdonald, J., Hogan, J., 2017. Design of monopoles for
Wind Turbine for Offshore System Development. Technical Report NREL/TP-500-
offshore wind turbines in 10 steps. Soil Dyn. Earthq. Eng. 92, 126–152.
38060.
Barari, A., Bagheri, M., Rouainia, M., Ibsen, L.B., 2017. Deformation mechanisms for
Kallehave, D., Byrne, B.W., Thilsted, C.L., Mikkelsen, K.K., 2015. Optimization of
offshore monopile foundations accounting for cyclic mobility effects. Soil Dyn.
monopiles for offshore wind turbines. Philos. Trans. 373, 1–15.
Earthq. Eng. 97, 439–453.
Kim, H.G., Kim, B.J., 2018. Feasibility study of new hybrid piled concrete foundation for
Bocher, M., Mehmanparast, A., Braithwait, J., Shafiee, M., 2018. New shape function
offshore wind turbine. Appl. Ocean Res. 76, 11–12.
solutions for fracture mechanics analysis of offshore wind turbine monopile
Krahl, P.A., Carrazedo, R., Debs, M.K.E., 2018. Mechanical damage evolution in
foundations. Ocean Eng. 160, 264–275.
UHPFRC: experimental and numerical investigation. Eng. Struct. 170, 63–77.
CCES, 2018. Technical specification for concrete-filled double skin steel tubular
Kuo, Y.S., 2008. On the Behavior of Large Diameter Piles under Cyclic Lateral Load. Ph. D
structures. http://www.doc88.com/p-0009122357132.html. China.
Thesis. Leibniz University, Hannover, Germany.
Chen, S.M., Zhang, R., Jia, L.J., Wang, J.Y., Gu, P., 2018. Structural behavior of UHPC
Li, M.H., Zong, Z.H., Liu, L., Lou, F., 2018. Experimental and numerical study on damage
filled steel tube columns under axial loading. Thin-Walled Struct. 130, 550–563.
mechanism of CFDST bridge columns subjected to contact explosion. Eng. Struct.
Christopher, D.J., Mohamed, A.M., Keri, L.R., 2017. Cost and Ecological Feasibility of
159, 265–276.
Using UHPC in Highway Bridge, NDOT Research Report No. 224-14-803. Nevada
Li, W., Wang, T., Han, L.H., 2019. Seismic performance of concrete-filled double-skin
Department of Transportation.
steel tubes after exposure to fire: Experiments. J. Constr. Steel Res. 154, 209–223.
DNV⋅GL, 2016. DNVGL-ST-0437: Loads and Site Conditions for Wind Turbines. Det
Liang, W., Dong, J.F., Wang, Q.Y., 2019. Mechanical behaviour of concrete-filled double-
Norske Veritas, Oslo, Norway.
skin steel tube (CFDST) with stiffness under axial and eccentric loading. Thin-Walled
Eom, S.S., Vu, Q.V., Choi, J.H., Park, H.H., Kim, S.E., 2019. Flexural behavior of
Struct. 138, 215–230.
concrete-filled double skin steel tubes with a joint. J. Constr. Steel Res. 155,
Ma, H.W., Yang, J., Chen, L.Z., 2017. Numerical analysis of the long-term performance of
260–271.
offshore wind turbines supported by monopiles. Ocean Eng. 136, 94–105.
GB 500 17-2017, 2017. Code for Design of Steel Structure. China Architecture & Building
Maryruth, B.P., 2014. XXL Monopile & Vibro-Hammers for Cheaper Offshore Wind. IQPC
Press, China.
GmbH, Germany.
Gentils, T., Wang, L., Kolios, A., 2017. Integrated structural optimisation of offshore wind
MIIC, 2018. Analysis of the Price Trend of World Steel Market, 2018, 5. Metallurgical
turbine support structures based on finite element analysis and genetic algorithm.
industry information center, p. 55.
Appl. Energy 199, 187–204.
Mirzazadeh, M.M., Green, M.F., 2017. Non-linear finite element analysis of reinforced
Gjersøe, N.F., Pedersen, E.B., Kristiansen, B., Hansen, N.E.O., Ibsen, L.B., 2015. Weight
concrete beams with temperature differentials. Eng. Struct. 152 (1), 920–933.
optimisation of steel monopile foundations for offshore windfarms. In: Proceedings
Morat� o, A., Sriramula, S., Krishnan, N., Nichols, J., 2017. Ultimate loads and response
of the Twenty-Fifth International Ocean and Polar Engineering Conference Kona. Big
analysis of a monopile supported offshore wind turbine using fully coupled
Island, USA.
simulation. Renew. Energy 101, 126–143.
GWEC, 2017. Global wind report: annual market update 2017. Global Wind Energy
Murphy, G., Igoe, D., Doherty, P., Gavin, K., 2018. 3D FEM approach for laterally loaded
Council report.
monopile design. Comput. Geotech. 100, 76–83.
Han, L.H., 2016. Concrete Filled Steel Tubular Structures: Theory and Practice, third ed.
Muskulus, M., Schafhirt, S., 2014. Design optimization of wind turbine support
Science Press, Beijing.
structures- A review. J. Ocean Wind Energy 1 (1), 12–22.
Han, L.H., Li, Y.J., Liao, F.Y., 2011. Concrete-filled double skin steel tubular (CFDST)
Negro, V., L� opez-Guti�errez, J., Esteban, M.D., Alberdi, P., Imaz, M., Serraclara, J.M.,
columns subjected to long-term sustained loading. Thin-Walled Struct. 49,
2017. Monopiles in offshore wind: preliminary estimate of main dimensions. Ocean
1534–1543.
Eng. 133, 253–261.

16
H. Ma and J. Yang Ocean Engineering 198 (2020) 106963

Nie, W., Liu, Q.Y., 2002. Dynamic Analysis of Offshore Structures. Harbin Engineering Stansby, P.K., Devaney, L.C., Stallard, T.J., 2013. Breaking wave loads on monopiles for
University Press, Harbin, China. offshore wind turbines and estimation of extreme overturning moment. IET Renew.
NPCA, 2015. Guide to manufacturing architectural precast UHPC elements. https://pre Power Gener. 7 (5), 514–520.
cast.org/wp-content/uploads/2015/02/UHPC-White-Paper.pdf. Str€
omblad, N., 2014. Modeling of Soil and Structure Interaction Subsea. Master’s Thesis.
Ou, L., Xu, W., Yue, Q., Ma, C.L., Teng, X., Dong, Y.E., 2018. Offshore wind zoning in Chalmers University of Technology, Sweden.
China: method and experience. Ocean Coast Manag. 151, 99–108. Thang, V., Marshall, P., Brake, N.A., Adam, F., 2016. Studded bond enhancement for
O’Kelly, B.C., Arshad, M., 2016. Offshore Wind Turbine Foundations- Analysis and steel-concrete-steel sandwich shells. Ocean Eng. 124, 32–41.
Design. Offshore Wind Farms Technologies, Design and Operation, pp. 589–610. Uenaka, K., 2016. CFDST stub columns having outer circular and inner square sections
Rad, H., Turaj, A., Paul, L.C., Molenaar, D.P., 2014. Integrated multidisciplinary under compression. J. Constr. Steel Res. 120, 1–7.
constrained optimization of offshore support structures. J. Phys. Conf. Ser. 555 (1), UHPC, 2017. Ultra high performance concrete made of NANODUR® Compound 5941.
1–10. http://www.dyckerhoff.com/online/download.jsp?idDocument¼315&instance¼1.
REN21, 2017. Renewables 2017 Global Status Report. Paris, France. Velarde, J., 2016. Design of Monopile Foundations to Support the DTU 10 MW Offshore
Scharff, R., Siems, M., 2013. Monopile foundations for offshore wind turbines – solutions Wind Turbine. Master’s Thesis. Norwegian University of Science and Technology.
for greater water depths. Steel Constr. 6 (1), 47–53. Wang, R., Han, L.H., Zhao, X.L., Rasmussen, K.J.R., 2016. Analytical behavior of concrete
Schmoor, K.A., Achmus, M., 2015. Optimum geometry of monopiles with respect to the filled double steel tubular (CFDST) members under lateral impact. Thin-Walled
geotechnical design. J. Ocean Wind Energy 2 (1), 54–60. Struct. 101, 129–140.
Schwanitz, V.J., Wierling, A., 2016. Offshore wind investments – realism about cost Wang, F.C., Han, L.H., Li, W., 2018. Analytical behavior of CFDST stub columns with
developments is necessary. Energy 106, 170–181. external stainless steel tubes under axial compression. Thin-Walled Struct. 127,
Senanayake, A.I.M.J., 2016. Design of Large Diameter Monopiles for Offshore Wind 756–768.
Turbines in Clay. PhD Thesis. University of Texas at Austin. Wang, W.Q., Wu, C.Q., Li, J., Liu, Z.X., Zhi, X.D., 2019. Lateral impact behavior of
Shafieifar, M., Farzad, M., Azizinamini, A., 2017. Experimental and numerical study on double-skin steel tubular (DST) members with ultra-high performance fiber-
mechanical properties of Ultra High performance concrete (UHPC). Constr. Build. reinforced concrete (UHPFRC). Thin-Walled Struct. 144, 106351.
Mater. 156, 402–411. Xu, H.B., 2015. Research on the Performance of HRB 500 Bars Reinforced Prestressed
Sheng, D., Eigenbrod, K.D., Wriggers, P., 2005. Finite element analysis of pile installation Ultra-high Performance Concrete Beams. PhD Thesis. Beijing University of
using large-slip frictional contact. Comput. Geotech. 32, 17–26. Technology, China.
Shimizu, M., Tatsumi, F., Ishikawa, T., Hattori, A., Kawano, H., 2013. Experimental study Yan, J.B., Zhang, W., Liew, J.Y.R., Li, Z.X., 2016. Numerical studies on shear resistance of
on ultimate strength of concrete filled double tubular steel with shear connector. Int. headed stud connectors in different concretes under Arctic low temperature. Mater.
J. Steel Struct. 13 (1), 49–54. Des. 112, 184–196.
Shirzadeh, R., Deveriendt, C., Bidakhvidi, M.A., Guillaume, P., 2013. Experimental and Ye, Y., Han, L.H., Guo, Z.X., 2017. Concrete-filled bimetallic tubes (CFBT) under axial
computational damping estimation of an offshore wind turbine on a monopile compression: analytical behaviour. Thin-Walled Struct. 119, 839–850.
foundation. J. Wind Eng. Ind. Aerodyn. 120, 96–106. Zhang, F.R., 2017. Dynamic Analysis of Steel Confined Concrete Tubular Columns
against Blast Loads. PhD Thesis. The University of Adelaide, Austria.

17

You might also like