1. Introduction
The offshore wind industry is growing continuously over the last few years. Recently, more wind turbines are being installed in deeper waters and further away from shore, which comprise a considerable share of the offshore wind resources with less turbulence [
1]. However, over deep-sea sites, the bottom-fixed concepts are no longer viable as these structures are not an economical solution for water depths more than 50 m [
2]. Therefore, several floating wind turbine system (FWTS) concepts are developed for use in deeper waters [
3,
4]. Among these floating concepts, the semisubmersible platform, stabilized mainly by buoyancy, is a promising technology and constitutes the most explored technology compared to the others. This is due to the applicability of these platforms in different water depths, good hydrodynamic performance, and facility in the installation process [
3,
5,
6].
Several examples of commissioned semisubmersible FWTSs are the 2 MW WindFloat Semisubmersible in Portugal [
7], 3x8.4 MW WindFloat Atlantic Semisubmersibles in Portugal [
8], 2 MW Fukushima Mirai 4-column Semisubmersible in Japan [
9], 7 MW Fukushima Shimpuu V-shape 3-column Semisubmersible [
9] and 5x9.5 MW plus 2 MW Kincardine Semisubmersibles in Scotland [
10]. Several FWTSs are under construction, such as 5MW Eolink 4-column Semisubmersible in France and 10MW concrete OO-Star Norway Semisubmersible [
8]. Some popular semisubmersible configurations are the WindFloat, Braceless platform, and VolturnUS-S, as shown in
Figure 1 [
11,
12,
13,
14,
15].
Figure 1 depicts the layout of two semisubmersible platforms.
Figure 2 (a) presents DeepCwind semisubmersible platform [
18] designed within the phase II of the Offshore Code Comparison Collaboration Continuation (OC4) project to support the National Renewable Energy Laboratory (NREL) offshore 5-MW wind turbine [
19]; while
Figure 1 (b) shows UMaine VolturnUS-S reference platform, which is developed to sustain the IEA 15-MW Offshore Reference Wind Turbine [
13]. Semisubmersible platforms may possess different offset columns. As shown by [
20], increasing these columns can reduce the platform motions. Although most of the semisubmersibles are made of steel, there is a tendency to use concrete for FWTS structures owing to lower construction cost, as well as superior durability and fatigue resistance of concrete compared to steel [
21,
22,
23,
24].
Nonetheless, FWTSs are less mature than their bottom-fixed counterparts owing to their more complicated operating conditions and higher prices. The FWTSs are subject to loads emanating from several sources, namely wave, current, wind, and, in certain locations, ice [
25,
26]. Although the similarities between FWTSs and floating platforms in the industry of oil and gas lead to a partial transfer of the technology [
27], there are numerous differences between the two industries, such as the aerodynamic loads acting on the wind turbines, which affect significantly the dynamics of FWTSs and give rise to several new technical challenges [
25,
28]. For instance, the coupled current-wave-wind loads acting on the FWTSs, as shown schematically in
Figure 3, may cause large platform motions along the degrees of freedom of motion of the structure, which combined with the blade aeroelastic deformations results in a highly dynamic inflow to the turbine rotor affecting the turbine structural integrity [
29,
30]. Additionally, to guarantee safe turbine operations, the platform motions induced by the offshore environmental loads need to be restrained to an acceptable limit [
31].
Accordingly, floating wind turbines are complex systems subject to coupled hydro-aerodynamic loads, and a better understanding of the behavior of these machines, which is substantial for their efficient design, requires the accurate evaluation of the nature and order of these coupled loading on the structure. The design of a cost-effective FWTS is a major obstacle to the industrialization of this technology [
21]. Cost reduction can be achieved by optimizing the floating platform’s size and weight, along with mooring, anchors, and installation and maintenance processes [
5,
21,
32]. Further cost reduction, particularly the levelized energy cost, can be achieved by improving the power production by enhancing the platform’s motion, a function of platform geometry, control system, and mooring [
33]. This is only possible by selecting a method able to analyze accurately the coupled hydro-aero-structural dynamics of FWTSs in each design stage while incorporating control algorithms, such as power and pitch control [
34,
35]. Thus, a rather detailed overview of the basic assumptions, formulations, accuracy, and computational demand of the most common methods capable of evaluating the coupled behavior of FWTSs in various design stages is required. This provides the necessary knowledge to identify the most suitable method at each design stage. For this purpose, based on the review carried out in this paper, a wide range of methods with different fidelities, as shown in
Figure 4, can be used.
Generally, from right to left, the fidelity increases. The higher the method’s fidelity, the more closely it represents reality by producing more detailed information. However, this comes at the cost of more computational demand and less efficiency. Therefore, the choice of a method at each design stage is a compromise between the accuracy and computational cost.
The simplest methods, as shown on the extreme-right of
Figure 4, are the linear frequency-domain models with the lowest level of fidelity, which, due to their ability to provide results in a short time period, are suitable in the early conceptual design phases for sizing and optimization purposes [
36]. The hydrodynamics and aerodynamics of the rigid FWTS structure are evaluated by strip theory based on potential flow (PF) together with Morison’s Equation (ME) and a simplified blade element momentum (BEM) approach or tabular data, respectively. In this stage, in addition to structure sizing and static analysis, one must also optimize the response amplitude operators (RAOs) [
37]. QuLAF [
38] and SLOW [
39] codes are examples of low-fidelity frequency-domain tools.
Fully coupled nonlinear time-domain models, used at the basic design stages, are categorized as mid-fidelity models (see
Figure 4). These models are more accurate than the low-fidelity tools, which come at the cost of higher computational demand. The aerodynamic part is modeled using either a BEM, Generalized Dynamic Wake (GDW) model, or Free Vortex Wake (FVW) method. Note that, in the case of FWTSs, the large blade deformations combined with the platform motions may lead to an interaction of the blades with their own shed vortices, which is a violation of basic assumptions of BEM method [
29,
40]. In this regard, FVW is better capable of modeling these complex physics than the BEM method while maintaining the computational cost to an acceptable level. The hydrodynamic loads are modeled based on PF, ME, or a combination of them. Further, the structural dynamics is modeled using multi-body dynamics (MBD) formulations, where flexibility is considered in selected components, such as blades and towers. Due to the efficiency and accuracy offered by the MBD for slender bodies, it has turned into a common tool for wind turbine structural simulations [
41].
The mid-fidelity simulation tools based on nonlinear time-domain methods are used for the analysis of dynamic responses of FWTSs under a variety of operating and extreme conditions. These tools, through a time-domain analysis, make it possible to assess properly the fatigue and extreme loads impact on FWTSs in various operating and extreme conditions [
42]. These tools are also excellent for the control design of the floating offshore wind plant, which may be more complex for this technology as the objective function, in addition to power production optimization, can be extended to reduce the structural loads and platform motions [
43,
44,
45,
46,
47,
48,
49,
50,
51]. The OpenFAST software from NREL is an example of these tools, which can perform the real-time coupled hydro-aero-servo-elastic simulations for FWTSs [
52,
53].
Next, as presented in
Figure 4, are the high-fidelity methods on the basis of both finite element analysis (FEA) and computational fluid dynamics (CFD). These tools are employed at last design phases, where more in-depth investigations are required for extreme conditions, as well as intricate flow conditions, e.g., vortex detachment from heave plates [
42]. These models involve less modeling than the other categories, which makes them computationally more expensive. These models are also suitable for fine tuning of the design, as well as the calibration of the low- and mid-fidelity models [
42,
54,
55,
56,
57,
58]. CFD uses either large eddy simulation (LES) or Reynolds-averaged Navier-Stokes (RANS) equations or hybrid LES-RANS, to predict the aero-hydrodynamic loads acting on FWTSs. Contrary to BEM and FVW methods, which rely on external input of aerodynamic loads and semi-empirical corrections for considering 3D effects, dynamic stall etc., CFD naturally accounts for all these effects. CFD can also better capture the turbulent wake behind a turbine, which leads to transient loads on the downstream machines in a wind farm.
Further, the FEA can better capture complex blade deformations at a higher computational cost than the MBD. Structural failure commonly emanates from local stress raisers, which the MBD cannot capture properly. Despite that, MBD can provide reasonable boundary conditions for the local finite element analyses.
The high-fidelity tools are an attractive option for evaluating FWTS dynamics; however, the results of these tools depend on several factors, such as temporal and spatial resolutions, turbulence modeling, free surface modeling, etc. Accordingly, reliable measurements from physical testing, including hybrid testing, full physical testing, and field testing, as shown in
Figure 4, are required to gain more confidence in these numerical tools.
Therefore, the low-, mid-, and high-fidelity models, as shown in
Figure 4, are obtained by combining numerical models able to evaluate the structural dynamics, hydrodynamics, and aerodynamics of FWTSs at different levels of accuracy. As one advance in the design stages, more accurate and detailed methods are required.
The present paper aims to provide a detailed overview of the most common coupled methods, as depicted in
Figure 4, including their basic assumptions, formulations, limitations, and costs, used for FWTSs, mainly those supported by a semisubmersible, to assist the choice of the most suitable method at each design stage of FWTSs. Note that several review papers are available about the hydro-aero-structural dynamic evaluation of FWTSs [
5,
8,
25,
30,
42,
59,
60,
61,
62].
In this regard, experimental testing and numerical models employed to assess FWTSs are reviewed by [
8]. However, full physical testing, field testing, numerical methods such as FVW methods, structural dynamics, and nonlinear second-order wave loads are not covered. An overview of the numerical methods, as well as the physical and hybrid tests, are covered by [
42]. However, the field tests, as well as the physical basis and formulations of the numerical models, are not presented. An overview of the coupled numerical models for optimizing an FWTS structure in conceptual, basic, and detailed design stages is given by [
5]. An overview of the coupled numerical tools and full physical tests for FWTSs is presented by [
25]. A review of the experimental and numerical methods is presented by [
59], without giving any detail about numerical methods. A general view of the numerical and physical models is given by [
30]. The physical and hybrid model tests is reviewed by [
60,
61,
62].
The present review is built upon the previous review efforts by giving a more detailed and up-to-date overview of the physical and numerical coupled modeling techniques for FWTSs, as presented in
Figure 4.
The paper is organized as follows:
Section 2 presents the coupled CFD-FEA modeling of FWTSs. Afterward, the nonlinear coupled time-domain methods are described in
Section 3. This is followed by the presentation of the linear frequency-domain models and mooring systems in
Section 4 and
Section 5, respectively. Moreover, some final notes about numerical models are given in
Section 6, and after that
Section 7 focuses on physical testing. Finally, a summary and conclusions are given in
Section 8.
5. Mooring system
A mooring system serves as a station-keeping system for a floating platform. Proper implementation of the mooring system is crucial for maintaining the wind turbine nearly fixed in its position in such a manner to guarantee the safe operation of the machine.
Generally, there are two types of mooring line systems: catenary and taut, as shown in
Figure 17.
In the case of the catenary system, the station keeping is performed through the weight of the portion of the mooring lines in contact with the seafloor [
21,
310]. In the case of the taut system, this is performed by the high tension in the cables. In other words, while the mooring line weight provides the restoring force in the catenary mooring system, the elasticity of the mooring line provides this force in the taut mooring system.
The catenary mooring lines are typically made of steel chains and/or wires [
311,
312], while the taut mooring lines are made of synthetic fibers, such as polyester, nylon or wires [
312,
313,
314,
315,
316,
317]. Recent studies have shown that a taut system composed of nylon may significantly reduce the cost of the floating wind turbine system, especially for shallow water applications, compared to the same system with polyester [
313,
315].
It is also possible to use a combination of materials for catenary applications (the semi-taut mooring system) to reduce the weight and cost of the mooring system [
318]. The chain is typically used to adjust the length and tension of the upper portion, where the mooring line is connected to the floater at the fairlead. Again, the chain is preferred due to its stiffness at the seabed, where the mooring line rests on the sea floor. Synthetic fibers, nylon, or wire can be used for cost and weight reduction. The Hywind project, for instance, uses a combination of steel chain and wire [
318].
The catenary configuration is usually applied to the floating platforms at water depths between 100 m to 250 m, where it can be considered an optimum solution for station keeping of the system [
319]. Based on
Figure 18, which shows the cost evolution of a traditional steel chain catenary system regarding water depth, in the case of shallow waters with a depth smaller than 100, the highly dynamic condition in which the floater operates calls for a more expensive station-keeping system [
320]. Additionally, the additional mooring line length increases the cost as the water depth exceeds the values larger than 250 m.
On the other hand, the taut configuration can be used in water depths smaller than 100 m or larger than 250 m [
320]. An example of a taut system used at a water depth shallower than 100 m is the Ideol platform installed close to the French coastline at a water depth of nearly 30 m [
321].
A floating platform equipped with a catenary system can have some horizontal movement. However, a floating platform with a taut system, typically has limited horizontal movement. The disadvantages of the catenary system are its more seabed disruption and larger footprint. On the other hand, as the anchor experiences relatively less loads, the catenary system has more economical anchors. Considering a taut system, the disadvantage is the higher cost of installation and anchoring due to the large horizontal and vertical loads acting on the anchor [
27,
322]. However, its lower seafloor disruption, smaller footprint, and suitability for shallow and deep waters are the main advantages of a taut system.
In the case of catenary systems, to increase the mooring line tension and, consequently, the restoring force acting on the floating platform, especially in the case of shallow waters, the clump weights can be used, as shown schematically in
Figure 19. It is also possible to reduce the mooring line dynamics and weight effects on the platform in the case of deep waters using buoyancy modules, as shown in
Figure 19.
One main challenge related to floating wind turbines is the high cost of anchoring for a single floating platform. An alternative to reduce these costs is using shared mooring or shared anchoring systems [
324,
325,
326].
However, the use of shared mooring lines is accompanied by a number of challenges as follows [
327,
328]:
The installation process is more complex.
Difficulty in towing a single floating wind turbine to a nearby port when it is necessary for operation and maintenance.
complexity in a mooring line may increase to a point where it is exposed to failure risk.
The Natural period of several mooring lines may have coupled effects with their surrounding floating wind turbines.
In the event of a mooring line or anchor failure, this scenario may affect several turbines.
The use of shared anchors is only beneficial under certain situations, which calls for a site-specific investigation [
328]. Saving due to shared anchors can be lost due to an increase in the length of mooring lines to respect the minimum spacing between turbines, which depends highly on the energy yield calculations of the specific site. Shared anchors are more feasible for sites where it is possible to have smaller turbines’ spacing [
327].
The coupled analysis of a floating wind turbine system is generally accompanied by a mooring line modeling, which is performed using either a quasi-static or dynamic method.
5.1. Quasi-static method
The quasi-static models can be as a simple linear stiffness matrix to more complex catenary equations, where Newton’s force equation is solved for each node of connection [
329]. These models consider that the mooring is in the balance between the anchor and the FWTS attachment point [
30]. Additionally, the dynamic effects, such as inertia, hydrodynamic drag force, and vortex shedding, are not considered [
65,
66,
67,
75,
82,
182].
Figure 20 illustrates schematically how the responses of a mooring line can be affected when dynamic effects are ignored. As can be seen, although the mooring line close to the fairlead tends to follow the floating platform motions, the lower portion of the mooring line may respond with a delay to this displacement. This can be regarded as a violation of the quasi-static assumption, which considers that the line is in static equilibrium at each time step [
330].
The quasi-static models tend to underestimate the restoring load, particularly for extreme sea states [
42,
331,
332], and as shown by [
333] they are unable to estimate reliably the fairlead load amplitudes.
5.2. Dynamic method
Considering or neglecting the dynamic effects of mooring lines for FWTSs is a hot topic in the hydrodynamics of these structures. Due to the smallness of mass and motions, the dynamics may be neglected in the case of mooring systems used in shallow waters. However, it is crucial to consider the dynamic effects of deep water conditions. The oil and gas platform and vessel demonstrate the importance of considering line dynamics for water depth larger than 150 m [
180]. Further, using dynamic models is essential for reliable estimation of the ultimate and fatigue loads in a mooring line and correct modeling of the peak tension in extreme events [
30,
173].
The dynamic models are based on either lumped-mass approach (similar to MBD) [
186,
195,
334,
335,
336], FEA [
182,
183,
299,
330,
299] or finite difference (FD) method [
337]. The hydrodynamic drag and added mass are considered for all the elements resulting from the line discretization.
The simplest model is the lumped mass, where the mooring line model is a set of compact masses connected through massless spring-damper elements [
195]. A sensitivity study may be required to determine the number of multibodies beyond which the results remain unchanged. The drawback of this method is that it neglects the mooring line torsional stiffness [
333,
338]. Due to its simplistic nature, the lumped-mass method is available in many simulation tools of FWTSs, such as OpenFAST and OrcaFlex [
180]
A higher fidelity solution can be achieved using either the FD or FE methods. However, the results obtained from lumped mass, FD, and FE methods are expected to be very similar under sufficient resolution [
30,
333,
339].
6. Final notes on numerical models
Table 1 summarizes the tools used to perform the coupled hydro-aero-structural analysis of floating wind turbines based on low- and mid-fidelity codes.
As can be inferred, most of these tools use BEM with various corrections to compute the aerodynamic loads. The main difference between the BEM, GDW, and FVW methods is their consideration of the dynamic inflow condition. A correction referred to as a dynamic inflow model is typically used within the BEM method for the better representation of the aerodynamic responses under unsteady conditions emanating from changes in incoming wind speed, rotor speed, and blade pitch angle, as well as platform motions. However, the dynamic inflow is directly considered by the GDW method. A limitation of the GDW method is the assumption of the smallness of the induced velocities compared to the free flow, which creates instability at low wind speeds. Finally, the FVW method is able to account for the dynamic inflow due to its basic formulation.
In this regard, the Offshore Code Comparison Collaboration (OC6) evaluates the validation of 54 aerodynamic numerical models for a floating wind turbine subject to large harmonic support platform motions in both surge and pitch directions [
340]. The validation uses wind tunnel measurements from a 1:75 scaled DTU 10 MW wind turbine. A wide range of modeling approaches, such as BEM, dynamic BEM (DBEM) (which considers the effects of dynamic inflow), GDW, FVW, and blade-resolved/actuator-based CFD methods, are evaluated. The tests are performed at a wind speed of 4.19 m/s and an angular velocity of the rotor of 240 rpm, equivalent to the scaled model’s rated condition. The measured and calculated torque moment and thrust force of the turbine during one harmonic surge motion of the platform with a frequency of 1 Hz and an amplitude of 0.035 of rotor diameter are presented in
Figure 21. Note that the numerical results are median values for each modeling technique. As can be inferred, the predictions provided by the FVW are comparable to those provided by CFD and even better in the case of thrust force. The poorest predictions are produced consistently by the BEM and DBEM methods.
Regarding the hydrodynamics, as shown in
Table 1, either the Morison equation or potential flow theory or a combination of both is employed. Second-order potential flow theory based on QTFs is employed to account for the fundamental second-order effects. Considering these effects is crucial because a moored floating platform, unlike an unrestricted platform, has natural frequencies in the horizontal plane, generally outside the frequency range of the incident waves. The natural frequencies in the horizontal plane are excited by the low second-order frequencies emanating from the second-order potential problem. As shown in the OC5 Phase II validation project [
173], underestimation of low-frequency loads can lead to a 20% underestimation of the loads and motions of a floating structure. This underestimation remarkably affects the correct estimation of the fatigue and ultimate loads acting on the floating structure. As suggested in the project of OC6 Phase I, improvement in the low-frequency predictions can be achieved by tuning the QTFs using data from physical tests or high-fidelity CFD simulations [
203].
Accordingly, a RANS-based CFD model, together with the VOF method, is used by [
57] to fine-tune the QTFs of an FWTS obtained from potential flow theory for a group of bichromatic regular waves, with a frequency difference correspondent to the natural frequency of the FWTS in surge or pitch direction. In this regard,
Figure 22 shows the phases and magnitudes of the QTFs related to the force in surge direction and the moment in pitch direction obtained from both CFD (new QTF) using the commercial finite volume code OpenFOAM and potential flow theory (old QTF) using the commercial code WAMIT. As can be seen, a considerable difference exists between the predictions; regarding the QTFs related to the force and moment magnitudes, this difference increases even more at higher frequencies, while regarding the phase, there is consistently a substantial difference between the predictions with CFD showing nearly an opposite trends for phase angle when compared with potential flow theory.
The frequency-dependent characteristics of a floating platform are obtained most commonly using either AQWA [
177,
186,
195,
196,
199] or WAMIT [
198]. As shown by [
195]. Generally, a good agreement exists between the two codes. AQWA can perform the time-domain simulations [
177,
186]. However, one needs to specify how the aerodynamic loads are computed. These loads can be either calculated using simple expressions within the software or through coupling with other codes such as OpenFAST.
As seen in
Table 1, the structural flexibility is accounted for through an MBD formulation, where flexibility is considered by either a modal method or finite element representation. The flexibility is generally applied to the wind turbines, and substructures are typically assumed to be rigid.
Note that in addition to the coupled numerical models presented so far, the coupled dynamic analysis of FWTSs may be carried out through other coupling schemes as follows:
An aeroelastic model is developed by [
341] by combining the CFD-based AL method with a beam solver to evaluate the couple aero-structural responses of the NREL 5MW wind turbine in both isolation and wakefield of an upstream wind turbine. It is shown that considering the aeroelastic deformations leads to an aggravation of the tower shadow effect.
A CFD solver based on URANS equations with a
model coupled with an FEA code based on non-linear Timoshenko beam elements is used by [
143] to investigate the aeroelastic behavior of the reference 10MW DTU wind turbine at various uniform wind speeds [
342]. [
343] evaluates the NREL 5 MW wind turbine’s aeroelastic behavior by coupling a URANS-based CFD model with a geometrically exact beam formulation FEA code under harmonic platform surge motions. The turbulence model in CFD is the
SST model.
[
107] proposes an aero-elastic model that considers the drivetrain dynamics through coupling the Hybrid RANS-LES CFD model responsible for the flow field prediction with an MBF. The model is applied to the NREL 5 MW wind turbine to investigate the interaction between turbine aerodynamics, flexible blades, and drivetrain dynamics. A coupled hydro-aero-elastic model is developed by [
75] to simulate the OC4 semisubmersible FOWT under a combined wind/wave operating condition. The hydrodynamic model is based on RANS with
SST model and VOF method, and the structural model is based on an MBF. An overset grid CFD solver based on the DDES model derived from the
SST coupled with an MBD solver is used by [
35] to predict the aeroelastic behavior of the floating NREL 5 MW wind turbine subject to an ABL. Blades and tower are considered as flexible one-dimensional structures.
- -
- coupled CFD-BEM-MBD models [
344]:
A CFD model, which is responsible for the platform hydrodynamics, is coupled to the aero-servo-elastic OpenFAST code by [
344] to simulate the OC4 DeepCWind FOWT dynamic responses under various operating conditions. More accurate results are obtained using this coupled tool compared to the OpenFAST alone simulations.
- -
- coupled CFD-PF models [
79]:
[
79] proposes a coupled fluid-structure interaction model based on LES to study FWTS motions in waves. The domain is divided into two parts: near-field, where a two-phase LES solver based on the level set method is used, and a far-field model, where an aerodynamic model based on LES without viscosity is combined with a hydrodynamic model based on a fully nonlinear-potential-flow-theory. The modeling of the subgrid-scale turbulence over the near-field region is performed using the dynamic Smagorinsky model, while the scale-dependent Lagrangian dynamic model is used over the far-field region.
- -
- coupled CFD-PF-MBD models [
345,
346]:
The wakefield of the OC4 semisubmersible FWTS is compared against its fixed-bottom counterpart under ABL condition by [
345]. The wind turbine rotors are represented using an AL method based on OpenFAST code, where the wind flow is simulated using the SOWFA LES solver from NREL with standard Smagorinsky model responsible for subgrid-scale turbulence modeling [
346]. The results indicate a small deviation between the wake patterns of the FWTS and its fixed-bottom counterpart.
- -
- coupled PF-FEA models [
347,
348]:
[
347] analyzes the ship-FWTS collision using a coupled linear PF-FEA model. The aerodynamic thrust force is considered to be a point load acting at the hub based on the thrust curve of the 5MW NREL wind turbine. The linear diffraction theory and an FEA model are combined by [
348] to evaluate the hydro-elastic behavior of a novel triangular floating platform capable of accommodating three wind turbines on its vertices.
8. Summary and Conclusions
Various methods with various fidelity levels are available to analyze the FWTS coupled hydro-aero-structural dynamics at each design stage.
The linear models in the frequency domain with the lowest level of fidelity are suitable in the early conceptual design phases for dimensioning and optimizing the FWTSs.
At the preliminary design stages, the mid-fidelity fully coupled nonlinear time-domain models are used to analyze the dynamic responses of FWTSs under various conditions to assess the operating, fatigue, and extreme loads correctly. These models are also excellent for the control design of FWTSs to reduce the structural loads and platform motions. The complex aerodynamic behavior of FWTS is better captured by the FVW method than by the BEM method. This is because the FVW method can characterize the wake evolution of a wind turbine over time, which may give more reliable predictions in the case of FWTSs due to the possibility of rotor movement into its own wake. Additionally, the accuracy of hydrodynamic modeling is increased by accounting for second-order wave loads using full sum- and difference-frequency QTFs. The QTFs can be calibrated using data from physical tests or high-fidelity CFD simulations to enhance the low-frequency predictions further. For the dynamic structural assessment of FWTSs, using nonlinear beam theories is preferred over the modal analysis in the case of large blade deformations.
The high-fidelity coupled CFD-FEA methods are used at the last design phases, where more in-depth investigations are required for extreme conditions and intricate flow conditions. These methods are suitable for fine-tuning the design, as well as the calibration of the low- and mid-fidelity models. However, due to the high computational demand of a coupled CFD-FEA method, the high-fidelity analyses of FWTSs are mainly restricted to aeroelastic and hydro-aerodynamic studies. In this regard, little attention has been given to the impact of the near-wall grid resolution on CFD simulations of FWTSs. It is common to couple high-fidelity CFD and FEA methods with low- and medium-fidelity models such as PF, MBD, or BEM to reduce the computational demand.
Modeling the mooring system is also part of all numerical modeling of FWTSs. In this regard, during a coupled analysis, the dynamic effects of mooring cables for floating wind turbine systems must be accounted for through the use of dynamic models such as lumped mass, FEA, and FD methods for a more reliable estimation of the ultimate and fatigue loads in a mooring line and correct modeling of the peak tension in extreme events.
Finally, the physical tests are used to validate the numerical models to gain more confidence in these numerical tools. However, the shortage of full-scale field measurements for FWTSs leads to a partial validation study of simulation tools. The physical tests also enhance our understanding of the physical aspect of the problem by revealing complex phenomena that may only be recognized using physical modeling techniques. This enhanced understanding eventually leads to an improvement in the low- and mid-fidelity numerical models. In the full physical testing to solve the issue of Reynold-number dissimilarity between the model and prototype, several solutions are outlined. Hybrid testing is an intriguing option to solve this issue, which can be categorized into numerically modeled aerodynamics and hydrodynamics.
Finally, based on the review carried out in this article, the following possible future research is identified to address challenges in modeling FWTSs dynamics that persist to date:
The accuracy of the mid-fidelity tools can be improved by incorporating the effect of vortex-induced vibration (VIV). This effect originates from currents or low-frequency waves and results in time-varying loading emanating from pressure fluctuations owing to vortex shedding from the platform components or mooring lines. The problem arises once the vortex shedding occurs at a frequency close to the structure’s natural frequency, which leads to the resonation of the two frequencies and large oscillation amplitudes.
The computational cost of CFD simulations of an FWTS in waves can be reduced by coupling a fully nonlinear potential solver capable of describing the nonlinear, three-dimensional wave field with CFD solvers governing the flow field in the FWTS vicinity. Using this strategy, it is possible to account for inherently nonlinear second-order effects.
A more systematic study is required to assess the impact of different materials for mooring lines on the platform’s motions and overall power production.
Due to wave second-order effects, more research is required to evaluate the structural fatigue damage for ultra-large FWTSs.
Modeling slender offshore structures in waves as rigid might lead to overestimating fluid forces. For example, modeling an offshore wind turbine foundation as completely rigid, commonly adopted in numerical and experimental analyses, may give very conservative wave forces. The size of offshore wind turbines has grown over the last few years, but the size of the platforms used to support these huge machines is also increasing. It may be crucial to consider the structural dynamics (hydro-elasticity) of these platforms in numerical and experimental studies, which are generally not considered.
During numerically modeled aerodynamics hybrid tests, the aerodynamic loads are typically calculated using a BEM method, which cannot accurately capture the effects of unsteady loading, dynamic inflow, and turbine-wake interactions. However, this issue can be avoided by using an FVW method.
One source of error in hybrid tests comes from the numerical models used. One solution to this problem is to avoid using these numerical models by coupling the two types of hybrid tests for an FWTS model via an Internet connection.
To gain confidence in the numerical and experimental results, detailed verification and validation studies are essential in quantifying errors and uncertainties. It is also crucial to identify the limitations of the hybrid testing by using the same FWTS model in both methods, i.e., numerically modeled aerodynamics and hydrodynamics. Further, in the case of CFD and FEA, sensitivity studies are required to test different element sizes, spatial and temporal discretization schemes, turbulence models, mooring line models, etc.
Artificial intelligence can be used to substitute the numerical subsystem in hybrid testing, thus increasing the efficiency of this type of test.
Field measurements about the impact of FWTS motions on their aerodynamic performance and power generation are crucial, especially for validating the numerical models and understanding where improvements are essential for these models.
Figure 1.
Some popular semisubmersible configurations
Figure 1.
Some popular semisubmersible configurations
Figure 2.
Two examples of semisubmersible platforms with three offset columns
Figure 2.
Two examples of semisubmersible platforms with three offset columns
Figure 3.
An FWTS subject to loads emanating from several sources, such as waves, current and wind
Figure 3.
An FWTS subject to loads emanating from several sources, such as waves, current and wind
Figure 4.
Methods for evaluating the coupled hydro-aero-structural dynamic behavior of FWTSs
Figure 4.
Methods for evaluating the coupled hydro-aero-structural dynamic behavior of FWTSs
Figure 5.
Free surface modeling in CFD
Figure 5.
Free surface modeling in CFD
Figure 6.
From left to right: wind turbine geometry, actuator disk, actuator line and actuator surface
Figure 6.
From left to right: wind turbine geometry, actuator disk, actuator line and actuator surface
Figure 7.
Percentage of CFD software used to study the coupled behavior of an FWTS
Figure 7.
Percentage of CFD software used to study the coupled behavior of an FWTS
Figure 8.
Percentage of turbulence models used in CFD codes
Figure 8.
Percentage of turbulence models used in CFD codes
Figure 9.
Percentage of methods used to model the turbulent boundary layer close to the blades: wall functions (WF), without WF, and actuator methods
Figure 9.
Percentage of methods used to model the turbulent boundary layer close to the blades: wall functions (WF), without WF, and actuator methods
Figure 10.
Linearized boundary conditions typically used in the linear potential flow theory
Figure 10.
Linearized boundary conditions typically used in the linear potential flow theory
Figure 11.
The radiation problem related to an FWTS
Figure 11.
The radiation problem related to an FWTS
Figure 12.
velocity potential related to the diffraction of the incident waves to an FWTS
Figure 12.
velocity potential related to the diffraction of the incident waves to an FWTS
Figure 13.
Elements along the blade span located at the radial position of r, which extend to affect the airflow momentum that travels across the circular ring area covered by these rotating elements due to their aerodynamic loads.
Figure 13.
Elements along the blade span located at the radial position of r, which extend to affect the airflow momentum that travels across the circular ring area covered by these rotating elements due to their aerodynamic loads.
Figure 14.
angle of attack, pitch angle, angle enclosed by the rotor plane and the total incident wind speed W
Figure 14.
angle of attack, pitch angle, angle enclosed by the rotor plane and the total incident wind speed W
Figure 15.
FVW method explanation
Figure 15.
FVW method explanation
Figure 16.
Multi-body representation of a wind turbine
Figure 16.
Multi-body representation of a wind turbine
Figure 17.
Mooring line systems: catenary and taut
Figure 17.
Mooring line systems: catenary and taut
Figure 18.
Cost evolution of a traditional steel chain catenary system regarding water depth [
319]
Figure 18.
Cost evolution of a traditional steel chain catenary system regarding water depth [
319]
Figure 19.
A catenary mooring system with clump weights and buoyancy modules [
323]
Figure 19.
A catenary mooring system with clump weights and buoyancy modules [
323]
Figure 20.
Possible response of a mooring line under quasi-static and dynamic assumptions
Figure 20.
Possible response of a mooring line under quasi-static and dynamic assumptions
Figure 21.
The measured and calculated (using a wide range of modeling approaches, such as BEM, DBEM, GDW, FVW, and CFD) torque moment and thrust force of a scaled turbine model during one harmonic surge motion of the platform with a frequency of 1 Hz and an amplitude of 0.035 of rotor diameter (Figure reproduced using the data given by [
340])
Figure 21.
The measured and calculated (using a wide range of modeling approaches, such as BEM, DBEM, GDW, FVW, and CFD) torque moment and thrust force of a scaled turbine model during one harmonic surge motion of the platform with a frequency of 1 Hz and an amplitude of 0.035 of rotor diameter (Figure reproduced using the data given by [
340])
Figure 22.
The phases and magnitudes of the QTFs related to force in surge direction and moment in pitch direction obtained from both CFD and potential flow theory using the commercial code WAMIT (Figure reproduced using the data given by [
57])
Figure 22.
The phases and magnitudes of the QTFs related to force in surge direction and moment in pitch direction obtained from both CFD and potential flow theory using the commercial code WAMIT (Figure reproduced using the data given by [
57])
Figure 23.
Numerically modeled aerodynamics hybrid tests
Figure 23.
Numerically modeled aerodynamics hybrid tests
Figure 24.
numerically modeled hydrodynamics hybrid tests
Figure 24.
numerically modeled hydrodynamics hybrid tests
Table 1.
Summary of the capabilities of coupled numerical tools
Table 1.
Summary of the capabilities of coupled numerical tools
Code |
Hydrodynamics |
Aerodynamics |
Structural dynamics |
Mooring system dynamics |
WAMIT |
Potential flow theory (Frequency domain) |
- |
Rigid Body/ Modal |
Quasi-static method |
AQWA |
Potential flow theory (Frequency/time domain) |
- |
Rigid body/ FEA |
Quasi-static/ Dynamic methods |
WINDOPT |
Potential flow theory (Frequency domain) |
- |
Rigid body |
Quasi-static/ Dynamic methods |
OpenFAST |
Potential flow theory/ Morison Equation (Time domain) |
BEM, GDW and FVW methods |
MBD, modal/FEA |
Quasi-static/ Dynamic methods |
BLADED |
Potential flow theory/ Morison Equation (Time domain) |
BEM method |
MBD, modal/FEA |
Quasi-static/ Dynamic methods |
OrcaFlex |
Potential flow theory/ Morison Equation (Time domain) |
BEM method |
MBD, modal/FEA |
Quasi-static/ Dynamic methods |
3DFloat |
Potential flow theory/ Morison Equation (Time domain) |
BEM method |
FEA |
Quasi-static/ Dynamic methods |
HAWC2 |
Potential flow theory/ Morison Equation (Time domain) |
BEM method |
MBD, modal/FEA |
Quasi-static/ Dynamic methods |
SIMA (SIMO/ RIFLEX) |
Potential flow theory/ Morison Equation (Time domain) |
BEM method |
MBD, modal/FEA |
Quasi-static/ Dynamic methods |
QBlade |
Potential flow theory/ Morison Equation (Time domain) |
BEM and FVW methods |
MBD, modal/FEA |
Quasi-static/ Dynamic methods |
Sesam/ Wadam |
Potential flow theory (Frequency domain) |
- |
Rigid body |
Quasi-static method |
Simpack |
Potential flow theory/ Morison Equation (Time domain) |
BEM, GDW, and FVW methods |
MBD, modal/FEA |
Quasi-static/ Dynamic methods |
SLOW |
Reduced potential flow theory/ Morison Equation (Time domain) |
Actuator model |
MBD, modal |
Quasi-static method |