Next Article in Journal
A Quick Pheromone Matrix Adaptation Ant Colony Optimization for Dynamic Customers in the Vehicle Routing Problem
Previous Article in Journal
Fluid–Structure Interaction Analysis of Manta-Bots with Self-Induced Vertical Undulations during Fin-Based Locomotion
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Motion Analysis of International Energy Agency Wind 15 MW Floating Offshore Wind Turbine under Extreme Conditions

by
Zengliang Chang
1,
Yueming Zheng
2,
Meng Qu
1,
Xingguo Gao
1,
Xiaojie Tian
2,3,* and
Guijie Liu
2,3
1
Shandong Electric Power Engineering Consulting Institute Co., Ltd., Ji’nan 250199, China
2
Department of Mechanical and Electrical Engineering, Ocean University of China, Qingdao 266100, China
3
Key Laboratory of Ocean Engineering of Shandong Province, Ocean University of China, Qingdao 266100, China
*
Author to whom correspondence should be addressed.
J. Mar. Sci. Eng. 2024, 12(7), 1166; https://doi.org/10.3390/jmse12071166
Submission received: 20 May 2024 / Revised: 27 June 2024 / Accepted: 3 July 2024 / Published: 11 July 2024
(This article belongs to the Section Ocean Engineering)

Abstract

:
In recent years, ultra-large-scale offshore wind turbines have attracted widespread attention. However, accurately evaluating the motion responses of offshore wind turbines under extreme conditions, especially for semisubmersible floating off-shore wind turbines, is often challenging. In order to assess the operational behavior of wind turbines under wind and wave loads, this paper adopted a numerical analysis method to solve the motion responses under extreme conditions. It specifically examines the motion responses of the IEA 15 MW wind turbine in terms of surge, heave, and pitch direction, focusing on environmental loads that occur once every 50 years. The results show that the wind turbine can still operate normally under the Ultimate condition. However, the average amplitude increased by 7% in the pitch direction and decreased by 4% in the heave direction compared to the rated condition. Under extreme conditions (occurring once every 50 years), with the wind turbine parked, the average amplitude in the surge direction reduced by 33%, while the average amplitude in the pitch direction reduced by 106%. Thus, it is essential to pitch the blades and brake the generator in extreme environmental conditions to ensure the safety of the wind turbine.

1. Introduction

At the 80th Marine Environment Protection Committee (MEPC 80), governments adopted the IMO’s revised greenhouse gas (GHG) strategy [1]. And the objective is to reach net-zero GHG emissions from international shipping close to 2050. Offshore wind energy, especially offshore wind turbines, has attracted considerable attention in power and renewables [2]. According to DNV’s Energy Transition Outlook 2023, the installed global floating offshore wind capacity will have grown from today’s 200 MW to almost 270 GW by 2050. The floating offshore wind turbine (FOWT) has a large structure and size, and the structural motion response is complex. In order to study the dynamic responses of the wind turbine, researchers have conducted a series of studies using coupling analysis software, aiming to accurately assess the characteristics of the structure.
In as early as 2005, the National Renewable Energy Laboratory (NREL) developed the popular 5 MW Reference Wind Turbine (RWT) [3], followed by the Technical University of Denmark’s (DTU) 10 MW RWT [4]. The NREL, using numerical simulation and model experiment methods, developed the OC3-Monopile and OC4-DeepC platforms to support the 5 MW RWT [5,6]. Meanwhile, a braceless steel platform was designed by the Norwegian Research Centre for Offshore Wind Technology to support the 5 MW Reference Wind Turbine (RWT) [7]. It consists of a central column, three side columns, and three pontoons; it was developed using numerical analysis methods. Additionally, the NAUTILUS-10 floating substructure [8] was provided by the LIFE 50+ project to support the DTU 10 MW RWT. In Japan, the FORWARD Semi-Sub 2 MW FOWT and V-shape Semi-Sub 7 MW FOWT have been developed and installed on a four-column floating platform [9,10]. An 8.4 MW wind turbine has been designed for installation on the WindFloat platform as part of the WindFloat Atlantic (WFA) project [11]. A concrete multi-column floating platform has been conceptually designed by Yokohama National University [12]. Moreover, several barge-type platforms have been designed and experimented with to support wind turbines [13,14]. With the development of deep water turbines and large turbines, elevated requirements are put forward for the stabilization of the floating wind turbines. When the wind turbines are operated under extreme conditions, they are a highly nonlinear system under floating body motions and extreme wave–structure interactions. The motion characteristics of the whole offshore wind structures are crucial for the generation of power, as well as their safety.
Yang et al. [15] developed a coupling framework based on FAST and AQWA through modifications on source codes of the user_force.dll of AQWA, obtaining the motion of multi-body platforms under the action of wind and waves. The framework has the ability to simulate multi-body platforms and to observe unique results, which contrasts with traditional tools. Guo et al. [16] developed a multibody modeling framework, namely TorqTwin. The framework can predict the motion response of the floating platform by calling on the source code of OpenFAST through Python library. However, the portability of the code is limited. Domene et al. [17] incorporated a direct air capture (DAC) system into the IEA 15 MW Reference Wind Turbine, analyzed the motion response, and compared it to the initial design. Chen et al. [18] used a coupling analysis software based on OrcaFlex to conduct a time domain simulation for 15 MW RWTs. The analysis considered three frequency regions (low frequency, wave frequency, and high frequency) and revealed that the tension response of the mooring line would be significantly affected by high-frequency motions originating from the superstructure. Using the computational fluid dynamics (CFD) method, Xue et al. [19] utilized the FOWT-UALM-SJTU solver to calculate the motion response of OC4 FOWT under different wave conditions and found that wind turbines with larger initial inclination angle recovered faster. Tran et al. [20] used CFD to analyze the motion responses of the OC4 semisubmersible FOWT based on an overset grid technique, and found that the motion of the floating platform has good consistency compared with the calculation results of FAST. Xu et al. [21] conducted a secondary development based on OpenFAST, analyzed the dynamic response of an FOWT in different stages of typhoon transit, and found that yaw errors would lead to an increased mooring load. Guo et al. [22] conducted an experimental study based on a 12 MW semisubmersible floating wind turbine, obtaining its motion and mooring response, and found that the motion response and mooring tension in the 60° and 90° directions were greater than those in the 0° direction. Xiang et al. [23] discussed the influence of different mooring lengths and counterweights on the motion response of semisubmersible wind turbines, and found that the motion of the floating platform in the pitch direction was affected. Zhang et al. [24] used different mooring models to compare the motion performances and mooring tension of FOWTs, and found that the quasi-static method still has a better calculation accuracy in most cases, but the finite element method is more suitable for fatigue calculations. Lozon et al. [25] conducted a dynamic analysis of shared mooring for floating wind farms and found that shared mooring does not cause an increase in dynamic response. Gao et al. [26] discussed the differences in the dynamic responses of FWOTs when the platform is considered as either rigid or flexible. The results show that accounting for the flexibility of the tower and platform leads to a significant increase in both the motions in the pitch direction and the tower-base force. Huang [27] constructed a shell-based Finite Element Model (FEM) to simulate motion responses under seismic loads, and demonstrated the flexible foundation of FOWTs presents a lower stress. Rinker et al. [28] compared the aeroelastic responses from HAWC2 and OpenFAST for the IEA 15 MW RWT and found that the aerodynamic loads agree quite well. Most recently, the IEA 22 MW ultra-large-scale RWT and a floating platform have been designed in IEA Wind TCP Task 55 [29]. While some significant and useful conclusions have been obtained, these studies are inadequate for understanding the motion responses of ultra-large-scale floating wind offshore turbines (FWOTs) under extreme conditions. Therefore, it is necessary to conduct a more in-depth analysis of wind turbine motion responses in extreme conditions.
In the present work, a classical semisubmersible floating offshore wind turbine, which consists of four columns, a heave plate, and three catenary chain mooring lines, developed by the International Energy Agency is studied. The hydrodynamic and motion characteristics under extreme conditions are crucial. The main focus is to utilize a software tool to predict the motion characteristics of floating offshore wind turbines and the hydrodynamic forces excited by highly nonlinear environmental loads.
The remainder of this paper is organized as follows: Section 2 describes the methodology, while Section 3 presents the model data of the FOWTs. In Section 4, the wind and wave loads utilized in the simulation are discussed, and a comparison with the results of the free decay is made to demonstrate the accuracy of the model. The results are presented in Section 5. Finally, the conclusions are drawn in Section 6.

2. Methodology

2.1. Dynamical Equation of FOWTs

Influenced by aerodynamic load, hydrodynamic load, and mooring load, the FOWT is a complex rigid–flexible coupled dynamic system. When coupled with wind and waves, its equation of motion in the time domain can be expressed as follows:
M + M A x ¨ + B R + B V x ˙ + K S + K M x = F W a v e + F W i n d + F M o o r i n g
where M is the mass matrix of the structure; M A is the added mass matrix of the platform at infinite frequency; B R is the radiation damping matrix; B V is the viscous damping matrix; K S is the hydrostatic stiffness matrix; K M is the mooring stiffness matrix; x , x ˙ , and x ¨ are, respectively, the displacement, velocity, and acceleration of each DOF; F W a v e is the wave load; F W i n d is the aerodynamic load; and F M o o r i n g is the mooring load.

2.2. Aerodynamic Loading

The most widely used methods for the aerodynamic load calculation of FOWTs are Blade Element Momentum (BEM), free vortex wake (FVW), and computational fluid dynamics (CFD) methods. In OpenFAST, BEM is utilized to primarily calculate the aerodynamic loads. The basic process of BEM involves calculating the thrust and torque of the wind turbine using momentum theory and blade element theory, respectively. Then, the induction factor and the actual angle of attack of the blade can be solved. According to the blade element theory, the force on a certain blade element of the wind turbine can be solved as follows:
d T = B 1 2 ρ V t o t a l 2 ( C l cos φ + C d sin φ ) c d r
d Q = B 1 2 ρ V t o t a l 2 ( C l sin φ C d cos φ ) c r d r
where d T and d Q are, respectively, the thrust and torque on the unit blade element; B is the number of blades; ρ is the density of air; V t o t a l is the relative speed of air flow over the blade; C l and C d are, respectively, the lift coefficient and drag coefficient; φ is the angle of indraft; c is the chord length of blade element; and r is the extension radius of the blade element.
After calculating the initial aerodynamic performance through the BEM method, the Generalized Dynamic Wake (GDW) model is used to further accurately calculate the aerodynamic loads in OpenFAST. And the GDW model is based on the Pitt–Peters–He acceleration potential method.

2.3. Hydrodynamic Loading

Hydrodynamic load calculation methods of FOWTs include the potential theory and the Morison equation, and sometimes a combination of the two methods is used. The potential theory is primarily suitable for large-scale components with a structural characteristic size-to-wave wavelength ratio greater than 0.2. The Morrison equation is applicable to structures with a ratio of structural characteristic size-to-wave wavelength of less than 0.2, and it is commonly used in the calculation of slender and small-scale components such as cross bracing, diagonal bracing of floating foundations, and mooring cables.
The potential theory assumes that the fluid is an ideal fluid and defines the velocity potential of the fluid in an irrotational flow field. The Laplace equation and the corresponding boundary conditions of the velocity potential are as follows:
φ ( X ) = φ I + φ D + j = 1 6 φ R
2 φ / x 2 + 2 φ / y 2 + 2 φ / z 2 = 0
φ / z = 0
φ / φ g 2 + 2 φ / t 2 = 0
φ / n = 0
where φ ( X ) is the velocity potential function; φ I and φ D are, respectively, the incident potential and the diffraction potential function; φ R is the radiation potential function; z is the depth of water; and n is the normal vector on the surface of a floating structure.
Morison’s equation defines the wave force on a structure as the sum of the drag force and the inertial force, as follows:
d F H = 1 2 C D ρ D u x | u x | d z + C M ρ π D 2 4 u x t d z
where C D is the drag force coefficient, ρ is the density of the fluid, D is the diameter of the member, u x is the incoming flow velocity in the vertical direction of the structure axis, and C M is the inertia coefficient.

2.4. Mooring Loading

The quasi-static method, lumped-mass method, and finite element method are often used to calculate the mooring loads. In this paper, the lumped-mass method is used to calculate the mooring load.
The mooring cable is divided into a number of spring–mass–damping systems in the lumped-mass method. Parameters such as mass, gravity, and buoyancy are concentrated on the nodes. The lower end of the first segment is connected to the anchor, and the upper end of the nth segment is connected to the platform. The equilibrium equation of the mooring system can be described as follows:
M i a i = F T i F T i 1 + F D i + F A i W i
F DN i = 1 2 ρ C DN D u N u N
F DT i = 1 2 ρ C DT ( π D ) u T u T
F A i = ρ C M V ( a i )
where a i is the acceleration of node i ; M i is the mass of node i ; W i is the gravity of node i ; F T i 1 and F T i are, respectively, the tension generated by two elements connected by node i ; and F A i , F D i , and F DT i are, respectively, the fluid inertia force, normal resistance, and tangential force of node i .

2.5. Process of Analysis

For the motion analysis of floating offshore wind turbines, the hydrodynamic model of structure is established in ANSYS-AQWA (ANSYS, 2022 R2), as indicated by the orange part in Figure 1.
The VolturnUS-S semisubmersible platform is used in the calculation, excluding the upper wind turbine structures and mooring cables. Basic parameters, based on the platform’s physical characteristics, are provided, and different incident wave periods are set for the analysis. After solving, an ASCII-formatted hydrodynamic database is generated. The database includes the hydrostatic stiffness matrix, additional mass matrix, radiation damping matrix, and wave excitation force of the structure at the different incident frequencies. These matrices can be programmatically converted into WAMIT format. Subsequently, the hydrodynamic coefficients are transferred to the next stage.
It is worth mentioning that the viscosity of the fluid plays an important role on the calculation [30]. Therefore, the viscous effects of the fluid are taken into account, and details about the viscous damping matrix can be found in another paper [31].
Subsequently, aerodynamic loads, hydrodynamic loads, structural flexibility, and the servo control system are all taken into account in OpenFAST(v3.4.1), as shown in Figure 1. This is another important part during the motion analysis process. OpenFAST is a typical aero-hydro-servo-elastic coupling framework and uses Kane’s method to solve the equations [32]. It takes into account the impact of the control system, and the solution results are more realistic [33]. Additionally, the finite element method (FEM) is utilized to account for the elasticity of the structure. The FWOT is calculated as a coupled system and the structural motion responses can be obtained.

3. Model Data

3.1. Description of the Floating Offshore Wind Turbine

The International Energy Agency (IEA) defines the 15 MW RWT with the aim of assisting researchers in exploring new technologies or design methodologies. This horizontal axis turbine, featuring three blades, is mounted on the VolturnUS-S semisubmersible platform. And the VolturnUS-S semisubmersible platform was designed by the University of Maine (UMaine). Table 1 presents the key parameters of the IEA 15 MW Reference Wind Turbine.
As depicted in Figure 2a, the platform comprises four columns and the heave plate is designed at the bottom to improve the performance in the heave direction. The plat-form is designed with center symmetry, and the wind turbine is arranged on the central column. The dimensions of the hull are given in the technical report [34] and the platform properties are shown in Table 2.

3.2. Description of Mooring System

The arrangement of the mooring system is shown in Figure 2b. The mooring system, designed for a water depth of 200 m, consists of three catenary chain mooring lines, with a horizontal angle of 120° between two chains. The diameter of each mooring line is 185 mm, with a length of 837.6 m, and the fairlead is designed at a depth of 14 m below the SWL. Unfortunately, the chain is designed for strength based on the displacement boundaries (with a peak surge-sway offset of under 25 m), and fatigue strength was not studied. The properties of the mooring system are summarized in Table 3.

4. Loads and Model Verification

4.1. Load Case

The wind and wave loads are designed based on the joint probability function, and current load is not considered in the analysis. Table 4 presents the selection of representative Design Load Cases (DLCs). The two-parameter JONSWAP spectrum is utilized to describe the waves [35], incorporating significant wave height (Hs) and wave spectrum peak period (Tp), with wind speed values in Table 4 representing the mean speed at hub height. Turbulent wind conditions are assumed in all scenarios, with the wind spectrum model defined as the IEC Kaimal spectral model. The Normal Turbulence Model (NTM) is adopted in DLC 1-2, while the Extreme Wind Speed Model (EWM), occurring once in 50 years, is utilized in DLC 3. The wind class is set as Class A, and turbulent intensity factors and other parameters can be found in IEC 61400-3 [36]. And the time history curve of wind speed and wave height are shown in Figure 3. In Figure 3, the relationships between time and wave height, as well as time and wind speed, are intuitively illustrated. The wave conditions in DLC 2 and 3 are significantly more severe compared to in DLC 1. Additionally, the wind speed under extreme conditions is approximately four times higher than that under the other conditions.
DLC 1 is the rated working condition, with a 10 m/s wind speed and a 1.54 m wave height. It is used as a reference for comparison. For the DLC 2, the wind speed and wave height (Hs) increases a lot, and it is used to simulate Ultimate conditions. The DLC 3 corresponds to the once-in-50-year wave conditions. Under this extreme condition, the wind turbine is parked and the blades are feathered, while in other cases, wind turbine operation is actively controlled (AC).

4.2. Free Decay Test

By applying initial boundary conditions in each direction of the structure, the natural frequency of the rigid-body modes of the system can be obtained [37]. The vibration frequency of the structure obtained through free decay testing is independent of the initial conditions. It is solely determined by the inherent characteristics of the system. The natural frequency of surge and sway is the same, and roll and pitch are also the same. Therefore, only the calculation results in the surge, heave, roll, and yaw directions are listed, as shown in Figure 4 and Table 5.
In Figure 4a, the time history curve of free decay shows a period of around 130 s in the surge direction, and its natural frequency is shown to be 0.008 Hz in the frequency domain in Figure 4e. In general, the free heave motion of the structure can also be used as part of the model’s validation [38]. And the period of the heave motion is approximately 20 s, corresponding to a natural frequency of 0.05 Hz in the heave direction, as shown in Figure 4b,f. Similarly, in Figure 4c,g, the period is around 30 s, and the natural frequency is 0.035 Hz in the pitch direction. In Figure 4d,h, the period is around 80 s, and the natural frequency is 0.012 Hz in the yaw direction. Comparison with the results from the reference [34] reveal a close consistency. The errors are all not greater than 0.001 Hz, which means the results are in good agreement with the reference data. Therefore, the constructed model is validated.

5. Results and Discussion

In this section, the motion characteristics, nacelle acceleration responses, and mooring forces of the IEA 15 MW Reference Wind Turbine are studied under different wind and wave combination conditions.

5.1. Motion Responses

The motion responses under different conditions are calculated and the simulation time is set to 3500 s. Then, the influence of different conditions on the motion response is obtained. Figure 5 shows the time history curve of the pitch, heave, and surge motions of the floating platform under different conditions. It can be seen from Figure 5 that the surge and sway motion are the main motion forms.
Transient effects can be observed in approximately the first 200 s and are not significant in the heave direction. The motion in all three directions starts from an initial position of 0 m. The most substantial change was observed in the surge direction, where the motion increased from 0 m to approximately 25 m under Ultimate 1 and Extreme conditions, and from 0 m to 15 m under the Operational 1 condition, before beginning to fluctuate. Similarly, in the pitch direction, the motion increased from 0 m to approximately 6 m under the Ultimate 1 and Extreme conditions, and from 0 m to 2 m under the Operational 1 condition, before beginning to fluctuate.
Table 6 describes the statistical parameters of motion response in each degree of freedom direction under different conditions. Here, ‘Max’, ‘Min’, and ‘Mean’ denote the maximum, minimum, and average values of motion response, respectively. ‘STD’ represents the standard deviation, and ‘Range’ indicates the difference between the maximum and minimum values (maximum–minimum). The values presented here are derived from data between 500 s and 3500 s, with transient results excluded from the statistical analysis.
Based on the calculation results, it is evident that when only the wave height and wave period are increased, the wind turbine can still operate normally, without a corresponding increase in wind speed (referred to as ‘Ultimate 1’). Furthermore, the responses in the surge and pitch directions exhibit minimal variation compared to the rated working condition (Operational 1). Specifically, the average motion response amplitude increased by 7% in the surge direction and decreases by 4% in the pitch direction.
Under the extreme environmental condition, occurring once in 50 years (referred to as ‘Extreme’), the average motion response amplitude is reduced by 33% in the surge direction, and is reduced by 106% in the pitch direction, compared to the rated working condition. This reduction is likely due to the pitched blades and braked generator.
However, the influence of waves on wind turbine response is most notable in the heave direction, where the ‘Range’ increased by a factor of 12 in ‘Ultimate 1’ and by a factor of 16 in ‘Extreme’ scenarios.

5.2. Nacelle Acceleration Responses

The acceleration of the nacelle has an important influence on the force exerted on the structure, and the nacelle acceleration responses serve as a reference for studying the tower responses. Therefore, the nacelle acceleration responses are obtained. Figure 6 presents the acceleration responses in the x and y directions, while Figure 7 displays the maximum (Max) values and standard deviation (STD) values of the response. The effects of transient phenomena are hardly noticeable in the acceleration response.
It can be seen from Figure 6 that the nacelle acceleration responses are all between −2 m/s and 2 m/s in the x direction, and are between −1 m/s and 1 m/s in the y direction under different conditions. In Figure 7, the maximum nacelle acceleration response is measured at 1.65 m/s2. According to relevant regulations, when the nacelle acceleration response exceeds 6 m/s2, the risk of structural failure is expected to rise. Therefore, the floating offshore wind turbine structure is within a safe range.
It can also be found that the acceleration response of the structure is notably affected by the wave height and wave period. And the acceleration response of structure in the windward direction (X direction) is significantly greater than that in the lateral direction (Y direction). There is little difference in the acceleration response between the Ultimate 1 and Extreme conditions in the X direction. However, the acceleration response is larger under both the Ultimate 1 and Extreme conditions compared to the rated working condition. Under the Extreme condition, the maximum acceleration response in the Y direction is approximately three times greater than that under rated conditions, with the standard deviation being approximately 3.6 times larger.

5.3. Mooring Force Responses

There is no doubt that the state of the mooring cable is particularly important for the safety of floating wind turbine structures. The force responses of three mooring lines can be calculated under different conditions. Figure 8 shows the time history curve of fairlead tensions, and Figure 9 shows the statistic characteristics of fairlead tensions.
Transient effects, observable within the initial 100 s, exhibit different trends in mooring cables distributed in two directions around the FOWT. Due to the presence of initial tension, the tension in the mooring lines initially changes from 2.5 MN.
The tension of fairlead 1 increased from 2.5 MN to approximately 5 MN under the Ultimate 1 and Extreme conditions, and from 2.5 MN m to 4 MN under the Operational 1 condition, before beginning to fluctuate. Similarly, the tension of fairleads 2 and 3 decreased from 2.5 MN to approximately 1.8 MN under the Ultimate 1 and Extreme conditions, and from 2.5 MN to 2 MN under the Operational 1 condition, before beginning to fluctuate.
From the results, it can be observed that the tension of the mooring cable increases with deteriorating environmental conditions. When the wind turbine is operational, the average tension of the cable on the windward side (fairlead 1) is greater than that on the leeward side. Under DLC3, the average tension of the cable on the windward side decreased, while the tension of the two mooring cables on the leeward side increased. In all conditions, the mooring cable tension is far below its breaking strength (22 MN [34]), indicating a large safety margin.

6. Conclusions

This paper focuses on the IEA 15 MW ultra-large-scale wind turbine structure, to solve hydrodynamic coefficients and to study the platform’s dynamic responses under different environmental conditions. The conclusions can be summarized as follows:
(1)
When facing extreme environmental conditions, accurately determining the timing of generator shutdown is essential. During such conditions, particular emphasis should be placed on monitoring the wind turbine’s motion response in the heave direction.
(2)
Wave conditions significantly influence the acceleration response of the wind turbine. In extreme environmental conditions, the nacelle acceleration response of the wind turbine notably increases even during shutdown. Therefore, careful observation of the tower’s acceleration response is necessary.
(3)
During wind turbine shutdown, the cable tension on the windward side decreases by approximately 20%, while the mooring tension on the leeward side increases by about 8%. In all scenarios, the mooring tension on the windward side is greater and should be prioritized.
In this study, the motion response of the IEA 15 MW floating wind turbine under extreme condition is investigated to gain its motion responses in challenging environmental conditions. This research holds significance as a reference for the safety monitoring of wind turbine structures. However, there is a gap in the research regarding whether wind turbines should be shut down when encountering complex sea conditions. It is necessary to explain the limitations of this study. The research method used in this study is singular. Due to the limitations of experimental equipment, experimental analysis could not be conducted, and there is no publicly available dataset for comparison. The accuracy of the analysis may be validated in future studies. Additionally, due to limitations in funds and personnel, this paper only analyzes a small number of extreme conditions. The accuracy of the analysis and the verification of a wider range of working conditions may be achieved in future research. Increasing the number of design conditions to examine the optimal timing for wind turbine shutdowns can also be beneficial in future studies.

Author Contributions

Conceptualization, Z.C.; Methodology, Z.C. and Y.Z.; Software, Y.Z. and X.G.; Validation, Y.Z. and X.G.; Formal analysis, Z.C.; Data curation, M.Q.; Writing—review and editing, X.T. and G.L.; Supervision, M.Q., X.T. and G.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research is funded by the Natural Science Foundation of Shandong Province of China (No. ZR2021ME014).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The raw data supporting the conclusions of this article will be made available by the authors on request.

Conflicts of Interest

Authors Zengliang Chang, Meng Qu and Xingguo Gao were employed by the Shandong Electric Power Engineering Consulting Institute Co. The remaining authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

References

  1. Wissner, N.; Healy, S. Raising Ambition Levels at the IMO for 2050, An Overview of the Key Issues at Stake at MEPC 80; European Parliamentary Research Service: Brussels, Belgium, 2023. [Google Scholar]
  2. Ren, Z.; Verma, A.S.; Li, Y.; Teuwen, J.J.E.; Jiang, Z. Offshore wind turbine operations and maintenance: A state-of-the-art review. Renew. Sustain. Energy Rev. 2021, 144, 110886. [Google Scholar] [CrossRef]
  3. Jonkman, J.; Butterfield, S.; Musial, W.; Scott, G. Definition of a 5-MW Reference Wind Turbine for Offshore System Development; National Renewable Energy Lab. (NREL): Golden, CO, USA, 2009. [Google Scholar]
  4. Bak, C.; Zahle, F.; Bitsche, R.; Kim, T.; Yde, A.; Henriksen, L.C.; Hansen, M.H.; Blasques, J.P.A.A.; Gaunaa, M.; Natarajan, A. The DTU 10-MW reference wind turbine. In Proceedings of the Danish Wind Power Research 2013, Fredericia, Denmark, 27–28 May 2013. [Google Scholar]
  5. Jonkman, J. Definition of the Floating System for Phase IV of OC3; National Renewable Energy Lab. (NREL): Golden, CO, USA, 2010. [Google Scholar]
  6. Robertson, A.; Jonkman, J.; Masciola, M.; Song, H.; Goupee, A.; Coulling, A.; Luan, C. Definition of the Semisubmersible Floating System for Phase II of OC4; National Renewable Energy Lab. (NREL): Golden, CO, USA, 2014. [Google Scholar]
  7. Luan, C.; Gao, Z.; Moan, T. Design and analysis of a braceless steel 5-mw semi-submersible wind turbine. In Proceedings of the International Conference on Offshore Mechanics and Arctic Engineering, Busan, Republic of Korea, 19–24 June 2016; p. V006T009A052. [Google Scholar]
  8. Yu, W.; Müller, K.; Lemmer, F. D4. 2 Public Definition of the Two LIFES50+ 10MW Floater Concepts; Technical Report; University of Stuttgart: Stuttgart, Germany, 2018. [Google Scholar]
  9. Yoshimoto, H.; Awashima, Y.; Kitakoji, Y.; Suzuki, H. Development of floating offshore substation and wind turbine for Fukushima FORWARD. In Proceedings of the International Symposium on Marine and Offshore Renewable Energy, Tokyo, Japan, 28–30 October 2013; pp. 28–30. [Google Scholar]
  10. Ishihara, T.; Yamaguchi, A.; Taki, S. The challenge to the world’s first floating wind farm. In Proceedings of the Grand Renewable Energy 2014 Proceedings, Tokyo, Japan, 27 July–1 August 2014. [Google Scholar]
  11. Cruciani, M. Offshore Wind Power Floating in its Industrial and Technological Dimension. Ifri 2019, 51, 11–42. [Google Scholar]
  12. Hasan, M.S.; Munbua, W.; Malta, E.B.; Gonçalves, R.T.; Fujiyama, C.; Maekawa, K. Conceptual Design of a Concrete Multi-Column Floating Platform Supporting a 10 MW Offshore Wind Turbine. In Proceedings of the ASME 2022 41st International Conference on Ocean, Offshore and Arctic Engineering, Hamburg, Germany, 21–30 July 2022. [Google Scholar]
  13. Dachollom, G.L.; Hejazi, F.; Yusuf, B. Development of Ultra High-Performance Fiber reinforced concrete barge for 5 MW wind turbine. Structures 2023, 53, 1349–1368. [Google Scholar] [CrossRef]
  14. Yang, R.-Y.; Wang, C.-W.; Huang, C.-C.; Chung, C.-H.; Chen, C.-P.; Huang, C.-J. The 1:20 scaled hydraulic model test and field experiment of barge-type floating offshore wind turbine system. Ocean Eng. 2022, 247, 110486. [Google Scholar] [CrossRef]
  15. Yang, Y.; Bashir, M.; Michailides, C.; Li, C.; Wang, J. Development and application of an aero-hydro-servo-elastic coupling framework for analysis of floating offshore wind turbines. Renew. Energy 2020, 161, 606–625. [Google Scholar] [CrossRef]
  16. Guo, F.; Gao, Z.; Schlipf, D. Torqtwin—An Open-Source Reference Multibody Modeling Framework for Wind Turbine Structural Dynamics. SSRN 2024. [Google Scholar] [CrossRef]
  17. Domene, G.A.; Crawford, C. Dynamic analysis of a floating wind turbine platform with on-board CO2 direct air capture. Ocean Eng. 2024, 308, 118205. [Google Scholar] [CrossRef]
  18. Chen, J.; Su, Z.; Ding, Y.; Liu, A.; Duggal, A. Motion and Tension Response Study of a 15MW Offshore Reference Floating Wind Turbine With a Semisubmersible. In Proceedings of the International Conference on Offshore Mechanics and Arctic Engineering, Boston, MA, USA, 7–8 December 2022; p. V001T001A002. [Google Scholar]
  19. Xue, Y.-j.; Yang, X.-l.; Zhao, W.-w.; Wan, D.-c. Numerical investigation of the coupled aero-hydrodynamic performances of a semi-submersible floating offshore wind turbine with inclined columns. J. Hydrodyn. 2024, 36, 316–330. [Google Scholar] [CrossRef]
  20. Tran, T.T.; Kim, D. A CFD study of coupled aerodynamic-hydrodynamic loads on a semisubmersible floating offshore wind turbine. Wind. Energy 2018, 21, 70–85. [Google Scholar] [CrossRef]
  21. Xu, H.; Wang, L.; Zha, X.; Rui, S.; Shen, K.; Guo, Z. Dynamic response of floating offshore wind turbine under different stages of typhoon passage. Appl. Ocean Res. 2024, 148, 104047. [Google Scholar] [CrossRef]
  22. Guo, J.; Liu, M.; Fang, Z.; Xiao, L.; Chen, W.; Pan, X. Motion and mooring load responses of a novel 12 MW semi-submersible floating wind turbine: An experimental study. J. Offshore Mech. Arct. Eng. 2024, 146, 1–26. [Google Scholar] [CrossRef]
  23. Xiang, S.; Zhenya, T.; Ronghua, Z. Investigation on the Influence of Anchor Chain Length and Weight on the Motion Response of Floating Wind Turbine Platforms. Ocean Eng. Equip. Technol. 2024, 11, 59. [Google Scholar]
  24. Zhang, F.; Wood, C.; Wang, Q.; Nekstad, O. A comparison Study of the Mooring Simulation in the Design and Analysis of Floating Offshore Wind. In Proceedings of the Offshore Technology Conference, Houston, TX, USA, 6–9 May 2024; p. D022S058R011. [Google Scholar]
  25. Lozon, E.; Hall, M. Coupled loads analysis of a novel shared-mooring floating wind farm. Appl. Energy 2023, 332, 120513. [Google Scholar] [CrossRef]
  26. Gao, S.; Zhang, L.; Shi, W.; Wang, W.; Wang, B.; Li, X. Dynamic Analysis of a 10 MW Floating Offshore Wind Turbine Considering the Tower and Platform Flexibility. J. Ocean Univ. China 2024, 23, 358–370. [Google Scholar] [CrossRef]
  27. Huang, H.S. Simulations of 10MW wind turbine under seismic loadings. Compos. Struct. 2022, 279, 114686. [Google Scholar] [CrossRef]
  28. Rinker, J.; Gaertner, E.; Zahle, F.; Skrzypiński, W.; Abbas, N.; Bredmose, H.; Barter, G.; Dykes, K. Comparison of loads from HAWC2 and OpenFAST for the IEA Wind 15 MW Reference Wind Turbine. J. Phys. Conf. Ser. 2020, 1618, 052052. [Google Scholar] [CrossRef]
  29. Zahle, F.; Barlas, T.; Lønbæk, K.; Bortolotti, P.; Zalkind, D.; Wang, L.; Labuschagne, C.; Sethuraman, L.; Barter, G. Definition of the IEA Wind 22-Megawatt Offshore Reference Wind Turbine. DTU Wind E 2024. [Google Scholar] [CrossRef]
  30. Ciba, E.; Dymarski, P. Modelling of the viscosity effect of heave plates for floating wind turbines by hydrodynamic coefficients. Acta Mech. Autom. 2023, 17, 469–476. [Google Scholar] [CrossRef]
  31. Gaertner, E.; Rinker, J.; Sethuraman, L.; Zahle, F.; Anderson, B.; Barter, G.E.; Abbas, N.J.; Meng, F.; Bortolotti, P.; Skrzypinski, W. IEA Wind TCP Task 37: Definition of the IEA 15-Megawatt Offshore Reference Wind Turbine; National Renewable Energy Lab. (NREL): Golden, CO, USA, 2020. [Google Scholar]
  32. Sarkar, S.; Fitzgerald, B. Use of Kane’s Method for Multi-Body Dynamic Modelling and Control of Spar-Type Floating Offshore Wind Turbines. Energies 2021, 14, 6635. [Google Scholar] [CrossRef]
  33. Abbas, N.; Zalkind, D.; Pao, L.; Wright, A. A reference open-source controller for fixed and floating offshore wind turbines. Wind. Energy Sci. 2021, 2021, 1–33. [Google Scholar] [CrossRef]
  34. Allen, C.; Viscelli, A.; Dagher, H.; Goupee, A.; Gaertner, E.; Abbas, N.; Hall, M.; Barter, G. Definition of the UMaine VolturnUS-S Reference Platform Developed for the IEA Wind 15-Megawatt Offshore Reference Wind Turbine; National Renewable Energy Lab. (NREL): Golden, CO, USA, 2020. [Google Scholar]
  35. Mazaheri, S.; Imani, H. Evaluation and modification of JONSWAP spectral parameters in the Persian Gulf considering offshore wave characteristics under storm conditions. Ocean Dyn. 2019, 69, 615–639. [Google Scholar] [CrossRef]
  36. Quarton, D.C. An International Design Standard for Offshore wind Turbines: IEC 61400-3; Garrad Hassan and Partners, Ltd.: Bristol, UK, 2005. [Google Scholar]
  37. Sundar, V. Ocean Wave Mechanics: Applications in Marine Structures; John Wiley & Sons: Hoboken, NJ, USA, 2017. [Google Scholar]
  38. Gao, J.; Mi, C.; Song, Z.; Liu, Y. Transient gap resonance between two closely-spaced boxes triggered by nonlinear focused wave groups. Ocean Eng. 2024, 305, 117938. [Google Scholar] [CrossRef]
Figure 1. Flowchart between AQWA and OpenFAST.
Figure 1. Flowchart between AQWA and OpenFAST.
Jmse 12 01166 g001
Figure 2. General arrangement of the FOWT.
Figure 2. General arrangement of the FOWT.
Jmse 12 01166 g002
Figure 3. Time history of wave height (top) and wind speed (bottom).
Figure 3. Time history of wave height (top) and wind speed (bottom).
Jmse 12 01166 g003
Figure 4. Free decay time histories and natural frequency at still water line: (a) Surge direction, (b) Heave direction, (c) Pitch direction, (d) Yaw direction—Time histories; (e) Surge direction, (f) Heave direction, (g) Pitch direction, (h) Yaw direction—Natural frequencies.
Figure 4. Free decay time histories and natural frequency at still water line: (a) Surge direction, (b) Heave direction, (c) Pitch direction, (d) Yaw direction—Time histories; (e) Surge direction, (f) Heave direction, (g) Pitch direction, (h) Yaw direction—Natural frequencies.
Jmse 12 01166 g004
Figure 5. Motion responses of the floating offshore turbine under different conditions.
Figure 5. Motion responses of the floating offshore turbine under different conditions.
Jmse 12 01166 g005
Figure 6. Response curves of the nacelle acceleration.
Figure 6. Response curves of the nacelle acceleration.
Jmse 12 01166 g006
Figure 7. Nacelle acceleration response statistics.
Figure 7. Nacelle acceleration response statistics.
Jmse 12 01166 g007
Figure 8. Mooring force response curves.
Figure 8. Mooring force response curves.
Jmse 12 01166 g008
Figure 9. Mooring force response statistics.
Figure 9. Mooring force response statistics.
Jmse 12 01166 g009
Table 1. Parameters of the IEA 15 MW Reference Wind Turbine.
Table 1. Parameters of the IEA 15 MW Reference Wind Turbine.
ParameterUnitValue
Number of blades[pcs]3
Rotor diameter[m]242
Cut-in wind speed[m/s]3.0
Rated wind speed[m/s]10.6
Power rating[MW]15
Hub height[m]150
Cut-out wind speed[m/s]25.0
Maximum rotor speed[rpm]7.6
Table 2. Parameters of the Maine VolturnUS-S semisubmersible platform.
Table 2. Parameters of the Maine VolturnUS-S semisubmersible platform.
ParameterUnitValue
Overall platform mass[t]17,854
Displaced water volume[m3]20,206
Draft (with mooring system)[m]20
Center of mass below SWL[m]14.94
Center of buoyancy below SWL[m]13.63
Platform roll inertia about center of mass[kg m2]1.251 × 1010
Platform pitch inertia about center of mass[kg m2]1.251 × 1010
Platform yaw inertia about center of mass[kg m2]2.367 × 1010
Table 3. Parameters of the mooring system.
Table 3. Parameters of the mooring system.
ParameterUnitValue
Effective diameter of the chain[mm]185
Unstretched mooring line length[m]837.6
Fairlead depth below SWL[m]14
Anchor depth below SWL[m]13.63
Extensional stiffness EA[N]3.27 × 109
Table 4. Design Load Case matrix for the IEC 15 MW wind turbine.
Table 4. Design Load Case matrix for the IEC 15 MW wind turbine.
DLCNameTurbulence ModeWind (m/s)Wave ConditionController
Hs (m)Tp (s)
1Operational 1NTM10.001.547.65AC
2Ultimate 1NTM10.008.1012.80AC
3ExtremeEWM 50 year42.5010.7014.20Parked
Table 5. Rigid-body natural frequencies.
Table 5. Rigid-body natural frequencies.
Natural Frequency [Hz]SurgeSwayHeaveRollPitchYaw
Present0.0080.0080.0500.0350.0350.012
Reference [34]0.0070.0070.0490.0360.0360.011
Table 6. Statistics of the platform motions for different load cases.
Table 6. Statistics of the platform motions for different load cases.
DOFNameMaxMinMeanSTDRange
Surge (m)Operational 124.68011.19018.2432.94513.490
Ultimate 124.61012.47019.5092.20912.140
Extreme19.9105.39412.2312.40414.516
Heave (m)Operational 1−0.244−0.741−0.4770.0790.498
Ultimate 13.129−3.404−0.4320.9626.533
Extreme4.548−4.191−0.1881.3408.739
Pitch (°)Operational 15.7561.4163.7890.8144.340
Ultimate 16.1100.9453.6480.9045.165
Extreme2.509−2.777−0.2240.9075.286
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Chang, Z.; Zheng, Y.; Qu, M.; Gao, X.; Tian, X.; Liu, G. Motion Analysis of International Energy Agency Wind 15 MW Floating Offshore Wind Turbine under Extreme Conditions. J. Mar. Sci. Eng. 2024, 12, 1166. https://doi.org/10.3390/jmse12071166

AMA Style

Chang Z, Zheng Y, Qu M, Gao X, Tian X, Liu G. Motion Analysis of International Energy Agency Wind 15 MW Floating Offshore Wind Turbine under Extreme Conditions. Journal of Marine Science and Engineering. 2024; 12(7):1166. https://doi.org/10.3390/jmse12071166

Chicago/Turabian Style

Chang, Zengliang, Yueming Zheng, Meng Qu, Xingguo Gao, Xiaojie Tian, and Guijie Liu. 2024. "Motion Analysis of International Energy Agency Wind 15 MW Floating Offshore Wind Turbine under Extreme Conditions" Journal of Marine Science and Engineering 12, no. 7: 1166. https://doi.org/10.3390/jmse12071166

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop