You are on page 1of 16

Ships and Offshore Structures

ISSN: 1744-5302 (Print) 1754-212X (Online) Journal homepage: http://www.tandfonline.com/loi/tsos20

A parametric study of spar-type floating offshore


wind turbines (FOWTs) by numerical and
experimental investigations

Sajad Rahmdel, Baowei Wang, Changwan Han, Kwanghoon Kim & Seonghun
Park

To cite this article: Sajad Rahmdel, Baowei Wang, Changwan Han, Kwanghoon Kim &
Seonghun Park (2015): A parametric study of spar-type floating offshore wind turbines
(FOWTs) by numerical and experimental investigations, Ships and Offshore Structures, DOI:
10.1080/17445302.2015.1073865

To link to this article: http://dx.doi.org/10.1080/17445302.2015.1073865

Published online: 19 Aug 2015.

Submit your article to this journal

Article views: 61

View related articles

View Crossmark data

Full Terms & Conditions of access and use can be found at


http://www.tandfonline.com/action/journalInformation?journalCode=tsos20

Download by: [Indian Institute of Technology Madras] Date: 25 January 2016, At: 23:33
Ships and Offshore Structures, 2015
http://dx.doi.org/10.1080/17445302.2015.1073865

A parametric study of spar-type floating offshore wind turbines (FOWTs) by numerical and
experimental investigations
† †
Sajad Rahmdel , Baowei Wang , Changwan Han, Kwanghoon Kim and Seonghun Park∗
School of Mechanical Engineering, Pusan National University, Busan, Republic of Korea
(Received 12 January 2015; accepted 14 July 2015)

The use of spar platforms as a substructure for floating offshore wind turbines (FOWTs) is a new concept that is developing
quickly in the offshore wind industry owning to the excellent stability and adaptability to different water depths. However, the
Downloaded by [Indian Institute of Technology Madras] at 23:33 25 January 2016

lack of studies about the dynamic response and design guidelines of spar-type FOWTs is a barrier to further development of
the offshore wind industry. Therefore, the goal of this study is to carry out dynamic response analysis and to develop design
guidelines for spar-type FOWTs. To achieve this goal, the dynamic responses of full-scale spar-type FOWT models with
different values of three design variables (spar diameter, depth, and concrete ratio) were first numerically obtained in the time
domain and experimentally validated by considering all environmental conditions such as wind, regular wave, and constant
current loads, as well as the mooring line loads. Then, regression and perturbation analyses, which were also validated by the
analysis of variance method, were performed to analyse the effects of the design variables and to propose design guidelines
of spar-type FOWTs.
Keywords: spar platform; floating offshore wind turbine (FOWT); hydrodynamic analysis; design guideline

1. Introduction barge, and two spar platforms. Using the IEC 61400-3 stan-
Wind power is a favourable source among the various types dard, various load cases were analysed, and the dynamic re-
of clean energy, and it is also one of the fastest growing parts sponses of these platforms were compared with each other
of the renewable energy industry, with an average annual (IEC 2009). From the comparison, it was concluded that
growth rate of 25% from the end of 2007 to 2012. Exclud- the barge platform has the highest dynamic motions. The
ing hydropower, wind power has the highest global energy differences in the ultimate and fatigue stresses between the
production capacity among all renewable energy sources. TLP, the semi-submersible, and the spar platforms were not
The total kinetic wind energy generated by the earth’s atmo- significant, except that the TLP had smaller stresses in its
sphere is approximately 900 Terawatts (Lorenz 1955), and tower compared with the spar and semi-submersible sys-
according to Miller’s research (Miller et al. 2011), about 68 tems (Robertson and Jonkman 2011). Kurian et al. (2012)
Terawatts of this wind energy could be harnessed. In spite performed both numerical and experimental studies on the
of this abundant wind power available, only 283 Gigawatts dynamic responses of classic and truss spar platforms con-
of this energy is currently harnessed (McCrone et al. 2012; sidering random wave and current forces. The dynamic
Zervos 2013), although it is estimated that, in 2020, this responses of a 1:100 scale model were measured in the fre-
amount will increase to 1 Terawatt (Gsänger and Pitteloud quency domain from experiments. The damping ratio and
2012). Wind power can be harnessed both onshore and off- natural periods of the system were obtained using a free-
shore. By the end of 2012, about 98% of the wind turbines decay test and the response amplitude operators (RAOs)
were installed onshore; however, because of the various ad- were calculated for heave, surge, and pitch motions. From
vantages of offshore wind power, global trends are moving good agreement between experimental results and numeri-
from onshore to offshore generation. cal results, it was concluded that coupled wave and current
The application of spar platforms to the wind indus- forces would result in higher surge, heave, and pitch re-
try is a new concept, and a limited number of studies are sponses for the classic spar compared to the truss spar.
available (Seebai and Sundaravadivelu 2013; Ku and Roh Furthermore, using the same random waves, the responses
2014). Robertson and Jonkman (2011) studied the dynamic in surge, heave, and pitch for both types of spar platforms
responses of six offshore wind turbine platforms, includ- increased as the current velocities increased. The experi-
ing two tension leg platforms (TLPs), a semi-submersible, a mental results of classic and truss spar platforms subjected


Corresponding author. Email: paks@pusan.ac.kr

Both authors contributed equally to this work.


C 2015 Taylor & Francis
2 S. Rahmdel et al.

to multidirectional waves also show that multidirectional figurations that are generated by the method of Design of
waves generate smaller dynamic motions in comparison to Experiment (DOE) are presented, and then, methods for the
long-crested waves (Kurian et al. 2012). numerical simulation of these full-scale models, consider-
Utsunomiya et al. (2013) developed a hybrid spar with ing all environmental loads (wind, waves, and current), are
precast concrete segments at the lower portion and steel described. In Section 3, scale model simulations and ex-
segments at the higher portion of the spar. An at-sea exper- periments conducted to validate the current method of the
iment using a 1/10-scale model of the hybrid-SPAR model full-scale model simulation are explained. In Section 4, all
mounting a 1 KW wind turbine was then conducted. The the simulation and experimental results are presented and
experimental data were compared with simulation results discussed. Then, based on the results of full-scale model
acquired from a preliminary simple in-house code and it simulations, design guidelines for spar-type FOWTs are
was concluded that hybrid-SPAR FOWT was feasible for finally suggested.
being installed in offshore areas (Utsunomiya et al. 2013).
Liu et al. (2013) scrutinised non-linear motions of spar
2. Full-scale model simulations
platform hull by means of numerical and experimental anal-
The configuration of the Pusan National University (PNU)
Downloaded by [Indian Institute of Technology Madras] at 23:33 25 January 2016

yses. Then, the results of response amplitude and time histo-


ries of the structure were obtained for various wave frequen- 2.5 MW spar-type FOWT which is fabricated in PNU is
cies. It was perceived that the response of the structure is used for full-scale model simulations (Figure 1).
sensitive to different wave frequencies (e.g., when wave fre-
quency approached heave natural frequency, large chaotic 2.1. Definition of motions
motions were revealed on both heave and pitch motions)
The spar-type FOWT is a floating platform which is at-
(Liu et al. 2013).
tached to the seabed by a mooring line system. The entire
The coupling between the wind turbine and the hull,
system is considered to undergo rigid body motions (Agar-
as well as the interface between the hydrodynamic loading
wal and Jain 2003; Veritas 2010). A coordinate system is
and the structural response for WindFloat – the world’s
generally defined with the origin in the mean water level
second full-scale floating wind turbine – was studied by
(MWL), where hydrodynamic problems are encountered.
Roddier et al. (2009). In addition, hydrodynamic analysis
The rigid body motions of a floating platform have six de-
of the hull as well as the coupling hydrodynamics of the
grees of freedom (DOFs), three of which are translational,
hull with wind turbine aerodynamics was described in the
and the others are rotational.
literature (Cermelli et al. 2009). Considering the coupling
effects between the substructure and wind turbine, fatigue
analysis, as well as strength analysis, was carried out on the
structure of FOWTs (Aubault et al. 2009).
However, limited studies have conducted to fully un-
derstand the dynamic behaviours of spar-type FOWTs. The
International Electrotechnical Commission (IEC) investi-
gated design standards for offshore wind turbines (IEC
2009), but design standards considering spar platforms
were not covered. The ISO offshore structural design stan-
dards and DNV-OS-C201 are other sources of design stan-
dards that are used in the oil and gas industry, but none of
these standards are specifically made for spar-type FOWTs
(ISO 2003; DNV 2011, 2012). Recently, DNV, ABS, and
ClassNK have presented offshore standards for the design of
floating wind turbine structures, but none of these standards
provide detailed guidelines about how to reduce dynamic
motions of the structure by selecting appropriate design
variables (Kyokai 2012; DNV 2013; ABS 2014).
A barrier to the continued growth and development of
the offshore wind turbine industry is the lack of design
standards for FOWTs. Therefore, the aim of this study is to
analyse the dynamic behaviour of spar-type FOWTs with
both numerical and experimental results in the time domain
Figure 1. PNU 2.5 MW spar-type floating offshore wind turbine
and to suggest design guidelines of spar-type FOWTs. The and three design variables that are total draft (TD), spar diameter
brief outline of the current study is as follows. In Section 2, (SD), and concrete ratio (CR). (This figure is available in colour
13 full-scale spar-type FOWT models with different con- online.)
Ships and Offshore Structures 3

Then, the linearised free-surface condition reduces to

∂φ ω2 ∂φ ω2
− φ=0 − φ=0 (2)
∂z g ∂z g

where ω is the wave frequency. Seabed boundary conditions


are ∇φ = 0 when z → ∞ (for deep water) ∂φ ∂z
= 0 at z =
−d (for shallow water depth d).
To simplify the problem, the velocity potential is de-
composed into the incident wave velocity potential (ϕI ),
the diffracted wave velocity potential (ϕD ), and the radi-
ated wave potential in six DOFs (ϕj with j = 1, 2, . . . , 6).
The total velocity potential due to unit-amplitude incident
wave is acquired by applying a linear superposition of ve-
Downloaded by [Indian Institute of Technology Madras] at 23:33 25 January 2016

locity components. The total velocity potential is defined


by
⎡ ⎤

6
φ = ϕe−iωt = ⎣(ϕI + ϕD ) + ϕj Xj ⎦ e−iωt (3)
j =1

where Xj is the six-DOF structural motion. Here, the inci-


dent wave velocity potential of finite water depth d is

−igζ cosh [k (z + d)] eik(x cos θ+y sin θ+α) e−iωt


ϕI e−iωt =
ω cosh (kd)
Figure 2. Environmental loading conditions with (a) direction
of environmental loads, and (b) thrust curve as a function of hub (4)
wind speed for PNU 2.5 MW turbine. (This figure is available in
colour online.) where θ is the wave direction and ζ is the wave elevation.
The wave frequency ω is related to the wave number k by
the following equation:
• Translational motions: η1 = surge (along x); η2 =
sway (along y); η3 = heave (along z). ω2 = gk tanh (kd) (5)
• Rotational motions: η4 = roll (about x); η5 = pitch
(about y); η6 = yaw (about z). The hydrodynamic pressure (P) on the surface of the
structure is calculated from the Bernoulli equation (Wilson
2003) as follows:
In this study, the directions of the environmental loads
are defined as the angles between the propagation and pos- ∂φ
itive x-axis measured in the counterclockwise direction P = −ρ (6)
∂t
(Figure 2a).
where ρ is the water density. The hydrodynamic pressure
of the structure is then integrated over the wetted surface
2.2. Calculation of wave load of the body to find various fluid forces acting on the struc-
The inertia and diffraction forces which act on the main ture. Finally, the wave drag force acting on the body of the
body of the spar-type FOWT are calculated using the linear structure is calculated by the equation, 12 ρCd DV |V |, where
diffraction theory, and the wave drag forces acting on the Cd = 1 is the drag coefficient and D is the element diameter.
body of the structure are calculated using Morison’s equa-
tion. To obtain the wave drag forces, the governing equation 2.3. Environmental conditions
of velocity potential (φ) and the relative velocity between
the flow and the body (V ) are first given by The operating environmental parameters were selected
from the three-dimensional (3-D) contour surface of the
joint distribution for wind and waves (Johannessen et al.
∇ 2φ = 0 2002) and, to generate the 3-D contour surface, the simul-
(1)
V = ∇φ taneous wind and wave measurements from the northern
4 S. Rahmdel et al.

Table 1. Environmental parameters.

Wind Wave

Thrust force (N) Hub height (m) Speed (m/s) Height, Hs (m) Period, T (s) Current Speed (m/s)

2.37E05 80 11.3 3.1 10.1 1.0

North Sea between 1973 and 1993 were used. The typical forces, and the quadratic transfer functions for a slowly
current speed in this region was selected based on offshore varying wave drift force calculation that was obtained by
regulations (Veritas 2010). By considering uniform wind the far-field method (Figure 4). The output data of AQWA
field and fixed blade pitch angle, the thrust force was cal- Hydrodynamic Diffraction were then inputted to the AQWA
culated for the PNU spar-type FOWT using GH-Bladed Hydrodynamic Time Response module in ANSYS AQWA
software (version BI011) and applied to the centre of the along with other environmental loads (wind, current, wave,
Downloaded by [Indian Institute of Technology Madras] at 23:33 25 January 2016

nacelle (hub height) in the current study. The incident di- mooring, etc.). The equation of motion in the time do-
rection of the environmental loads was 0◦ (Table 1), and main was solved to obtain the time series for the dynamic
the regular wave was inputted with an angular frequency of behaviour. A time step of 0.5 s was selected, and the simu-
∼0.62 rad/s and a wave height of 3.1 m (Table 1). lations were run for 2200 s.
The superstructure was simplified by ignoring rotational
movements of blades, and the thrust force generated by
blades was calculated by GH-Bladed software (Figures 2b 2.5. Design of experiment (DOE)
and 3). In the present simulation, a constant wind speed of In the current study, three design variables (depth, spar di-
11.3 m/s was considered at hub height and a corresponding ameter, and concrete ratio), each one with three levels, were
thrust force of 237 kN was generated from the thrust force selected (Figure 1, Table 2) because those are the most in-
curve as a function of hub wind speed as shown in Figure 2b. fluential design variables for the dynamic behaviours of
spar-type FOWTs. With these design variables, 13 different
spar-type substructure configurations were generated by the
2.4. Simulation procedure
Box-Behnken Design (BBD) method (Box and Behnken
ANSYS AQWA suite (version 14.0) was used to carry out 1960) and modelled using the ANSYS Design-Modeler
the dynamic analysis. The AQWA Hydrodynamic Diffrac- (version 14.0) (Figure 5). For the dynamic behaviour anal-
tion model in ANSYS AQWA was first used as a preproces- ysis of rigid bodies, the required information of mass and
sor to obtain the hydrostatic loads, first-order wave exciting inertia properties was calculated by including a contribu-
tion from the superstructure in the current study (Table 3);
moments of inertia for the models were computed in terms
of their centres of gravity (CGs).

2.6. Full-scale model simulations


For all full-scale models, the same specifications of the
mooring catenary system as those in the OC3-Hywind
project (Jonkman and Musial 2010) were used. These cate-
nary lines were connected to the spar hull at a depth of 70
m below the MWL, while the connection points of the cate-
nary lines were spread equally around the exterior surface
of the spar hull. The anchors were placed on the sea bed at
a water depth of 320 m and with a radius of 853.9 m from
the centreline (CL) of the spar platform. The catenary lines
had an un-stretched length of 902 m.

2.7. Regression analysis


For most response surface methodology (RSM) prob-
lems, the relationships between the system responses
Figure 3. Simplified floating offshore wind turbine system. (This and variables are unknown. Normally, a first-order-, a
figure is available in colour online.) second-order, or a higher order polynomial could be used
Ships and Offshore Structures 5

Table 2. Design variables and their levels.

Levels

Design variables Notation Unit Low (−1) Medium (0) High ( + 1)

Depth D m 100 110 120


Spar diameter SD m 8 9 10
Concrete ratio CR – 0.4 0.7 1.0
Downloaded by [Indian Institute of Technology Madras] at 23:33 25 January 2016

Figure 4. Procedure of full-scale model simulation.

to approximate the relationship between the responses and If there is a curvature in the system, a higher order
variables. These mathematical relations are only reasonable polynomial such as a second-order model should be used
in a certain range of variables, not in the overall domain. (Montgomery 2001) by:
If the response is well modelled by a linear function of
variables, the approximating function is a first-order model 
k 
k 
(Montgomery 2001): y = β0 + βi xi + βii xi2 + βij xi xj + ε
i=1 i=1 i<j

y = β0 + β1 x1 + β2 x2 + · · · + βk xk + ε (7) (8)

where Xi or j are the variables, while β i and β ij are the regres-


sion coefficients for the main effects of the variables and for
interaction effects between the variables, respectively. The
coefficients in Equations (7) and (8) can be estimated us-
ing the least squares method (Montgomery 2001). The last
term, ε, is the stochastic error. Almost all RSM problems
use one or both of the aforementioned models.

2.8. Perturbation analysis


To determine which design variables are significant for each
motion, a perturbation analysis was conducted using regres-
sion analysis results. The perturbation plots help compare
the effects of all variables at a particular point in the design
Figure 5. Geometric representation of Box–Behnken design space. In this study, the midpoints (coded 0) of all variables
(BBD) for three design variables. were selected as the reference points. The responses were
6 S. Rahmdel et al.

Table 3. Mass, centre of gravity (CG), and moment of inertia for full-scale simulation models generated by design of experiments (DOE).

Design variables Inertia with respect to CG (kg·m2 )

Model no. D SD CR Mass (kg) CG (m) (z-axis) Ixx Iyy Izz

1 −1 0 −1 6.523E6 −57.410 9.997E9 9.997E9 1.015E8


2 −1 −1 0 5.162E6 −53.836 8.785E9 8.785E9 6.632E7
3 −1 +1 0 8.063E6 −61.211 1.160E10 1.160E10 1.492E8
4 −1 0 +1 6.532E6 −58.397 1.034E10 1.034E10 1.015E8
5 0 0 0 7.185E6 −64.848 1.273E10 1.273E10 1.117E8
6 0 +1 −1 8.869E6 −66.789 1.399E10 1.399E10 1.642E8
7 0 −1 +1 5.679E6 −60.491 1.092E10 1.092E10 7.299E7
8 0 +1 +1 8.869E6 −68.465 1.462E10 1.462E10 1.642E8
9 0 −1 −1 5.679E6 −59.902 1.067E10 1.067E10 7.299E7
10 +1 −1 0 6.195E6 −66.804 1.328E10 1.328E10 7.967E7
11 +1 0 −1 7.838E6 −70.553 1.519E10 1.519E10 1.220E8
+1 +1 −71.826
Downloaded by [Indian Institute of Technology Madras] at 23:33 25 January 2016

12 0 7.838E6 1.578E10 1.578E10 1.220E8


13 +1 +1 0 9.675E6 −74.850 1.782E10 1.782E10 1.793E8

Figure 6. Scale model of the PNU spar-type floating offshore wind turbine: (a) dimensional parameters and (b) configuration of
mooring lines.

plotted by changing only one variable over its range while


holding the other variables constant.
Table 4. Moment of inertia for the scale models of the PNU
3. Scale model simulations and experiments spar-type floating offshore wind turbine with different centres of
gravity (CG1 and CG2) for simulations and experiments.
Simulations and experiments were conducted using the
scale model of the PNU spar-type FOWT to validate our Items CG1 CG2
full-scale model simulation.
Centre of gravity −0.77 −0.661
(m) (from the top
3.1. Scale model prototype of the hull)
Moment of inertia 9.23735 9.8374
Two 100:1-scaled prototypes of the PNU spar-type FOWT in roll (kg·m2 )
with different CGs, holding the dimensional parameters Moment of inertia 9.24018 9.8403
and the total masses of the scale models unchanged, in pitch (kg·m2 )
were constructed for experimental measurements (Figure 6, Moment of inertia 0.06466 0.06592
in yaw (kg·m2 )
Table 4).
Ships and Offshore Structures 7

Table 5. Dimensions of mooring lines installed in the scale mod- model simulation (Section 2), while the experiments were
els of the PNU spar-type floating offshore wind turbine. carried out in a water tank with 100 m length, 8 m width,
Mooring lines Dimension and 3.5 m depth with the scaled prototypes moored at 30
m downstream from the wave maker (Figure 7). In experi-
Mooring line length (m) 4.34 ments, a wave maker was used to create various wave forces
Total mass (kg) 21 with frequencies ranging from 2.09 rad/s (∼0.333 Hz) to
Number 3
Pretension (N) 4.8
6.28 rad/s (∼1 Hz) for different scale models (Figure 7), and
different wave heights were generated to simulate various
In-line stiffness (N/m) K1 7.25 environmental conditions and to find the dynamic responses
K2 17.05 of models in those conditions (Table 6). A data acquisition
system, which amplifies and transfers the measured data to
Fairlead locations (m) L1 0.683
L2 0.163 computer, and a camera were placed on the right side of
the water tank. The height and frequency of waves were
measured by a wave gauge (Figure 7). A trinocular cam-
Downloaded by [Indian Institute of Technology Madras] at 23:33 25 January 2016

The draft size of the scale models was about 1.2 m, era system consisting of a trinocular camera, three LEDs,
and three light-emitting diodes (LEDs) were attached to the a camera controller, and a computer were used to measure
nacelle of scale models for experiments. All scale mod- the position and attitude of the LEDs attached to the nacelle
els were moored to a water tank by three catenaries and of the scale models. The camera captures the movement of
two kinds of fairlead locations and stiffness of mooring the structure. The camera controller then sends the data to
lines were used (Table 5). The same experimental condi- the computer, and the computer calculates the position and
tions for 100:1-scaled models were applied to scale model attitude of the structure in real time (Figures 7 and 8).
simulation. A summary of the simulation and experimental analy-
sis process is illustrated in Figure 8. The time-series data of
pitch motions acquired from both the scale model simula-
3.2. Experimental set-up and procedure tions and the experiments were converted to the frequency
Four cases were considered for scale model simulations domain using Fast Fourier Transform (FFT). The RAO val-
and experiments under only the wave loading condition ues of the pitch motions were calculated by the following
(Table 6). equation at an angular frequency of ∼0.62 rad/s under reg-
In each case, scale models with different combinations ular wave:
of CGs, moments of inertia, fairlead locations, and mooring
line stiffness were used. In scale model simulations, we em- Pitch amplitude
RAO = (9)
ployed the same methodology as described in the full-scale Wave amplitude

Table 6. Scale model simulation and experimental cases and corresponding loading conditions.

Regular wave loading conditions (LC) for experiments and simulations

LC1 LC2 LC3

Wave Wave Wave Wave Wave Wave


frequency height frequency height frequency height
Case no. (rad/s) (cm) (rad/s) (cm) (rad/s) (cm)

Case 1 K1 2.62 3.5 3.14 3.1 4.19 4.2


L1
CG1

Case 2 K2 2.62 4.0 3.14 3.5 4.19 4.2


L1
CG1

Case 3 K1 2.62 2.4 3.14 3.1 4.49 4.2


L2
CG1

Case 4 K1 2.09 2.9 3.14 3.7 4.19 2.4


L1
CG2
8 S. Rahmdel et al.

The time series of the surge, pitch, and heave motions


were obtained by running hydrodynamic motion analysis in
ANSYS AQWA software for different loading conditions.
Time-series results were later used in regression analysis
to model a mathematical equation which represents the
dynamic response of the structure. As an example, Figure 10
illustrates the results of the simulation for Model 3. As
shown in Figure 10c, the initial static equilibrium heave
position was shifted upwards because a fixed point on the
top of the hull was considered as the reference point. The
structure oscillated by approximately 0.7 m (mainly by wave
load).
Downloaded by [Indian Institute of Technology Madras] at 23:33 25 January 2016

4.3. Regression analysis results


Using the mean values of the time series of surge, pitch
and heave motions (Table 7), the regression models are
proposed for each motion by calculating the regression co-
efficients with Design Expert software (version 7).
The generated regression models for each of the dis-
placements with the design variables in the respective DOFs
are as follows:

(a) Regression model for surge motion

1  
Figure 7. Experimental device: (a) layout of a wave tank and (b) = 0.040 + 3.023 × 10−3 A
sensor locations in experiments. (This figure is available in colour surge
   
online.) + 7.478 × 10−3 B + 5.560 × 10−4 C
   
− 2.027 × 10−3 AB − 3.009 × 10−3 B 2
4. Results and discussion  
− 4.357 × 10−4 C 2
4.1. Scale model simulations and experiments
In order to validate the accuracy of our numerical analysis, (b) Regression model for pitch motion
the simulation and experimental results of pitch motions
for the scale model of the PNU spar-type FOWT were com- 1
pared. For more accuracy, the coupling effects between the = 0.55 + (0.074) A + (0.16) B
pitch
structure and its mooring line system were considered. The
pitch motion results obtained from the numerical analysis + (0.016) C + (0.013) AB
 
were in good agreement with the experimental results under + 7.051 × 10−3 BC
 
all loading conditions (Figure 9). − (0.016) B 2 − 9.263 × 10−3 C 2

4.2. Full-scale model simulations (c) Regression model for heave motion
The effects of the selected design variables on the dynamic 1
motions of the PNU spar-type FOWT were investigated by = 1.04 − (0.037) A + (0.42) B
heave
the numerical analysis described in Section 2, using its full-
+ (0.015) C − (0.028) AB − (0.24) AC
scale model. Similarly to our previous scale model simula-
tions and experiments, the coupling between the structure + (0.05) A2 − (0.063) B 2 − (0.12) C 2
and the mooring line, as well as hydrodynamic forces (both + (0.28) A2 C − (0.2) AB 2
inertia and drag) on the mooring lines, was also consid-
ered to improve accuracy in the simulation. With the wind, where A is the depth of spar, B is the spar hull diameter, and
wave, and current forces distributed in the same direction C is the concrete ratio in the ballast section of spar.
(0◦ along the positive x-axis), the motions in the surge, These regression models were validated from an anal-
pitch, and heave were significant among six-DOF motions ysis of variance (ANOVA) (Table 8). The adequacy of the
(Table 7). models was assumed at the 95% confidence level. In this
Ships and Offshore Structures 9
Downloaded by [Indian Institute of Technology Madras] at 23:33 25 January 2016

Figure 8. Scale model simulations and experiments: (a) simulation procedure, (b) experimental set-up, and (c) Fast Fourier Transform
(FFT). (This figure is available in colour online.)

method, the calculated F-value of the developed models dicate that the developed models are sufficiently accurate
should be larger than the tabulated value of the F-table for (Figure 12).
the 95% confidence level, which can ensure that the gener-
ated regression model is adequate. The calculated F-values
were 234.18, 2912.6, and 247.66, which were significantly 4.4. Perturbation analysis results
larger than the tabulated F-table: 4.28, 4.88, and 3.84 for The dynamic responses in the surge and pitch were both
the surge, pitch, and heave motions, respectively. In addi- sensitive (i.e., steep slope) to the depth and the spar diame-
tion, the P-values of all the regression models were much ter, whereas the concrete ratio is relatively insensitive (i.e.,
lower than 0.05, which indicates that the regression models flat slope) to the two responses (Figure 13). The changes
are adequate at the 95% confidence level. in the spar diameter and depth have significant influence
The validity of the developed models can be also proven on the pitch motion, as shown in the perturbation plot of
by the adjusted correlation coefficient (adjusted R-squared). pitch motion of Figure 13b. For the displacement in heave,
This coefficient is an indicator of the variability in the ob- the spar diameter is dominant over the other two design
served output by each variable and by interaction of the two variables (Figure 13c).
design variables. The coefficients of the regression models The dynamic behaviours of surge, heave, and pitch mo-
for the surge, heave, and pitch were 99.15%, 98.80%, and tions were also analysed in terms of interaction between two
99.94%, respectively, showing the highly predictive ability design variables from the 3-D surface perturbation plots of
of the regression models. Furthermore, a residual analysis the regression models. Similar conclusions to those from
was carried out to check the accuracy of the regression mod- the 2-D perturbation plots were made. The surge motion is
els obtained. The results that the residuals are distributed sensitive to the depth and the spar diameter. When the spar
almost along straight lines suggest that the residual errors diameter and the depth size increase, the structure will ex-
are distributed normally (Figure 11), and the linearly dis- perience smaller motion in the surge direction (Figure 14a).
tributed points along straight lines in the figures of the actual The pitch motion decreases with increase in the spar diam-
values versus predicted values by the regression models in- eter and the depth, and vice versa. However, the concrete
10 S. Rahmdel et al.
Downloaded by [Indian Institute of Technology Madras] at 23:33 25 January 2016

Figure 10. Time series data of (a) surge, (b) pitch, and (c) heave
motions for Model 3.

ratio is relatively insensitive to the pitch motion (Figure


14b and 14c). For heave motion, the spar diameter is dom-
Figure 9. Comparisons of response amplitude operators (RAO) inant over the other design variables, and the displacement
between the numerical and experimental results of pitch motions in heave decreases substantially with the increase in spar
under four different loading conditions: (a) case 1 (K1, L1, and
diameter (Figure 14d).
CG1), (b) case 2 (K2, L1, and CG1), (c) case 3 (K1, L2, and
CG1), and (d) case 4 (K1, L1, and CG2). (This figure is available
in colour online.)
Ships and Offshore Structures 11

Table 7. Average values of the time series of surge, heave, and pitch motions computed from full-scale model simulations.

Design variables Displacements

Model no. Run no. D (A) SD (B) CR (C) CG (m) (z-axis) Surge (m) Heave range (m) Pitch (◦ )

1 11 100 9 0.4 −57.410 27.465 2.092 4.833


2 1 100 8 0.7 −53.836 40.859 1.236 10.542
3 2 100 10 0.7 −61.211 23.015 0.579 2.684
4 13 100 9 1 −58.397 27.512 0.651 4.317
5 3 110 9 0.7 −64.848 24.363 0.963 3.316
6 12 110 10 0.4 −66.789 22.880 0.780 2.283
7 8 110 8 1 −60.491 33.068 1.980 7.178
8 10 110 10 1 −68.465 22.158 0.806 1.996
9 7 110 8 0.4 −59.902 34.945 2.494 7.912
10 6 120 8 0.7 −66.804 29.055 2.619 5.341
11 4 120 9 0.4 −70.553 23.896 1.137 2.823
−71.826
Downloaded by [Indian Institute of Technology Madras] at 23:33 25 January 2016

12 5 120 9 1.0 23.088 1.014 2.508


13 9 120 10 0.7 −74.850 22.082 0.843 1.619

4.5. Effects of design variables on the stability of 4.5.1. Effects of changes in spar diameter on dynamic
spar-type floating offshore wind turbines motions
(FOWTs)
Using the results of this study, the effects of design vari- ◦ Increases in the spar diameter of the structure greatly
ables on the stability of spar-type FOWTs are summarised reduce the heave, surge, and pitch motions of the
as follows, and they are considered as the influences of structure.
changes in each design variable – while maintaining other ◦ Increases in the spar diameter can decrease the struc-
design variables constant – on the dynamic motions of the tural dynamic motions, and as a result, the overall
structure. These guidelines are applicable for the situation stability of the structure will increase. However, by
studied in this paper and further studies are required before increasing the spar diameter, the weight of the struc-
their applications on all conditions could be confirmed. ture increases, which leads to economic issues.

Table 8. Summary of analysis of variance (ANOVA) for the regression models of surge, heave, and pitch motions.

Source of variation Regression model Residual error Total Adjusted R-squared Predicted R-squared
DOF 6 6 12 99.15% 98.22%

Surge motion SS 5.65E−04 2.41E−06 5.68E−04


MS 9.42E−05 4.02E−07
F-value 234.18
P <0.0001

F-table (0.05, 6, 6) = 4.28

Heave motion DOF 10 2 12 98.2%


SS 2.27 6.721E−03 2.28
MS 0.23 3.361E−03
F-value 67.56
P <0.0147

F-table (0.05, 3, 9) = 3.84

Pitch motion DOF 7 5 12 99.94% 99.80%


SS 0.26 6.31E−5 0.26
MS 0.037 1.26E−5
F-value 2912.6
P <0.0001

F-table (0.05, 7, 5) = 4.88

Note: SS stands for sum of squares, DOF stands for degree of freedom, and MS denotes the mean square. MS is the SS to DOF ratio, and the F-value is
the ratio of the MS of the regression model to the MS of the residual error.
12 S. Rahmdel et al.
Downloaded by [Indian Institute of Technology Madras] at 23:33 25 January 2016

Figure 11. Normal distribution plots as a function of residual


errors that are calculated from differences in the dynamic motions
between simulations and regression models: (a) surge, (b) pitch, Figure 12. Comparison between the mean values of dynamic
and (c) heave. (This figure is available in colour online.) motions obtained by simulations and their corresponding values
from regression models: (a) surge, (b) pitch, and (c) heave. (This
figure is available in colour online.)

◦ An optimal spar diameter should be selected so that


the structure has the lowest weight while maintain- 4.5.2. Effects of changes in depth on dynamic motions
ing the stability of the structure. Selection of this
optimal spar diameter is strongly affected by the ◦ With increases in the depth of the structure, the surge
environmental conditions of the operational region. and pitch motions decrease.
Ships and Offshore Structures 13
Downloaded by [Indian Institute of Technology Madras] at 23:33 25 January 2016

Figure 13. 2-D perturbation plots for the design variables: (a)
surge, (b) pitch, and (c) heave.

4.5.3. Effects of changes in concrete ratio on dynamic


motions

◦ Compared to other design variables, the concrete ratio


has a trivial influence on the dynamic motions of
the structure. Figure 14. 3-D surface perturbation plots between (a) the surge
and the two interacted variables of spar diameter (SD) and total
Overall, it is concluded that among all design variables, draft (TD), (b) the pitch and two integrated variables of SD and TD,
the spar diameter has the highest influence on the stability of (c) the pitch and the two integrated variables of SD and concrete
the structure. The accurate dimension of this design variable ratio (CR), and (d) the heave and the two integrated variables of
SD and TD. (This figure is available in colour online.)
14 S. Rahmdel et al.

can highly increase the stability. The depth also has great funded by the Korea Government Ministry of Trade, Industry
influence on the stability, but the influence of the depth and Energy; and by Leading Foreign Research Institute Recruit-
is less than the influence of the spar diameter. Increases ment Program through the National Research Foundation of Korea
(NRF) funded by the Ministry of Science, ICT & Future Planning
in the depth increase the overall stability of the structure, [Grant No. 2013044133].
although they increase the heave motion of the structure as
well. Compared with the spar diameter and the depth, the
concrete ratio has the least influence on the overall dynamic
motions of the structure. ORCID
Regarding three design variables, the spar diameter and Sajad Rahmdel http://orcid.org/0000-0002-8310-3149
the depth are related to the overall shape of the substruc- Baowei Wang http://orcid.org/0000-0002-2985-9906
ture, while the concrete ratio is responsible for moving the
centre of mass as low as possible. As expected, higher slen-
derness ratio of the cylindrical substructure with higher References
depth and lower spar diameter reduced the stability of spar- Agarwal A, Jain A. 2003. Dynamic behavior of offshore spar
Downloaded by [Indian Institute of Technology Madras] at 23:33 25 January 2016

type FOWTs. However, it was unexpected that the concrete platforms under regular sea waves. Ocean Eng. 30:487–516.
ratio shows the least influence on the stability of spar-type [ABS] American Bureau of Shipping. 2014. Guides for build-
FOWTs. ing and classing floating offshore wind turbine installations.
Houston (TX): American Bureau of Shipping.
Aubault A, Cermelli C, Roddier D. 2009. Wind float: a floating
5. Conclusions foundation for offshore wind turbines – part III: structural
analysis. In: Proceedings of the ASME 2009 28th Interna-
The goals of the present study were (1) to understand the tional Conference on Ocean, Offshore and Arctic Engineer-
dynamic behaviours of spar-type FOWTs in various wave, ing; 2009 May 31–Jun 5. Honolulu (HI): American Society
current, and wind environments; (2) to evaluate the most of Mechanical Engineers.
influential design variable; and (3) to provide design guide- Box GE, Behnken DW. 1960. Some new three level designs for the
study of quantitative variables. Technometrics. 2:455–475.
lines for designers of spar-type FOWTs.
Cermelli C, Roddier D, Aubault A. 2009. Wind float: a floating
Thirteen spar-type FOWT full-scale models were de- foundation for offshore wind turbines – part II: hydrodynam-
signed, and coupled dynamic numerical analyses in the time ics analysis. Proceedings of the ASME 2009 28th Interna-
domain were carried out by considering all environmental tional Conference on Ocean, Offshore and Arctic Engineer-
conditions. The results of the numerical simulation were ing; 2009 May 31–Jun 5. Honolulu (HI): American Society
of Mechanical Engineers.
verified by comparing them with the scale model exper-
DNV. 2011. DNV-OS-C201: structural design of offshore units
imental results. Then, a regression analysis for the sim- (WSD method). Oslo: Det Norske Veritas (DNV).
ulation results was conducted and validated by ANOVA. DNV. 2012. DNV-OS-C106: structural design of deep draught
Finally, the effects of the design variables on the dynamic floating units (LRFD method). Oslo: Det Norske Veritas
motions of the current spar-type FOWT were quantitatively (DNV).
DNV. 2013. DNV-OS-J103: design of floating wind turbine
obtained from a perturbation analysis and some design
structures. Oslo: Det Norske Veritas (DNV).
guidelines were provided. Gsänger S, Pitteloud J-D. 2012. World wind energy report. Bonn:
World Wind Energy Association (WWEA).
[IEC] International Electrotechnical Commission. 2009. IEC
61400-3. Wind turbines – part 3: design requirements for
Acknowledgements offshore wind turbines. Geneva: IEC.
This work was supported by the Human Resources Development [ISO] International Organisation for Standardisation. 2003.
program (No. 20113020020010-11-1-000) of the Korea Institute ISO/CD 19901-4: natural gas industries. Specific require-
of Energy Technology Evaluation and Planning (KETEP) grant ments for offshore structures: part 4: geotechnical and foun-
funded by the Korea government Ministry of Trade, Industry dation design considerations. Geneva: International Organi-
and Energy; and by Leading Foreign Research Institute Recruit- sation for Standardisation.
ment Program through the National Research Foundation of Korea Johannessen K, Meling TS, Haver S. 2002. Joint distribution for
(NRF) funded by the Ministry of Science, ICT & Future Planning wind and waves in the northern north sea. Int J Offshore Polar
(No. 2013044133). Eng. 12:1–8.
Jonkman J, Musial W. 2010. Offshore code comparison collabo-
ration (OC3) for IEA task 23 offshore wind technology and
Disclosure statement deployment. Contract. 303:275–3000.
No potential conflict of interest was reported by the authors. Ku N, Roh M-I. 2014. Dynamic response simulation of an offshore
wind turbine suspended by a floating crane. Ships Offshore
Struct. 1–14. doi:10.1080/17445302.2014.942504
Kurian V, Ng C, Liew M. 2012. Experimental investigation on
Funding dynamic responses of spar platforms subjected to multi-
This work was supported by the Human Resources Development directional waves. In: Proceedings of the Business Engineer-
program (No. 20113020020010-11-1-000) of the Korea Institute ing and Industrial Applications Colloquium (BEIAC); 2012
of Energy Technology Evaluation and Planning (KETEP) grant Apr 7–8. Kuala Lumpur: IEEE.
Ships and Offshore Structures 15

Kyokai NK. 2012. ClassNK guidelines for offshore floating wind tory, US Department of Energy, Office of Energy Efficiency
turbine structure. Tokyo: Nippon Kaiji Kyokai. and Renewable Energy.
Liu L, Zhou B, Tang Y. 2013. Study on the nonlinear dynamical Roddier D, Cermelli C, Weinstein A. 2009. Wind float: a floating
behavior of deepsea spar platform by numerical simulation foundation for offshore wind turbines – part I: design basis
and model experiment. J Vibr Control. 20(10):1528–1537. and qualification process. In: Proceedings of the ASME 2009
Lorenz EN. 1955. Available potential energy and the maintenance 28th International Conference on Ocean, Offshore and Arctic
of the general circulation. Tellus. 7:157–167. Engineering; 2009 May 31–Jun 5. Honolulu (HI): American
McCrone A, Usher E, Sonntag-O’Brien V, Moslener U, Grüning Society of Mechanical Engineers. p. 845–853.
C. 2012. Global trends in renewable energy investment 2012. Seebai T, Sundaravadivelu R. 2013. Response analysis of spar
Frankfurt: UNEP Collaborating Centre, School of Finance & platform with wind turbine. Ships Offshore Struct. 8:94–101.
Management and Bloomberg New Energy Finance. Utsunomiya T, Matsukuma H, Minoura S, Ko K, Hamamura H,
Miller L, Gans F, Kleidon A. 2011. Estimating maximum global Kobayashi O, Sato I, Nomoto Y, Yasui K. 2013. At sea exper-
land surface wind power extractability and associated climatic iment of a hybrid spar for floating offshore wind turbine using
consequences. Earth Syst Dyn. 2:1–12. 1/10-scale model. J Offshore Mech Arctic Eng. 135:034503.
Montgomery D. 2001. Design and analysis of experiments. 5th Veritas DN. Oct. 2010. DNV-OS-E301: offshore standard-position
ed. New York (NY): John Wiley & Sons. p. 427. mooring. Oslo: Det Norske Veritas (DNV).
Robertson AN, Jonkman JM. 2011. Loads analysis of several Wilson JF. 2003. Dynamics of offshore structures. 2nd ed. Hobo-
Downloaded by [Indian Institute of Technology Madras] at 23:33 25 January 2016

offshore floating wind turbine concepts. The Twenty-first In- ken (NJ): John Wiley & Sons.
ternational Offshore and Polar Engineering Conference; 2011 Zervos A. 2013. Renewables 2013 global status report. Paris:
Jun 19–24. Maui (HI): National Renewable Energy Labora- Renewable Energy Policy Network for the 21st Century.

You might also like