Download as pdf or txt
Download as pdf or txt
You are on page 1of 34

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

net/publication/370956528

Global Design Methodology for Semi-Submersible Hulls of Floating Wind


Turbines

Preprint · January 2023


DOI: 10.2139/ssrn.4455440

CITATION READS

1 303

5 authors, including:

Shuaishuai Wang Torgeir Moan


Norwegian University of Science and Technology Norwegian University of Science and Technology
32 PUBLICATIONS 344 CITATIONS 573 PUBLICATIONS 16,599 CITATIONS

SEE PROFILE SEE PROFILE

Zhen Gao
Shanghai Jiao Tong University
254 PUBLICATIONS 6,645 CITATIONS

SEE PROFILE

All content following this page was uploaded by Shuaishuai Wang on 26 September 2023.

The user has requested enhancement of the downloaded file.


1 Global design methodology for semi-submersible hulls
2 of floating wind turbines
3 Wei Lia,b, Shuaishuai Wangc*, Torgeir Moanc,d,e, Zhen Gaoc,e,f, Shan Gaoa,b
4
5 aKey Laboratory of Far-shore Wind Power Technology of Zhejiang Province, Hangzhou
6 Zhejiang China, 311122
7 bPowerchina Huadong Engineering Corporation Limited, Hangzhou Zhejiang China, 311122

8 cDepartment of Marine Technology, Norwegian University of Science and Technology (NTNU),

9 Trondheim, Norway
10 dFaculty of Maritime and Transportation, Ningbo University, Ningbo, China

11 eCentre for Autonomous Marine Operations and Systems (AMOS), Norwegian University of

12 Science and Technology (NTNU), Trondheim, Norway


13 fSchool of Naval Architecture, Ocean and Civil Engineering, Shanghai Jiao Tong University,

14 Shanghai, China.
15
16 Corresponding author: shuaishuai.wang@ntnu.no
17
18 Abstract
19 Global design of the hull is a significant part of the entire design cycle for floating wind
20 turbines (FWTs). It serves as the basis for determining the overall global dimensions of a
21 given floater layout based on the key serviceability and safety requirements. However,
22 because floating wind turbines still are in the infancy stage of development, there is a lack
23 of effective methods and procedures to determine the global dimensions of the floaters.
24 Moreover, the global response characteristics of different hull layouts are not yet clear. To
25 address these limitations, this paper deals with a global design methodology, illustrated for
26 a semi-submersible support structure of a 10-MW FWT. The design methodology and
27 procedures are described and explained in detail. Global design criteria and required values
28 are specified for the case study. A sensitivity study is conducted as a background to
29 determine the floaters' global dimensions. Intact stability is further studied to evaluate the
30 safety of the floaters in operation under wind loads. Global aspects of structural design are
31 analyzed based on wave-induced RAOs of cross-sectional forces and moments and
32 representative environmental conditions for an offshore site. Finally, the main findings and
33 conclusions are summarized. The study provides an essential methodology for global design
34 of the mainstream semi-submersible hulls of FWTs.
35
36 Keywords: Global design methodology; Semi-submersible hull; Floating wind turbine;
37 Sensitivity study; Intact stability; Structural load effects.
38
39 1. Introduction
40 Interest in FWT is overgrowing. GWEC's offshore wind market report in 2021 shows that
41 around 16.5 GW of floating wind installations are forecasted by 2030, although there was
42 only 17MW installed capacity in 2020 [1]. The vast prospect of the floating wind turbine
43 (FWT) lies in that around 80% of offshore wind resources worldwide are in far shore regions,
44 with waters deeper than 60m. However, the bottom-fixed offshore wind turbine
45 installations will not be more economically competitive than the FWT in such deep water

1 | 33
46 areas [2]. Currently, four mainstream FWT types are classified by different substructures
47 and mooring systems: tension leg hull (TLP), semi-submersible, spar, and barge. Among the
48 four concepts, the semi-submersible is the most popular, which occupied about 75.2% of
49 the global floating substructure market share in 2020 [3]. This is due to the following
50 reasons [4]: 1) the semi-submersible FWTs can be deployed in a wide range of water depths,
51 typically from 30m shallow water to deep water; 2) the semi-submersible FWTs are usually
52 moored by catenary lines, which are cheaper than the tension-leg system in TLP; 3) the
53 transportation and installation of the semisubmersible FWTs are relatively simple and cost-
54 effective. A wide variety of semi-submersible hull concepts have been proposed for
55 supporting horizontal axial wind turbines. Most proposed concepts are categorized into two
56 types: 1) the floater has three columns, and the turbine is mounted on one side column; 2)
57 the floater has four columns, and the turbine is mounted on the central column. Another
58 feature that leads to the variety of layouts is that columns are connected by beams or
59 braces. Those characteristics of the structural layouts are summarized from many of the
60 current well-known semi-submersible FWT concepts, such as 5MW Tri-floater[5], 5MW
61 DeepCwind [6], 5MW CSC [7], 8.4 MW WindFloat [8], 10MW OO-Star [9], 10MW DUT Semi
62 FOWT [10], 10MW SPIC [11], 12MW INO WINDMOOR [12], 15MW UMaine VolturnUS-S [13].
63 Offshore wind turbines (OWTs) should be designed to satisfy the requirements of
64 serviceability and safety throughout their intended service life while minimizing the lifecycle
65 costs. Prior knowledge and experiences accumulated from the design of offshore oil and gas
66 hulls and onshore wind turbines serve as a basis for FWT design. To date, the available
67 guidelines for FWT design include IEC TS 61400-3-2 [14](currently under development),
68 DNVGL-ST-0119 [15], and NI572 [16], while they are still immature yet due to the limited
69 operational experience and need to be improved.
70 These standards refer to many other guidelines applied to well-developed marine
71 structures, such as oil and gas platforms. However, the failure consequences, as well as
72 uncertainties in loads and load effects, are different for FWT hull structures. For instance,
73 the failure of oil and gas platforms will typically result in human injury, fatalities, and
74 environmental pollution. However, FWTs are unmanned during the operation, and the
75 consequences of failures would only cause economic loss. On the other hand, oil and gas
76 hulls are mainly subjected to deck loads and wave loads but with limited wind loads. In
77 contrast, in addition to carry the vertical loads from the tower, rotor, and nacelle weight, as
78 well as the wave loads, the FWT hulls also have to withstand huge wind loads. Furthermore,
79 the tight coupling of the aero-hydro-servo-elastic of the FWT system results in strong
80 nonlinearity in the structural load effects. Moreover, FWT will experience various modes,
81 such as operating, parked, faulted, and emergency shutdown, which further exacerbates the
82 complexity of its dynamic responses. The complex and variable environmental loads in FWT
83 support structures result in more significant uncertainties in load effects than those in the
84 oil and gas hulls. If the failure of wind turbines in a farm is correlated, multiple turbine
85 failures might occur. The target safety level for FWT hulls should be based on considerations
86 of the consequences of failure and the trade-off between safety and costs. Then, the safety
87 factors in semi-probabilistic design procedures should be calibrated based on reliability
88 analysis and a comprehensive numerical and experimental assessment of the inherent
89 uncertainties in the design checks.
90 Design is an iterative process, which is subdivided into the three phases, as follows:

2 | 33
91 • selection of floater concept (barge, semi, spar, TLP, etc.) and selection of concept
92 among alternatives of semi-submersible type (number of columns, pontoons and
93 connections between columns- beams, braces).
94 • determination of global dimensions (distance between columns, draught and cross
95 section size of the columns, pontoons and other elements in the water.
96 • determination of hull structure scantlings (thickness of shell plating, stiffener,
97 frames and bulkhead spacing and dimensions.
98 To choose the best concept, full analysis from phases 1-3 for each concept should be
99 conducted. This study focuses on the phase two. Several studies, e.g., [17-21], were
100 conducted to determine the global dimensions of FWT hulls. However, many of today's
101 studies estimate hydrodynamic loads by simplified methods that are not applicable for large
102 semi-submersible hulls. A comprehensive and accurate global design procedure for semi-
103 submersible hulls has not yet been presented in the open literature.
104 Motivated by the background described above, this study deals with a global design for
105 a semi-submersible support structure for a 10-MW FWT. A comprehensive global design
106 methodology is presented, including a description of the design procedure, design criteria,
107 and global parameters. The main layout of the hull, with four columns connected by
108 pontoons and a turbine mounted on the central column, is chosen. Global dimensions of the
109 hulls are obtained based on criteria of serviceability, intact stability, and motion natural
110 period, which are described in detail in Section 2. The dynamic characteristics of global
111 motions and cross-sectional loads are studied. The objectives of the study are:
112 • to present a useful and effective methodology and procedure for the global design
113 of FWT semi-submersible hulls;
114 • to gain an insight into the global response characteristics of the hulls;
115 • to determine global design parameters of the hulls for future detailed design.
116 The global design methodology and procedure, as well as the main results and findings
117 of a case study, are presented in the following sections. More details about the global
118 design work with an application on a case study of the 10-MW semi-submersible FWT are
119 presented in a research project report by Wang et al. [22].
120 While this paper focus on global design methodology and analysis, detailed engineering
121 design needs to be based on coupled dynamic analysis of wind and wave induced load
122 effects and ULS/FLS design checks, considering all possible wind and wave conditions.
123
124 2. Global design methodology
125 Fig. 1 shows the methodology and procedure for design of the semi-submersible hulls.
126 The global design procedure is generally divided into four steps and each step consists of
127 data input, method of analysis and design results. As indicated in Section 1, the design
128 process consists of three design phases. The design step 1 corresponds to the design phases
129 1, while the design phase 2 refers to the design steps 2 and 3. In this study, the design phase
130 2; namely, the design steps 2 and 3 are focused to update the global dimensions of the
131 floater, based on the outcome of the assessment in each step. The design step 4 refers to
132 global aspects of structural design based on wave-induced RAOs of the relevant load effects
133 and representative environmental conditions. This analysis is on the borderline between
134 global issues (phase two) and detailed design (phase three). Global responses of the floater
135 in Steps 1-3 are obtained under the free-floating conditions, while simplified mooring
136 properties are considered in Step 4.

3 | 33
137 In Step 1, the offshore site data (water depth and environmental conditions) and wind
138 turbine properties (structural, mass, and aerodynamic properties) are first given. Then the
139 main configuration, e.g., three/four columns, with/without braces, steel/concrete concepts,
140 etc., of the floater is selected by judgment based on the existing knowledge and experiences
141 about the actual design of semi-submersibles, and based on engineering and research
142 studies as well as the design criteria that are applied. More details about the main
143 configuration information are given in Section 2.1. Next, the global baseline dimensions of
144 the floater, e.g., draft, freeboard, column/pontoon parameters, equivalent plate thickness,
145 etc., are determined according to the existing design and engineering experience. Finally,
146 some global parameters are identified for further sensitivity analysis, while others are fixed
147 (but might be changed in the design iterations).
148 In Step 2, initial design criteria are used to determine the relevant range of the floater's
149 global dimensions. The initial design criteria consist of the serviceability (static heeling angle
150 under maximum mean wind thrust) and motion natural periods in heave, roll, and pitch of
151 the floater, which are described in Sections 2.1.1 and 2.1.4, respectively. Based on Eqs. (1)-
152 (5), the overturning moment due to mean wind thrust force, overall mass, and moment of
153 inertia (including wind turbine structure, floater steel, and ballast mass), added mass, and
154 hydrostatic stiffness in heave, roll, and pitch of the floater are needed to calculate the static
155 heeling angle and natural periods. The added mass and hydrostatic stiffness are calculated
156 using HydroD (WADAM module) code with the panel, structural, and mass models with
157 sufficient accuracy. Subsequently, a sensitivity analysis of the global response characteristics
158 to the identified design parameters of the floater is conducted. Based on this, essential
159 floater parameters are found, and the relevant range of the design parameters is determined
160 based on a check of the initial design criteria. Then, a decision to update the main
161 configuration or global dimensions is made based on evaluating the obtained design space
162 and considerations for following design steps.
163 In Step 3, the intact stability criterion is employed to narrow down further the design
164 space obtained in Step 2. Descriptions of the intact stability criterion are shown in Section
165 2.1.2. The input data includes the wind directions, heeling and righting moment curves, and
166 freeboard, which are used for the intact stability design checks. Sensitivity analysis of the
167 intact stability performance to the floater's global parameters is conducted to reduce the
168 global design cases. Finally, a decision to update the main configuration or global parameters
169 in Steps 1 and 2 is made based on the assessment of characteristics of the intact analysis
170 results of the obtained design space.
171 In Step 4, the dynamic performance of selected hull layouts is studied and compared to
172 finalize the global design layouts and parameters. The main considerations for the dynamic
173 performance are motion responses due to wave actions that include the motion response
174 amplitude operators (RAOs) and the RAOs of cross-sectional loads of the hull in moored
175 condition. While the calculation of structural load effects is mainly associated with the
176 detailed structural design, RAOs of cross-sectional loads might also be useful in the
177 assessment of the global performance of the hull. The RAO analysis for the global motions
178 and cross-sectional loads is carried out in HydroD with a set of regular waves with different
179 periods and propagation directions, and using a simplified mooring system (linear spring).
180 Based on assessing the dynamic characteristics of different hull layouts, a judgment is made
181 to choose the final global design schemes and output the floater's global geometry, mass,
182 and hydrostatic/hydrodynamic properties. Moreover, the cross-sectional load effects related
183 to structural strength might provide a reference to decide on updating the main

4 | 33
184 configuration (especially with/without braces/trusswork) or the global dimensions in the last
185 three Steps.
186 The global design criteria and the sensitivity analysis method are described in greater
187 detail in Sections 2.1 and 2.2, respectively. Reference floating wind turbine system used in a
188 case study is described in Section 3. Results of the case study are presented in Section 4,
189 including e.g., details about serviceability performance, intact stability, and dynamic
190 responses of global motions and internal forces and moments.
191

192
193 Fig. 1. Global design methodology and procedure.
194
195 2.1 Global design criteria
196 Design of the floating support structures should satisfy the serviceability and safety
197 throughout the intended service life. Serviceability refers to the power production and
198 quality, which is closely associated with the hull heeling angle. The safety implies that the
199 annual failure probability of the turbine system during the service life should not exceed the
200 target value, for example, specified in design rules. In practice, this is ensured using
201 appropriate characteristic values of load effects and resistance and suitable safety factors.
202 To realize the requirements of serviceability and safety, the design should first and
203 foremost satisfy the criteria of static heeling angle, stability, and structural integrity. The

5 | 33
204 stability criteria consist of intact stability and damaged stability criteria. The intact stability
205 requirements are mandatory in the hull structure design, while the damaged stability
206 criterion is optional. The intact stability implies that the FWT should be capable of
207 maintaining stability during the operation under the largest rotor thrust loading condition
208 and during the parked under the severe storm condition.
209 The structural integrity will be checked in the detailed design stage to avoid failures in
210 the ultimate limit state (ULS), fatigue limit state (FLS), accidental limit state (ALS), and
211 serviceability limit state (SLS). However, it is not conveniently performed in the global initial
212 design phase. In this paper, design for structural integrity are addressed at a global levelby
213 limiting the global structural load effects by proposing restrictions on the natural periods,
214 and the characteristic frequencies in RAOs of the cross-sectional loads of the hull. The
215 characteristic frequencies of the cross-sectional load effects are related to the phase angles
216 of wave loads on different parts of the semi-submersible hull. In the present work, they
217 correspond to the splitting/prying forces due to the phase angles of the wave loads on the
218 central and outer columns.
219 In the present work, the global design for the FWT hull structure is carried out based on
220 the design requirements of the static heeling angle, natural periods and mean offset of the
221 hull, and the criterion of intact stability. In addition, the characteristic frequencies in RAOs of
222 the hull cross-sectional loads are studied, which closely link to the structural strength
223 performance. Moreover, detailed structural design is not considered. These aspects are
224 addressed as follows. Sections "2.1.1 Serviceability" and "2.1.4 Motion natural period
225 criteria" constitute the "Initial criteria" in Step 2, and Section "2.1.2 Intact stability criteria"
226 corresponds to the criteria in Step 3.
227
228 2.1.1 Serviceability criteria
229 Serviceability limit state (SLS) design of structures includes factors such as durability,
230 deflection/displacement or accelerations that reduce the overall performance of the power
231 production system, cracking of reinforced concrete that can lead to corrosion, and excessive
232 vibration, etc. A detailed study on the SLS assessment of semi-submersible FWTs is
233 conducted by Wang and Moan [23], focusing on the main criteria of floater tilt angle and
234 nacelle accelerations. The floater tilt angle should be assessed from the free-floating and
235 moored as well as static and dynamic conditions. The static analysis in the free-floating
236 condition is closely linked to the intact stability performance, while the dynamic analysis in
237 the moored condition is tightly associated with the power performance.
238 Since the global design study aims at determining a reasonable global layout, a
239 simplified serviceability criterion is considered in the present work.
240 The serviceability is evaluated by the static heeling angle under the maximum mean
241 thrust force of the wind turbine in a free-floating mode, which is expressed as follows:
𝑀
242 𝜃=𝐶 (1)
55
243 where 𝑀 presents the overturning moment, 𝐶55 is the hydrostatic stiffness in pitch (or roll)
244 that is calculated by the Sesam module Wadam [24] based on the following equation:
245 𝐶55 = 𝜌𝑔(𝐼55 + 𝑉𝑤 𝑍𝐵 ) − 𝑚𝑔𝑍𝐺 (2)
246 where 𝜌 is the density of the water; 𝑔 is the acceleration of gravity; 𝐼55 is the waterplane
247 moment of inertia in pitch; 𝑉𝑤 is the volume of the wet part of the hull structure; 𝑚 is the
248 mass of the wind turbine structure (including the mass of rotor, nacelle and tower and the
249 steel mass and ballast mass of the hull structure); 𝑍𝐵 is the vertical center of buoyancy; 𝑍𝐺 is
250 the vertical center of gravity. The hydrostatic stiffness in Eq. (1) can also be obtained

6 | 33
251 considering the nonlinear geometrical effect of the hull. However, a linear assumption as
252 expressed in Eq. (1) often suffices.
253 The overturning moment consists of the contributions of the rotor thrust force and the
254 wind drag force on the tower, as expressed in the following equation:
255 𝑀 = 𝐹𝑇 ∙ 𝐻ℎ + ∑𝑁 𝑖=1 𝐹𝑖 ∙ 𝐻𝑖 (3)
256 Where N represents number of tower section, 𝐹𝑇 is the rotor thrust force, 𝐻ℎ is the vertical
257 distance between the hub and the center of the floatation of the semi-submersible
258 waterplane area. 𝐹𝑖 represents aerodynamic drag loads on the ith section of the tower,
259 which are estimated by the drag term of the Morison formula, where the non-dimensional
260 aerodynamic coefficient is specified as 0.8, which is based on the relationship of the drag
261 coefficients and the Reynolds number summarized in the OC4 report [6]. 𝐻𝑖 presents the
262 vertical distance between the center of the ith section of the tower and the floatation center.
263 The overturning moments would decrease with the increase of heeling angle, and a
264 conservative calculation based on the vertical condition is conducted in the present work.
265 According to common practice in stability analysis, e. g., DNVST-0119 [15], the level arm,
266 when calculating the wind overturning moment, shall be taken as the vertical distance from
267 the center of (drag) pressure of all wind-exposed surfaces to the center of lateral (drag)
268 resistance of the underwater body of the unit or to the level of the mooring line attachment
269 points (fairleads). The consideration of the underwater pressure center is based on the
270 assumption of a freely drifting structure under constant speed. For a moored structure, it is
271 specified (in some standards) that the heeling lever should be conservatively assumed by
272 fairlead. In the present work, the center of the floatation is chosen to calculate the
273 overturning moment, which is more conservative than the alternative method.
274 Table 1 lists the overturning moments induced by rotor thrust and tower drag forces
275 under rated and parked conditions, where the mean wind speeds at hub are 11.4 m/s and
276 14.2 m/s, respectively. It is seen that the thrust-induced overturning moment is dominating
277 in the rated condition, while the contributions of the tower load and thrust force are similar
278 in the parked condition. The total overturning moment under the rated condition is much
279 larger than that under the parked condition; thus, it is used for the static heeling angle and
280 intact stability analyses. It is noted that the contributions of drag force on the floater due to
281 the current on the overturning moment are not considered in the present work. The
282 overturning moments in Table 1 are calculated for the DTU 10-MW reference wind turbine,
283 which are shown in Section 3.1 with more details. It is noted that the thrust force used in
284 this study is based on the land-based turbine, which is larger than that for FWT due to
285 different control systems. Therefore, the SLS results represented by the resultant tilt angle
286 will be conservative. Detailed discussions about the tilt angle estimation by different wind
287 turbine models are shown in the study of Wang and Moan [23].
288
289 Table 1. Overturning moments under rated and parked conditions.

Overturning moment (Nm) Rated condition Parked condition


Rotor thrust force contribution 1.73e8 0.20e8
Tower drag force contribution 0.03e8 0.21e8
Total 1.76e8 0.41e8
290
291 Requiring the heeling angle to be unnecessarily small might result in the overdesign of
292 the hull and excessive costs, while allowing a large heeling angle would, to some extent,
293 reduce power output. The study of Wayman [25] indicated that the FWT will significantly

7 | 33
294 reduce the power efficiency beyond a steady-state pitch angle of 10 degrees. Therefore, a
295 good compromise should be considered by limiting the heeling angle within a reasonable
296 range. A threshold range of the static heeling angle of 5-10 degrees in the free-floating
297 condition is used as the serviceability requirement in the conceptual design based on
298 experiences from the wind industry.
299
300 2.1.2 Intact stability criteria
301 Floating stability implies a stable equilibrium and reflects a total integrity against
302 downflooding and capsizing. The stability requirements cover both static and dynamic
303 stability. According to the DNVGL-ST-0119 [15], stability shall either be demonstrated by a
304 quasi-static evaluation on the moment-heeling curves or by time domain stability evaluation.
305 In the present work, the quasi-static stability evaluation is addressed. In addition, sufficient
306 stability shall be documented for a set of load cases and various wind turbine modes,
307 including normal operating, survival (extreme storms), fault, temporary (installation,
308 inspection, and changing of draught), and transit (tow-out) conditions. In this work, the
309 normal operating and survival conditions are considered.
310 The intact stability is evaluated by the requirements, that 1) GM = C55/ (where C55 is
311 given by Eq. (2) and  is the displacement) shall be larger than 0.3, 1.0 m for temporary
312 phases of short duration and for operation, respectively; 2) the area under the righting
313 moment curve to the second intercepted angle shall be equal to or larger than 130% of the
314 area under the heeling moment curve to the same angle; 3) the righting moment curve shall
315 be positive over the entire range of angles from upright to the second intercept 4)
316 watertight integrity shall be ensured at the first intercept. For the four-column semi-
317 submersible layout, as illustrated in Fig. 2, the righting moment curve varies in different
318 wind directions; thus, the intact stability is evaluated using the righting moment data in the
319 most critical condition.
320 As mentioned above, the approach to determine the overturning moment based on the
321 floatation center or fairleads gives a conservative estimate. However, the quasi-static area
322 requirement used to check the stability is uncertain, and it seems that regulators are
323 cautious to introduce a more physically correct estimate of the overturning moment and
324 hence the determination of the intercepts with the righting moment curve. The fact that
325 the stability criterion, commonly used, is a simplified quasi-static force-heeling area criterion
326 relating to wind loads only, has led to the proposal of a dynamic approach based on time
327 domain simulations where both the effect of waves and wind are considered. However, as
328 mentioned in current codes, e. g., DNV-ST-0119 [15], there is limited documentation of the
329 instability criterion and safety factor for such an approach. Currently, this option does not
330 seem to be viable.
331 The intact stability is closely correlated to the freeboard of the outer columns because
332 the restoring moment will decrease sharply when an outer column is submerged. The
333 righting moment curve proposed in the DNV-ST-0119 [15] is obtained based on the quasi-
334 static calculation in the still water condition, where the angle at the outer column
335 submerging is larger than the condition that the dynamic hull motions and wave elevations
336 are considered. However, the semi-submersible hull might, in principle, not be capsizing due
337 to the inertia effect even if the outer column is submerging. Therefore, the safety of the
338 intact stability should be checked by carrying out a dynamic stability check based on a time
339 domain simulation analysis.
340

8 | 33
341 2.1.3 Damage stability criteria:
342 Damage stability criteria for unmanned wind turbines are not mandatory but optional
343 to the owner of the turbine, and are not pursued in this study. However, it is noted that the
344 present concept involves pontoons which are fully filled with water ballast and damage to
345 the pontoons does not imply flooding of the hull. Moreover, damage stability could be
346 introduced based on a cost-benefit analysis considering the frequency of damage.
347
348 2.1.4 Natural period criteria with respect to rigid-body motions
349 The selection of natural periods of rigid-body motions of the wind turbine system in the
350 design process is closely related to the structural load effects and power performance.
351 Significant motion responses in heave and pitch (or roll) will increase cross-sectional loads of
352 the hull due to the structural inertial loads and the varying gravity and buoyancy-induced
353 bending moments. In addition, especially large pitch motions will affect the power
354 generation and quality. To “remove” the resonant responses of the hull motions in heave
355 and pitch, the natural periods should be outside the range of first-order wave load excitation
356 (4-25s). The natural periods in heave and pitch are estimated by the following equations
357 based on a linear assumption of the restoring effect.
1
𝑀 +𝐴 2
358 𝑇33 = 2 ∗ 𝑝𝑖 ( 33𝐶 33 ) (4)
33
1
𝑀 +𝐴 2
359 𝑇55 = 2 ∗ 𝑝𝑖 ( 55𝐶 55 ) (5)
55
360 where 𝑀33 and 𝑀55 are mass and mass moment of inertia about original y-axis, respectively,
361 of the wind turbine structure (including the mass of the simplified rotor and nacelle, as well
362 as tower, hull structure, ballast); 𝐴33 and 𝐴55 are the added mass in heave and pitch,
363 respectively, of the floating wind turbine at the corresponding resonance frequency, which
364 are calculated in Wadam [24]. 𝐶55 is calculated in Eq. (2). 𝐶33 is the hydrostatic stiffness in
365 heave that is calculated by Wadam [24] based on the following equation:
366 𝐶33 = 𝜌𝑔𝐴𝑤 (6)
367 where 𝐴𝑤 is the water plane area in still water.
368 The limits for the heave and pitch natural periods are considered in the sensitivity study,
369 which are: 𝑇33 ≥ 20𝑠, 𝑇55 ≥ 25𝑠, respectively. There is a compromise for the design
370 requirements of the heave natural period and the heeling angle. To satisfy the serviceability
371 requirement by limiting the static heeling angle within the given range, sufficient pitch
372 stiffness 𝐶55 is needed, while the heave stiffness 𝐶33 will be likely increased simultaneously,
373 posing the risk of heave motion resonances. In this stage, the contribution of the mooring
374 system is not included.
375 To reduce the mean offset and thus avoid large dynamic loads of the power cable, and
376 to reduce the mooring line tensions due to the low-frequency wind and second-order wave
377 excitation, the mooring system is designed to have adequate stiffness and to make the
378 natural periods in the surge, sway and yaw motions within the range of 60-120 s. The six-
379 DOF natural periods of the wind turbine system under the moored condition are given in
380 Section 4.6.
381 According to the discussions about the conceptual design criteria above, the required
382 values for design criteria shown in Fig. 1 are summarized in Table 2. The required design
383 values are used to carry out the case study in the present work, while it is noted that other
384 criteria and assessments in Fig 1, which are not listed in Table 2, are also considered in the
385 conceptual design of the present case study.

9 | 33
386 Table 2. Summary of the design criteria and the required values) in three steps in Fig. 1.

Steps in Fig. 1 Design criteria Required values


Static heeling angle 5-10 deg (free-floating)
Step 2 Heave natural period ≥20 s
Roll/pitch natural period ≥25 s
GM (initial =0) GM  1.0 m for operation
Step 3
Area ratio ≥1.3
Natural periods in surge, sway, yaw in
60-120 s
moored condition
Natural period in heave in moored
Step 4 ≥20 s
condition
Natural period in roll/pitch in moored
≥25 s
condition
387
388
389 2.2 Sensitivity study
390 A sensitivity study of global responses with relation to the performance of the
391 serviceability as well as heave and pitch natural periods is carried out to determine the initial
392 floater geometry. The primary purpose of the sensitivity study is to determine the
393 parameters that are most effective for achieving the initial design criteria, and based on this,
394 further determining a relevant design space. It is considered as a part of the global design
395 procedure and is a simplified design optimization. As illustrated in Fig. 2, the configuration of
396 the floater mainly consists of four columns connected by three pontoons. Therefore, the
397 leading global parameters for the sensitivity study are related to the columns and pontoons.
398 The design variables and variation range used in the sensitivity study are listed in Table 3 and
399 are illustrated in Fig. 2. The coordinate system refers to a right-handed rule, which origin at
400 the central column center in mean water level. The XY-plane coincides with the mean water
401 level and X-axis is positive from one column side to two column side, and Z-axis is positive
402 upwards. Detailed description for the baseline model is shown in Section 3.2.
403

404
405 Fig. 2. Sketch of the baseline model of the hull structure [9].
406
407
408 Table 3: Design variables and variation ranges.

Design variables Unit Baseline value Variation range Interval


Draft m 22 14-30 2
Column centerline spacing m 37 29-45 1
Outer column radius m 6.5 2.5-8 0.5

10 | 33
Outer pontoon radius m 8 8-12 1
Pontoon width m 16 8-16 1
Pontoon height m 7 3-11 1
409
410 The baseline values of the design variables are used based on previous experiences, and
411 in this work, the public study [9] relating to the OO-Star design is referred. As illustrated in
412 Fig. 2, two constraints in the floater dimensions exist in the present layout; namely, 1) the
413 diameter of the outer column should be equal to or smaller than the diameter of the outer
414 pontoon, 2) the diameter of the outer pontoon should be equal or larger than the pontoon
415 width. In the baseline model, the outer pontoon diameter is equal to the pontoon width,
416 which can be seen in Fig. 3. When one design variable is used for the sensitivity study, all the
417 other variables remain the baseline values accordingly. Therefore, the variation values of the
418 outer pontoon radius are equal to or larger than the baseline value of 8 m, and the
419 variations values of the outer column radius and pontoon width are equal to or smaller than
420 8 m and 16 m, respectively.
421 The sensitivity of the global responses to the variation of design variables is studied and
422 reported in Section 4.1. The sensitivity is represented by the sensitivity index, which is
423 defined as:
424 Sensitivity index (variation of responses):
425
∆𝑌𝑖
426 χ= (𝑋𝑖 ) (7)
𝑌𝑖
427
428 where 𝑋𝑖 represents the design variables shown in Table 3; 𝑌𝑖 denotes global responses for
429 the baseline variables; ∆𝑌𝑖 represents difference of the variable responses relative to the
430 baseline responses.
431 Sensitivity indices between different responses, such as the static heeling angle and
432 natural periods, are comparable to identify the sensitivity of the responses to variables. This
433 is because the sensitivity index is expressed along with the relevant change of the design
434 variables, namely, the variation of the design variables is defined in the similar manner:
435
∆𝑋𝑖
436 ϒ= (8)
𝑋𝑖
437 where ∆𝑋𝑖 represents difference of the design variable values relative to the baseline values.
438 In the sensitivity analysis, when one variable changes, all the other variables remain
439 fixed. The global responses in the sensitivity study are obtained by the hydrodynamic
440 analysis using the HydroD code. All the panel, structural, and mass models are updated in
441 each case. Therefore, the ballast distributions and the resultant displacement, added mass
442 and hydrostatic stiffness properties, overall mass properties, as well as CoB and CoG
443 information are updated accordingly. In all the cases, the pontoons are fully filled with the
444 water ballast, and only the proportions of the water ballasts inside the outer columns and
445 outer pontoons vary to achieve the equilibrium between the overall mass and buoyancy.
446
447
448
449
450
451
452

11 | 33
453 3. Description of baseline hull structure, as well as the reference wind turbine
454 and mooring system
455 3.1. 10-MW reference wind turbine
456 The Technical University of Denmark (DTU) 10-MW wind turbine [26] is used in this
457 study for conducting the global design of the floaters. A simplified wind turbine structure is
458 used where the rotor and nacelle are simplified as the mass points located on the tower top.
459 In the present work, the cross-sectional and mass properties of the tower are taken
460 from the OO-Star model [27], where the steel density is considered as 8243 kg/m3 to
461 account for the additional bolts, stiffeners, and flanges. The main properties of the wind
462 turbine model used in this study are listed in Table 4.
463
464 Table 4. Main specifications of the simplified 10-MW wind turbine model.

Parameter Value
Rated power (MW) 10
Cut-in, rated, and cut-out wind speed (m/s) 4, 11.4, 25
Hub height (m) 119.0
Rotor mass (kg) 230717
Rotor CoG (m) (-7.07, 0, 119)
Nacelle mass (kg) 446006
Nacelle CoG (m) (2.69, 0, 118.08)
Tower mass (kg) 1.249E+06
Tower CoG (m) (0, 0, 50.16)
465 CoG: center of gravity.
466
467 3.2. Baseline model of the hull structure
468 The layout of the baseline hull structure concept, as shown in Fig. 2 in Section 2.2, is
469 determined based on an existing 5-MW-CSC steel semi-submersible hull designed by Luan et
470 al. [7] and a 10-MW OO-Star concrete hull designed by Dr. techn. Olav Olsen AS in the LIFES
471 50+ project [9]. The present steel hull consists of three outer columns and one central
472 column; the outer columns and the central column are connected by three pontoons at a
473 low position near the hull's draft. Beams or trusses are not considered in that concept; thus,
474 the structural integrity of the hull depends on the pontoons connecting the columns. In
475 addition, heave plates are attached to the bottom of the pontoons to obtain a suitable heave
476 natural period to avoid resonance. Detailed information for the OO-Star concrete hull is
477 shown in the LIFES 50+ reports [9, 27].
478 Fig. 3 shows the arrangement of the 11 compartments in the hull structure. Table 5
479 shows the percentage of water filling in each compartment for the four layouts used for the
480 case studies in Section 5. Since CM1-CM7 are fully flooded by Water ballast in all the layouts,
481 the ballast weight is concentrated at the bottom of the hull to realize a lower center of
482 gravity to obtain good stability. In addition, the mass distribution does not vary significant
483 along with the hull motions. Besides, the pontoons are located at the low position of the hull,
484 subjected to enormous hydrostatic pressures externally, while filling the ballast water into
485 the pontoons alleviate large pressures on the plate.
486

12 | 33
487
488 Fig. 3. Arrangement of the compartments and water ballast distribution.
489
490 Table 5. Percentage of water filling in each compartment for the four layouts in Table 8.
Compartment Layout 1 Layout 2 Layout 3 Layout 4
CM8 3.01% 8.65% 0.80% 26.92%
CM9 3.51% 9.09% 0.87% 27.23%
CM10 3.51% 9.09% 1.08% 27.13%
491 Note: CM1-CM7 are 100% and CM11 is 0% water-filled for all the layouts.
492
493 The semi-submersible steel hull is designed with the following material properties:
494 density = 7850 [kg/m3]; Yield strength = 235 [MPa]; Young's modulus = 2.11*105 [MPa]; and
495 Poisson's ratio = 0.3. In the global design phase, the detailed internal structure design is
496 unavailable; instead, an equivalent plate thickness is used to account for shell plate, internal
497 stiffness, etc., for a preliminary weight estimate. This study assumes an equivalent plate
498 thickness of 50 mm for the whole hull structure.
499
500 3.3. Mooring system
501 In the present work, the mooring system is considered in Sections 5, "Response
502 amplitude operator analysis," while the results in all the other sections do not include the
503 contributions of the mooring system.
504 In Section 5, three linear springs are used to represent the mooring system. Each spring
505 consists of the pretension, vertical and horizontal stiffness, which are 2.0e6 N, 2.4e4 N/m,
506 and 1.9e5 N/m, respectively, as estimated from the catenary mooring system. In addition,
507 each spring connect the outer column of the hull at the fairlead position with an angle of 27
508 deg from the sea surface.
509
510 4. Results and discussions
511 4.1 Sensitivity study
512 In this section, the sensitivity of the heave and pitch natural periods (𝑇33 and 𝑇55 ), static
513 heeling angle (𝜃) are studied to the variation of design variables, including the draft, column
514 centerline spacing, outer column radius, outer pontoon radius, pontoon width, and pontoon
515 height. The intention is to gain good insights into the global response characteristics of the
516 hull. The responses are calculated based on the hydrostatic and hydrodynamic analysis in
517 Wadam [24], where the coupling between different DOFs and the mooring system is not
518 considered. The sensitivity results are illustrated in Fig. 4. Sensitivities of important design
519 variables of the column centerline spacing and outer column radius are shown in Appendix A.
520 All three global responses are very sensitive to the variations of the column spacing and
521 outer column radius. For a given parameter variation range from -0.2 to 0.2, the static

13 | 33
522 heeling angle shows the most significant sensitivity, followed by the pitch natural periods,
523 and the sensitivity of the heave natural period is the smallest.
524 Compared to the column spacing and outer column radius, the sensitivities of the static
525 heeling angle and pitch natural period to the variation of other variables are negligible. Both
526 the static heeling angle and pitch natural period decrease significantly as the column spacing
527 or the outer column radius increases, especially in the parameter variation range from -0.2
528 to 0.
529 Different sensitivity characteristics are observed in the heave natural period. In addition
530 to the design variables of the column spacing and outer column radius, the heave natural
531 period is also very sensitive to the variation of the pontoon width. Moreover, the heave
532 natural period decreases as the outer column radius increases, while it increases for
533 increasing column spacing or pontoon width, and the pontoon width results in a more
534 significant sensitivity.
535

536
537

538
539 Fig. 4. Comparison of sensitivities of the heave and pitch natural periods and static heeling angle of
540 the hull structure to the variation of different design variables.
541
542 Overall, two effective design measures will satisfy the initial design criteria: 1) large
543 column spacing and small outer column radius; 2) small column spacing and large outer
544 column radius. In these two design measures, a good compromise will be achieved for the
545 performance of the static heeling angle and pitch natural period. However, the dynamic

14 | 33
546 performance of the heave natural period will be different. Measure 1) will result in large
547 heave natural periods, implying that the resonant response could be avoided, while measure
548 2) will imply a risk of heave motion resonance due to the decreasing heave natural period.
549 Increasing the pontoon width will be an effective way to remedy the heave resonance risk in
550 measure 2), because the heave natural period increases significantly as the pontoon width
551 increases.
552
553 4.2 Response analysis of two alternative design measures
554 The sensitivity study in last section shows that the two dominating parameters for
555 global design are column spacing and outer column radius, and summarizes two effective
556 design measures; namely, considering large column spacing with small outer column radius,
557 or vice versa, while keeping other parameters as constant. In this section, detailed
558 responses of the heave and pitch natural periods and static heeling angle of the alternative
559 design measures are studied, which aims at determining the initial design space. In each
560 design measure, 48 design options are obtained by only changing the column spacing and
561 outer column radius, while all the other parameters are kept at the baseline values shown in
562 Table 3. In design measure 1), the column spacing varies from 38m to 45m with an interval
563 of 1m, and the outer column radius varies from 5m to 6.25m with an interval of 0.25m.
564 Similarly, the design measure 2) is formed by a combination of eight options for the column
565 spacing varying from 29m to 36m with an interval of 1 m and six options for the outer
566 column radius varying from 6.75m to 8m with an interval of 0.25m. A detailed discussion of
567 the global responses is presented in the following.
568 Figs. 5 and 6 illustrate the contour lines of the global responses of hull structure under
569 different combinations of the column spacing and outer column radius in design measures 1
570 and 2, respectively.
571 In design measure 1, all the heave and pitch natural periods exceed the 20s and 30s,
572 respectively, which implies that all the design options satisfy the requirement of the heave
573 and pitch motion natural periods. The static heeling angles that satisfy the initial design
574 criterion are located in the area marked by the red circle in Fig. 5, which covers a broad
575 range of design options with various combinations of column spacing and the outer column
576 radius. In design measure 2, as illustrated in Fig. 6, all the pitch natural periods are larger
577 than 25s, satisfying the design requirement. However, most of the heave natural periods are
578 lower than the required threshold value of the 20s. A few design options with a natural
579 period close to the 20s are located in the area with a small outer column radius and large
580 column spacing. Although there is a broad range of design options that satisfy the criterion
581 of the static heeling angle, as marked in the red circle, the compromise with the
582 considerations of the heave natural period limits the final design option. However, the static
583 heeling angle characteristic is most important and will be prioritized when choosing the
584 design parameters.
585

15 | 33
586

587
588 Fig. 5. Global responses of heave and pitch natural periods and static heeling angle of the hull
589 structure under different combinations of the column spacing and outer column radius in the design
590 measure 1.
591

592

16 | 33
593
594 Fig. 6. Global responses of heave and pitch natural periods and static heeling angle of the hull
595 structure under different combinations of the column spacing and outer column radius in the design
596 measure 2.
597
598 4.3 Intact stability analysis
599 In this section, intact stability analysis of semi-submersible hulls is conducted. Based on
600 the sensitivity study in Section 4.2, two representative design layouts are selected in the
601 design measures 1 and 2, respectively, for the following intact stability and response
602 amplitude operator (RAO) studies. The two layouts correspond to the layouts 1 and 4 in
603 Table 9 below. By referring to the baseline dimensions in Table 3, only the column centerline
604 spacing and outer column radius are different in the two design layouts, while other
605 parameters remain the same.
606 The righting moment differs in different wind directions relative to the hull. Fig. 7
607 illustrates the top view of the hull layout of design layout 1 and the seven different incoming
608 wind and wave directions from 0° to 60°, which covers all the heeling conditions due to the
609 symmetry of the structure. It is, for instance, noted that the wind direction 60° gives the
610 same results as the wind direction 180°.
611
612
613

17 | 33
614
615 Fig. 7. Top view of the hull layout of design layout 1 and the incoming wind/wave directions.
616
617 Fig. 8 shows the righting moment curves of the hull in the seven wind directions and the
618 constant overturning moment curve, for the intact stability analysis of design layout 1.
619 Significant differences are found in the righting moment curves when the heeling angle is
620 large. The most critical condition for the intact stability check is the 60° wind direction for
621 the specific hull layout. It is noted that the nonlinearity for directions of 0°-40° is due to the
622 transition of the waterplane area from the columns to include pontoons.
623 The waterplane moment of inertia in pitch, 𝐼55 , is found to be the main cause to C55 in
624 Eq. (2) for semi-submersible FWTs. Three important conditions are shown in Figs. 8-9.
625 Conditions 1 and 2 are selected for the wind direction of 0°, and condition 3 is selected for
626 the 60° wind direction. Condition 1 refers to the transition position from slow line increase
627 to rapid quick increase of the righting moment, which is due to the transition and increase of
628 the waterplane area from the outer column to the pontoon end. This is because, as shown in
629 Fig. 9 (a), the pontoon end starts to come out of the water, but the columns on the right
630 have not been submerged. Condition 2 refers to the maximum righting moment reached in
631 the approximately linear increase phase, and the righting moment curve becomes nonlinear
632 after the heeling angle of this condition. As shown in Fig. 9 (b), in condition 2, the outer
633 columns on the right side are submerged; meanwhile, the pontoon on the left side has
634 already come out of the water. Condition 3 refers to the peak position of the righting
635 moment curve. As shown in Fig. 9 (c), in this condition, the outer column on the left side is
636 submerged while the outer columns on the right side are still watertight.
637 Similar characteristics for the wind direction in the range between 10° and 50° can be
638 inferred from the situations in the 0° and 60° wind directions.

18 | 33
639
640 Fig. 8. Intact stability analysis of the design layout 1 (righting moment curves (RMC) for various wind
641 directions vs. overturning moment curve (OMC)).
642

643
644 Fig. 9. Illustration of the situations of floating conditions 1-3 pointed out in Figure 10.
645
646 Fig. 10 shows the variation of GM with the heeling angle for wind directions 0° and 60°
647 of the design layout 1. In wind directions 0° and 60°, the GM is larger than the requirement
648 of 1.0 m at the initial heeling angle (a requirement applied for the initial GM for semi-
649 submersible MODUs. It is currently unclear what will be the requirements for floating wind
650 turbines. The unclear situation is partly because IEC standard 61400-3-2 for floating wind
651 turbines in the current draft version refers to the IMO standard for MODUs).

652
653 Fig. 10. GM under the 0° and 60° wind directions for the design layout 1.
654

19 | 33
655 Table 6 shows the first intersect angles and the area ratios between the righting and
656 overturning moment curves to the second intercept angle for the seven wind direction
657 conditions. The first intersect angles are the same in all seven cases, and the value is very
658 close to the static heeling angle θ. However, the area ratio varies significantly with the wind
659 direction, and the values of the 40°, 50°, and 60° heeling cases do not satisfy the ratio
660 requirement of 1.3 in the intact stability criterion.
661
662 Table 6: The first intersected angles and area ratios between the righting and overturning moment
663 curves for different wind direction conditions.

Direction (deg) 0 10 20 30 40 50 60
First
8.14 8.14 8.14 8.14 8.14 8.14 8.14
intersect (deg)
Area ratio 2.95 2.84 2.49 1.91 1.15 1.05 1.02
664 Note: the area ratio is defined as the ratio of the righting and overturning moment curves to the
665 second intercepted angle.
666
667 The intact stability of design layout 1 can be improved by increasing the outer column
668 radius (as indicated above) or the freeboard. The improvements in the intact stability based
669 on these two measures are illustrated in Figs. 11.
670 In the analysis, except for the outer column radius and the freeboard, all the other
671 parameters remain unchanged. Accordingly, buoyancy, mass, and hydrodynamic properties
672 are updated with the panel, mass, and structural models in HydroD analysis. For different
673 outer column radius conditions, the angles for submergence of different outer column radii
674 are the same because of the consistent freeboard and column spacing. However, the peaks
675 increase for increasing outer column radius due to the increase of the hydrostatic stiffness
676 (slope), which results in the increase of the area under the righting moment curve, and,
677 therefore, better intact stability. The righting moment slopes are approximately consistent
678 between the five freeboard cases from the upright to the submerged conditions, which is
679 because those different freeboards have a minimal influence on the vertical center of gravity
680 and, thus, the hydrostatic stiffness. The submergence angles increase as the freeboard
681 increases, which leads to increased peak values and, thereby, the areas under the righting
682 moment curves.
683
684 Tables 7 lists the first intercept angles and the area ratios of the four outer column
685 radius and five freeboard cases. The first intersect angle decreases as the outer column
686 radius increases because of the increasing hydrostatic stiffness, and conversely, the area
687 ratio increases significantly. The scheme satisfies the intact stability criterion when the outer
688 column radius is 5.6 m. On the other hand, the vertical center of gravity is slightly increased
689 as the freeboard increases, which leads to a minor increase of the first intercept angle.
690 Simultaneously, the area ratio increases significantly as the freeboard increases, and the hull
691 satisfies the intact stability criterion when the freeboard is increased to 15 m.
692

20 | 33
693
694 (a) Four outer column radius conditions (b) Five freeboard conditions
695 Fig. 11. Comparison of the righting moment curves for (a) different outer column radii and (b)
696 different freeboards based on the design layout 1 in the case of 60° wind direction.
697
698 Table 7: The first intersect angles and area ratios between the righting and overturning moment
699 curves for different parameters of the outer column radius and the freeboard.

Variable Parameter (m) First intersection (deg) Area ratio


5.25 8.14 1.02
5.4 7.24 1.17
Out column radius
5.5 6.73 1.29
5.6 6.28 1.35
11 8.14 1.02
12 8.21 1.11
Freeboard 13 8.27 1.19
14 8.35 1.27
15 8.43 1.34
700 Note: the area ratio is defined as the ratio of the righting and overturning moment curves to the
701 second intercepted angle.
702
703 When the outer column radius increases from 5.25 m to 5.6 m, or the freeboard
704 increases from 11 m to 15 m, and all the other parameters are kept fixed for the design
705 layout 1, the area ratio requirement of 1.3 is practically fulfilled for both. The corresponding
706 total surface area of the two floaters increased by 207.32 m2 and 395.84 m2, respectively.
707 This implies that increasing the freeboard requires almost twice as much material as
708 increasing the outer radius. This provides some reference for the design layout selection.
709 However, it should be noted that the stress levels of the two floaters might be different.
710 Therefore, accurate evaluation for design selection should be based on a detailed structural
711 analysis.
712
713 5 Global aspects of structural design
714 In the structural design of pontoons and columns of the floater, ultimate strength (ULS)
715 and fatigue (FLS) design criteria need to be satisfied. While ULS depends on extreme loads,
716 fatigue is often dominated by frequent, more moderate load conditions. ULS criteria have
717 large influence on the steel weight while fatigue criteria affect the local geometry in hot spot
718 areas. In the present paper on initial global design, some considerations about the
719 relationship between global layout and extreme load effects in the pontoons and columns

21 | 33
720 are made. While both global cross-sectional forces/moments and lateral pressure on the
721 pontoons and columns affect the design, the focus herein is on the global effects. Both wind
722 and wave loads play a role. While the global wind-induced cross-sectional loads only
723 essentially depend on the rotor size and hub height, wave-induced load effects depend on
724 the global layout of the floater. Hence, the focus herein is on wave -induced loads and their
725 effects and especially their dependence on the global geometry.
726 Rigid motion resonance of the turbine under wave loads is important for the structural
727 load effects, which again depend on stiffness and mass as mentioned in Section 2.1.4, with
728 the cross-section area of the side columns and distance between central and side columns
729 as main parameters. Wave excitation forces of importance for the structural response,
730 depend on the horizontal distance between columns, since opposing loads on columns
731 cause internal forces. This distance is also directly connected to the length of pontoons.
732 Moreover, motion-induced inertia loads play an important role. It is noted that inertia
733 effects (accelerations) increase with increasing frequency for a given wave height.
734 In Section 5.1, the response amplitude operators (RAOs) for motions and internal
735 forces/moments are presented and discussed. In Section 5.2, example environmental
736 conditions are identified, based on different approaches, and used together with the RAOs
737 to obtain an approximate estimate of extreme load effects in alternative layouts as a basis
738 for possibly reduce the load effects.
739
740 5.1 Response Amplitude Operator (RAO) analysis for wave-induced motions and cross-
741 sectional load
742 This section deals with the responses amplitude operators (RAOs) of the wave-induced
743 motions and cross-sectional bending moments of different hull layouts. Four layouts are
744 selected based on the analysis in Sections 4.2 and 4.3, to investigate the dynamic
745 characteristics of the semi-submersible hull structures. Table 8 shows the main global
746 parameters and response characteristics of the four hull layouts. The simplified wind turbine
747 models for the dynamic RAO analysis are shown in Fig. 12, where layouts 2 and 4 are shown
748 as examples. It is noted that cross-sections 1, 2, and 3 refer to the section near the central
749 column, in the middle and the one near the outer column, respectively. The four models are
750 developed in HydroD (WADAM), where the same wind turbine structure (tower and
751 simplified rotor and nacelle, as described in Section 3.1) and simplified mooring system
752 (linear springs, as described in Section 3.3), are used. Moreover, the four floaters are
753 assumed to have the same equivalent plate thickness of 50 mm and the same compartment
754 arrangement. However, the panel, mass, and structural models are different due to the
755 different floater geometries; therefore, the displacement, ballast, mass, and hydrodynamic
756 properties differ.
757
758
759
760
761
762
763
764
765
766

22 | 33
767 Table 8: Main global dimensions and characteristic behavior of the four selected hull layouts1.

Layouts Layout1 Layout2 Layout3 Layout4


Draft (m) 22 22 22 22
Column centerline spacing (m) 45 45 45 35
Outer column radius (m) 5.25 5.6 5.25 6.8
Freeboard (m) 11 11 15 11
Outer pontoon radius (m) 8 8 8 8
Pontoon width (m) 16 16 16 16
Pontoon height (m) 7 7 7 7
T33 (s) 25.33 24.29 25.33 18.94
T55 (s) 38.26 33.68 38.96 32.03
θ (deg) 7.87 6.05 8.16 7.29
First intersect angle (deg) 8.14 6.28 8.43 7.82
Area ratio 1.02 1.35 1.34 1.35
768 1: T33 and T55 represent the heave and pitch natural periods, respectively; θ represents the static
769 heeling angle; M33 and M55 denote the displacement and the mass moment of inertia about the y-
770 axis, respectively; A33 and A55 represent the added mass in heave and pitch, respectively; C33 and C55
771 denote the restoring stiffness in heave and pitch, respectively.
772

773
774 (a) layout 2 (b) layout 4
775 Fig. 12. Simplified wind turbine models for the hydrodynamic analysis.
776
777 Typically, the RAOs consist of curves with peaks and valleys, sometimes a zero value.
778 Understanding the reasons for this behavior is an important background for the design.
779 Typically, peaks in motion RAOs are caused by resonances or by global load distribution in
780 the regular wave that maximizes the total excitation force. Valleys (even zeros) are caused by
781 forces acting in opposite directions and causing cancellation. As an example, consider the
782 heave motion in Fig. 13. The heave resonance frequency (0.26 rad/s) results in a peak. The
783 zero response most likely can be explained by considering the excitation of the pontoons as
784 follows: assume that the bottom of the pontoon has a surface of A b and the top of the
785 pontoon Au < Ab because of the columns. The dynamic pressure in waves decreases with the
786 distance from the surface, depending on wave frequency. The pressure on the pontoon
787 bottom and top are denoted pb and pu > pb, respectively. The excitation force (and motion) is
788 zero if Au pu = Ab pb, as indicated at a frequency of about 0.3 rad/s in Fig. 13. There is a peak

23 | 33
789 at 0.39 rad/s. The reason for this is that the RAO for heave motion generally increases with
790 reduced frequency, towards the value of 1.0, but then in the vicinity of a natural frequency, a
791 large motion value is caused by the resonance phenomenon. In addition, these three
792 phenomena act together in the range of frequencies from resonance and upwards the
793 particular phenomenon that causes zero excitation. When the peak value for frequencies
794 above 0.26 rad/s is not closer to 1.0, it is due to the fact that the pressure resultant on the
795 pontoons is zero.
796 Similar considerations can be made for pitch motions.
797 Peaks in structural responses (say, the bending moment in pontoons) are caused by
798 resonant motions or by load distribution in regular waves, including the direct load as well as
799 inertia forces especially due to the heave and pitch motions (accelerations). The response in
800 the pontoons depends on the direct wave loads and the inertia effects due to motions, not
801 only on the pontoons but also the columns since columns and pontoons are integrated. In
802 particular, a wave condition that causes opposing forces on the different columns (so-called
803 splitting/prying load conditions), causes large internal forces/moments in the structure. In
804 particular, the resultant horizontal force might be zero with correspondingly, a zero
805 horizontal motion. The frequency range for waves causing splitting/prying conditions for the
806 present structure is about 0.7-0.9 rad/s. The phenomena that affect the structural responses
807 act together, however, sometimes with a phase difference. Hence, it is difficult to assign the
808 reason for the peaks in the RAOs. For instance, peaks in the (pitch) accelerations occur at
809 similar frequencies as the splitting/prying forces on the columns, but the pitch acceleration
810 might be 90 degrees out of phase with the splitting/prying forces. The following discussion is
811 therefore meant to be indicative explanations of the RAOs.
812 Fig. 13 shows the RAOs of the heave and pitch motions as well as the internal bending
813 moments in cross-section 2 for layouts 1, 2, and 3. RAOs with wave periods in the range of 4-
814 25 s are calculated since wave energy is in the period range. In addition, two incoming wave
815 directions of 0° and 180° are considered because the most significant cross-sectional loads
816 might occur in these two wave directions according to the study [28].
817 In the present work, a 5% critical damping is included to account for viscous damping in
818 the six-DOF motions. In addition to the heave natural frequency, other frequencies where
819 the motion peaks occur are inconsistent with those of the cross-sectional bending moment
820 peaks, which implies that the dynamic motion behavior cannot fully reflect the
821 characteristics of structural load effects.
822
823
824
825

24 | 33
826
827 Fig. 13. RAOs of the heave and pitch motions as well as the bending moments in the cross-section 2
828 for layouts 1, 2, and 3.
829
830 Figs. 14 (a) and (b) show the RAOs of the heave and pitch motions and accelerations,
831 respectively, for layouts 2 and 4. In the heave motion RAOs, significant peaks appear at the
832 0.26 rad/s and 0.33 rad/s of layouts 2 and 4, respectively, which correspond to heave
833 resonance. Significant pitch motion responses occur at the low wave frequencies (close to
834 0.25 rad/s) because they are close to the pitch motion natural frequencies. In general, the
835 heave motion of layouts 2 and 4 is very close, while the pitch motion of layout 4 is
836 significantly larger than that of layout 2 for low frequencies (close to 0.25 rad/s) while it is
837 lower in the frequency range of 0.4 – 0.8 rad/s.
838 As shown in Fig. 14 (b), the RAOs of the hull accelerations partly reflect motion
839 resonance but also the fact that acceleration level for a unit wave amplitude, which increase
840 with increasing frequency until about 0.5-0.8 rad/s, especially in the pitch acceleration RAOs.
841

25 | 33
842
843 (a) motion RAOs
844

845
846 (b) acceleration RAOs
847 Fig. 14. RAOs of the heave and pitch motions and accelerations for layouts 2and 4.
848
849 The RAOs for the pontoon bending moments for layouts 2 and 4 are shown in Fig. 15. It
850 is found that the RAO amplitudes between the two layouts are close, while the locations of
851 peak values in layout 2 are generally lower than in layout 4. The RAO characteristics for 0°
852 and 180° wave directions are very different for the two layouts, as shown in the plots.
853 Critical frequencies in Fig. 15 and corresponding response characteristics are summarized in
854 Table 9. The results in Fig. 15 and Table 9 imply that heave motion resonance, inertia loads in

26 | 33
855 heave and pitch, and splitting/prying force on columns significantly affect the cross-sectional
856 bending moments.
857 In the 0° and 180° wave directions, the bending moments are influenced by the inertia
858 loads in heave and pitch, respectively. In addition, the low-frequency heave and pitch
859 motions also affect the structural internal bending moments due to the variations of the
860 gravity- and buoyancy-induced loads.
861 In addition, the cross-sectional bending moments at the frequencies of 0.83 rad/s in
862 layout 2 and 0.91 rad/s in layout 4 are caused by the splitting/prying force at a critical phase
863 of column loads. Critical responses will occur at the pontoon cross sections at a wavelength
864 of approximately twice the distance between the central and outer columns. More detailed
865 explanations can be found in the standard DNVGL-RP-C103 [29].
866

867
868 (a) cross-section1 (b) cross-section 2 (c) cross-section 3
869 Fig. 15. RAOs of the cross-sectional bending moments for layouts 2 and 4.
870
871 Table 9: A summary of frequencies and the corresponding dynamic response characteristics that
872 appear in Fig. 15.
Frequency Wave
Layout Dynamic characteristics
(rad/s) direction
0.26 180° Heave resonance
0.77 180° Inertial loads in pitch
Layout 2
0.83 0° Splitting/prying force on columns (wavelength: 90m)
1.08 0° Inertial loads in heave
0.33 0°, 180° Heave resonance
0.83 180° Inertial loads in pitch
Layout 4
0.91 0° Splitting/prying force on columns (wavelength: 70m)
1.21 0° Inertial loads in heave
873
874
875

27 | 33
876 5.2 Preliminary comparison of global structural integrity between alternative layouts.
877 As mentioned above, the wind loads depend on the rotor size and control and is the
878 same for the layouts considered herein. Moreover, these features are decided for other
879 reasons, such as serviceability and intact stability, than consideration of structural integrity.
880 In the following, the effect of the layout of the hull on global wave-induced load effects and
881 hence structural integrity, is addressed, based on selected 50-year sea states and the use of
882 the RAO’s (for the bending moment in the pontoon) in Section 5.1.
883 Typical environmental conditions for floating wind turbines can be represented by the 5
884 sites considered by the study [30]. In this connection, the severe conditions at site 14 are
885 considered as an example. The environmental 50-year contour surface is shown in Fig.16.
886 The most important environmental conditions refer to conditions with rated wind speed
887 (11.4 m/s at the hub), cut-out wind speed (25 m/s at the hub, but 24 m/s is applied since this
888 is more critical due to the effect of the control system [31]), and parked condition in extreme
889 environmental condition, corresponding to a 50-year condition.
890 The standard IEC 61400-3 [32] (and the forthcoming IEC 61400-3-2) recommends a
891 simplified method for identifying the load cases. The 50-year values of Hs for given wind
892 speed are determined based on a simplified inverse first order reliability (IFORM) method
893 and a range for Tp is assumed. The Weibull distribution is assumed for the marginal
894 distribution of the mean wind velocity and the conditional distribution of Hs for a given wind
895 velocity for the site 14 (Li et al., 2015). Then, for 10 min wind speeds at the hub of 11.4 m/s
896 and 24 m/s (corresponding to 7.86 m/s and 16.55 m/ at 10 m level and 1 hour average), the
897 conditional Hs is given in the Table 10.
898 A more refined approach is to determine the contour line for Hs-Tp for a given Uw. Fig.
899 17 shows the 50-year environmental contour surface obtained based on the joint
900 distribution consisting of marginal distribution of Uw, conditional distribution of Hs for given
901 Uw, and conditional distribution of Tp for given both Uw and Hs. Fig. 18 shows the contour
902 lines of Hs-Tp for given mean wind speed for rated and cut-out conditions. Some of the
903 combinations are indicated in the Table 10.
904

905
906 Fig. 16. 50-year 3D environmental contour surface (complete method) [30].

907

28 | 33
908
909 Fig. 17. Fifty-year rated and cut-out conditions contour lines for site 14 (complete method) [30].

910
911 Table 10: Conditional values of Hs and Tp for given 10 min mean wind speeds at the hub.
Wind speed IEC 61400-3 approach to estimate Contour based on joint Hs, Tp
Expected
(10 min average, 50-year joint extreme in terms of (Hs, Tp) pairs
Hs (m)
at the hub (m/s)) Hs Tp (Fig. 18).
(4.8 m, 7 s), (5.4 m, 8 s), (6.4 m,
11.4 2.46 m 10.15 m1) Assumed range2) 10 s),
(7.5 m, 15 s), (6.5 m, 20 s)
(6.1 m, 7 s), (7.0 m, 8 s), (8.5 m,
24.0 4.87 m 9.55 m1) Assumed range2) 10 s),
(10.5 m, 15 s), (6.8 m, 20 s)
912 1) Using the IFORM based on Weibull distribution of conditional Hs, given Uw.
913 2) Assumed Tp range of, say, 5-10 s – i. e., for the given wave heights.
914
915 The design value is obtained as the maximum of the responses for all combination of Uw,
916 Hs, Tp. Based on RAOs of the cross-sectional loads shown in Fig. 15, it is seen that Tp in the
917 range of 5-10 s and above 20 s are critical because peaks are experienced in the ranges. This
918 fact underlines the importance of analyzing structural responses with environmental
919 conditions that Tp is in the range of 5-10 s. In addition, the environmental conditions with
920 largest Hs on the contour lines are important for structural response assessment. Another
921 important consideration is to avoid resonance by requiring the natural heave period to be
922 larger than, say, 20 s.
923 It is noted that in the case study, the simplified approach in IEC 61400-3 [32] by using a
924 constant Hs that correspond to a 50-year Uw-Hs and Tp in the range of 5-10s, is conservative.
925 This is because RAO peaks are located in this Tp range where the Hs are considerably smaller
926 than the constant Hs, as seen in the contour lines in Fig. 17. However, the approaches of
927 determining environmental conditions for structural response assessment should be
928 compared for more case studies with different RAOs. Another interesting observation is that
929 the expected Hs are significantly smaller than the Hs in contour lines, as shown in Table 10,
930 which implies that considering the most probable or expected Hs for extreme value analysis
931 for structural ULS design check is not sufficient.
932 First, the load effects for the rated and cut-out wind conditions, may be compared. Since
933 the Hs for the cut-out condition is 25-35 % larger than for the rated condition, the load
934 effects will be correspondingly larger. However, the wind loads are significantly larger for the
935 rated condition, so both conditions need to be considered in detailed design. In the present

29 | 33
936 context the main aim is to assess the possible influence of the global hull geometry on the
937 global structural load effects and, hence, the structural integrity, which are mainly associated
938 with wave loads. This comparison will be done with reference to the rated condition, based
939 on the relevant (Hs, Tp) pairs (Fig.17, Table10) and the RAOs in Fig.15. In the following
940 simplified comparisons, Hs is used as a representative wave height corresponding to the
941 spectral peak frequency (2/Tp), and the “extreme” load effect is the product of the RAO
942 and Hs at a given frequency.
943 Then, considering the RAOs for direction 0, it is seen that two frequency ranges are of
944 interest with respect to extreme load effects, namely 0.8-1.3 rad/s and 0.26-0.33 rad/s
945 (resonance). Noting that Hs increases monotonically with increasing Tp (decreasing
946 frequency) – up to Tp = 15 s, from about 3.4 m at 1.2 rad/s to 5.3 m at 0.8 rad/s, it appears
947 that there is a limited difference in the “extreme” load effect between the two layouts. For
948 the relatively large RAOs in the “resonance” range, the wave height is about 6.7 m (at 0.33
949 rad/s) and 5.0 m (for 0.26 rad/s). The RAO at 0.33 rad/s for layout 4 is slightly smaller than
950 that at 0.26rad/s for layout 2. Hence, the resultant load effectsforthe two layoutsare so high
951 and comparable. Similar comparisons can be made for the wave direction 180. For the
952 frequency range 0.8-1.3 rad/s, layout 2 seems to have slightly smaller global load effects
953 than layout 4. The performance in the resonance range is similar to the case with waves in
954 the 0 direction.
955 Similar analysis method can be made for the column design, which is shown in detail in
956 the analyses by Wang et al. [33]..
957 In addition to the wave loads considered in this section, the local lateral pressure, static
958 loads in the still water, and dynamic load effects under the low-frequency wind and second-
959 order wave excitation should be included. It particular, it is noted that the wind induced
960 response is large for the rated wind speed. Thus, a fully coupled dynamic analysis is
961 necessary to capture the time domain cross-sectional loads and local pressure loads for the
962 structural strength check. These features are further addressed in the detailed structural
963 design of the FWT semi-submersible hull [28, 34].
964
965 6. Concluding remarks
966 This study outlines a global design methodology and procedure for semi-submersible
967 hull structures of floating wind turbines (FWTs) (see Fig. 1), and presents an application of
968 this methodology to a 10-MW FWT. Global design criteria are introduced in detail, including
969 serviceability, intact stability, and motion natural period. Since floating wind turbines are
970 relatively novel, there are limited operational experiences, and the design standards are still
971 not mature. For this reason, criteria and methods for assessing serviceability, notably tilt and
972 intact stability, are discussed.
973 The hull structure layout consists of four columns connected by pontoons, and the wind
974 turbine is mounted on the central column. A comprehensive sensitivity study is conducted
975 with the main performance features of heave and pitch natural periods and static heeling
976 angle of the hulls under maximum mean wind turbine thrust loads, which aims to identify
977 the important design variables and useful design measures. Intact stability is studied
978 considering various wind directions, different hull layouts (differs in the combination of
979 outer column radius and centerline spacing), and different hull parameters (outer column
980 radius and freeboard). Global aspects of structural design based on wave-induced RAOs of
981 the relevant load effects and representative environmental conditions for an offshore site,
982 are studied. The main findings are summarized as follows:

30 | 33
983
984 • The column centerline spacing and outer column radius are the design variables that
985 have the most significant influence on the global response performance in terms of
986 the heave and pitch natural periods and the static heeling angle. Increasing the
987 column centerline spacing while decreasing the outer column radius, as well as
988 decreasing the column centerline spacing while increasing the outer column radius
989 are two effective measures to reasonably determine attractive floater global
990 dimensions. Additional attention should be paid to the heave resonance risk in the
991 latter measure. Increasing the pontoon width is an effective way to mitigate the
992 heave resonance risk due to the high sensitivity.
993
994 • For the typical four-column hull layout presented in this study, the static heeling
995 angles under the maximum mean wind turbine thrust are approximately the same in
996 different wind directions. In contrast, the utilization factor of the intact stability, i.e.,
997 the ratio between areas under the righting and heeling moment curves, differs
998 significantly. Increasing the outer column radius and freeboard are two effective
999 ways to increase the intact stability of the hulls.
1000
1001 • It is documented that the mooring system actually have a large influence both on
1002 the heeling moment (lever) and the righting moment through the restoring effect,
1003 depending on the location of the mooring fairleads and the angle for the mean
1004 water line. However, the intact stability criteria (notably by the so-called are
1005 criterion) are subjected to significant uncertainties, among other things in view of
1006 the fact that besides the wind forces that are considered, there is an influence by
1007 the waves. On this background, it is understandable that the area requirement is
1008 based on a conservative assessment by neglecting the effect of the mooring system.
1009 However, the more physically correct analysis method by including the effect of the
1010 mooring system in the determination of the heeling and righting moment should be
1011 considered, when dealing with the serviceability requirement to the steady tilt angle.
1012
1013 • The cross-sectional loads of the hull are influenced by motions and accelerations in
1014 heave and pitch, and the column spacing relative to the wavelength (which
1015 influences the distributed wave loads on different parts and their resulting
1016 integration). Heave and pitch motion resonance will significantly increase the
1017 structural load effects; thus, an appropriate global design is essential to make the
1018 heave and pitch natural periods outside the range of significant wave energy. In
1019 addition, the critical wave periods that cause significant splitting/prying force on the
1020 hull columns should be considered in the structural design. In the present case study,
1021 simplified IFORM method to determine Hs based on the standard IEC 61400-3 will
1022 lead to conservative structural responses compared to the refined environmental
1023 contour line approach.
1024
1025 The global design methodology and procedure illustrated in the present work can
1026 significantly improve the design efficiency in the engineering practice and thus shorten the
1027 whole design cycle. However, the dynamic characteristics of the cross-sectional loads are
1028 studied under the first-order wave excitation in the frequency-domain, and it is of great
1029 interest to investigate the structural load effects under the wind and wave excitation by

31 | 33
1030 using the time-domain approach. In addition, this study only deals with the global design of
1031 the hulls, and the detailed structural design will be treated in another paper considering
1032 global and local, static and dynamic load effects.
1033
1034 Acknowledgment
1035 The authors would like to acknowledge the financial support from Key Laboratory of Far-
1036 shore Wind Power Technology of Zhejiang Province, Hangzhou Zhejiang China, 311122 and
1037 PowerChina HuaDong Engineering Corporation Limited, Hangzhou Zhejiang China,
1038 311122(No. 90814100/642005). Moreover, the second, third, and fourth authors would like
1039 to thank the support from the Research Council of Norway through the Centre for
1040 Autonomous Marine Operations and Systems (AMOS, Project No. 223254), Norwegian
1041 University of Science and Technology. In addition, the third author would like to
1042 acknowledge the support from Faculty of Maritime and Transportation, Ningbo University,
1043 Ningbo, China.
1044
1045 Reference
1046 [1] EWEC, Floating offshore wind-a global opportunity, 2022.
1047 [2] IEA, Offshore Wind Outlook. https://www.iea.org/reports/offshore-wind-outlook-2019., 2019.
1048 [3] M. Walter, S. Paul, B. Philipp, D. Patrick, M. Melinda, A. Cooperman, R. Hammond, M. Shields,
1049 Offshore Wind Market Report: 2021 Edition, (2021).
1050 [4] A. Du, Semi-Submersible, Spar and TLP – How to select floating wind foundation types?
1051 https://www.empireengineering.co.uk/semi-submersible-spar-and-tlp-floating-wind-foundations/.
1052 2021.
1053 [5] B. Bulder, M. van Hees, A. Henderson, R. Huijsmans, J. Pierik, E. Snijders, G. Wijnants, M. Wolf,
1054 Study to feasibility of and boundary conditions for floating offshore wind turbines, ECN, MARIN, TNO,
1055 TUD, MSC, Lagerway the Windmaster 26 (2002) 70-81.
1056 [6] A. Robertson, J. Jonkman, M. Masciola, H. Song, A. Goupee, A. Coulling, C. Luan, Definition of the
1057 semisubmersible floating system for phase II of OC4, National Renewable Energy Lab.(NREL), Golden,
1058 CO (United States), 2014.
1059 [7] C. Luan, Z. Gao, T. Moan, 2016, Design and analysis of a braceless steel 5-MW semi-submersible
1060 wind turbine, International Conference on Offshore Mechanics and Arctic Engineering, American
1061 Society of Mechanical Engineers, 2016, p. V006T09A052.
1062 [8] D. Son, C. Godreau, S. Diaz, T. Duarte, A. Peiffer, Design Process of the WindFloat Hull and
1063 Mooring System for WFA, Offshore Technology Conference, OnePetro, 2022.
1064 [9] W. Yu, K. Müller, F. Lemmer, D. Schlipf, H. Bredmose, M. Borg, T. Landbø, H. Andersen, LIFES50+
1065 D4. 2: Public definition of the two LIFES50+ 10 MW floater concepts, University of Stuttgart (2018).
1066 [10] Z. Zhao, W. Shi, W. Wang, S. Qi, X. Li, Dynamic analysis of a novel semi-submersible platform for
1067 a 10 MW wind turbine in intermediate water depth, Ocean Engineering 237 (2021) 109688.
1068 [11] Q. Cao, E.E. Bachynski-Polić, Z. Gao, L. Xiao, Z. Cheng, M. Liu, Experimental and numerical
1069 analysis of wind field effects on the dynamic responses of the 10 MW SPIC floating wind turbine
1070 concept, Ocean Engineering 261 (2022) 112151.
1071 [12] C. Souza, P. Berthelsen, L. Eliassen, E. Bachynski, E. Engebretsen, H. Haslum, Definition of the
1072 INO WINDMOOR 12 MW base case floating wind turbine, Tech. Rep. OC2020 A-044 (2021).
1073 [13] C. Allen, A. Viscelli, H. Dagher, A. Goupee, E. Gaertner, N. Abbas, M. Hall, G. Barter, Definition of
1074 the UMaine VolturnUS-S reference platform developed for the IEA Wind 15-megawatt offshore
1075 reference wind turbine, National Renewable Energy Lab.(NREL), Golden, CO (United States). 2020.
1076 [14] IEC-TS-61400-3-2, Wind energy generation systems - Part 3-2: Design requirement for floating
1077 offshore wind turbines. Geneva, Switzerland., 2019.

32 | 33
1078 [15] DNV, DNVGL-ST-0119: Floating wind turbine structures (Edition June 2021), Tech. rep. 2021. url:
1079 http://www. dnvgl. com, 2021.
1080 [16] B. Veritas, Classification and certification of floating offshore wind turbines, Guidance Note NI
1081 572 (2015).
1082 [17] E.E. Bachynski, T. Moan, Design considerations for tension leg platform wind turbines, Marine
1083 Structures 29(1) (2012) 89-114.
1084 [18] G. Ferri, E. Marino, C. Borri, Optimal dimensions of a semisubmersible floating platform for a 10
1085 MW wind turbine, Energies 13(12) (2020) 3092.
1086 [19] G. Ferri, E. Marino, N. Bruschi, C. Borri, Platform and mooring system optimization of a 10 MW
1087 semisubmersible offshore wind turbine, Renewable Energy 182 (2022) 1152-1170.
1088 [20] M. Karimi, M. Hall, B. Buckham, C. Crawford, A multi-objective design optimization approach for
1089 floating offshore wind turbine support structures, Journal of Ocean Engineering and Marine Energy
1090 3(1) (2017) 69-87.
1091 [21] S. Lefebvre, M. Collu, Preliminary design of a floating support structure for a 5 MW offshore
1092 wind turbine, Ocean engineering 40 (2012) 15-26.
1093 [22] S. Wang, T. Moan, Z. Gao, Conceptual global design of semi-submersible hulls for floating wind
1094 turbines. Technical report (number: 90814100/642002) , prepared for Power China, Department of
1095 Marine Technology, NTNU, Nov. 9, 2022. (Confidential). 2022.
1096 [23] S. Wang, T. Moan, Serviceability limit state assessment of semi-submersible floating wind
1097 turbines., 42nd International Conference on Ocean, Offshore & Arctic Engineering. Accepted., 2023.
1098 [24] DNVGL, SESAM User Manual, WADAM, Wave Analysis by Diffraction and Morison theory, Høvik,
1099 Norway (2019).
1100 [25] E.N. Wayman, Coupled dynamics and economic analysis of floating wind turbine systems,
1101 Massachusetts Institute of Technology, 2006.
1102 [26] C. Bak, F. Zahle, R. Bitsche, T. Kim, A. Yde, L.C. Henriksen, M.H. Hansen, J.P.A.A. Blasques, M.
1103 Gaunaa, A. Natarajan, The DTU 10-MW reference wind turbine, Danish wind power research 2013,
1104 2013.
1105 [27] A. Pegalajar-Jurado, H. Bredmose, M. Borg, J.G. Straume, T. Landbø, H.S. Andersen, W. Yu, K.
1106 Müller, F. Lemmer, State-of-the-art model for the LIFES50+ OO-Star Wind Floater Semi 10MW
1107 floating wind turbine, Journal of Physics: Conference Series, IOP Publishing, 2018, p. 012024.
1108 [28] S. Wang, T. Moan, Developing an efficient ULS analysis procedure for pontoon internal load
1109 effects of a semi-submersible hull structure for floating wind turbines, Under preparation (2023).
1110 [29] DNVGL, Recommended Practices (RP).“Column-Stabilised Units”, DNVGL-RP-C103, 2015.
1111 [30] L. Li, Z. Gao, T. Moan, Joint distribution of environmental condition at five european offshore
1112 sites for design of combined wind and wave energy devices, Journal of Offshore Mechanics and
1113 Arctic Engineering 137(3) (2015).
1114 [31] S. Wang, T. Moan, Z. Jiang, Influence of variability and uncertainty of wind and waves on fatigue
1115 damage of a floating wind turbine drivetrain, Renewable Energy 181 (2022) 870-897.
1116 [32] IEC61400‐3, International Standard ‘Wind Turbines–Part 3: Design requirements for offshore
1117 wind turbines’. International Electrotechnical Commission: Geneva, Switzerland., 2009.
1118 [33] S. Wang, T. Moan, Methodology of load effect anlysis for structural design of semi-submersible
1119 floating wind turbines: case study of column design., Under preparation (2023).
1120 [34] S. Wang, T. Moan, Z. Gao, Global structural load effect analysis of a semi-submersible hull for a
1121 10-MW floating wind turbine under still water, wind, and wave loads., Revision submitted to Marine
1122 Structures (2023).

1123

33 | 33

View publication stats

You might also like