Next Article in Journal
Distributed Secondary Control in Microgrids Using Synchronous Condenser for Voltage and Frequency Support
Next Article in Special Issue
Effect of Gaskets Geometry on the Performance of a Reverse Electrodialysis Cell
Previous Article in Journal
An Accurate Switching Transient Analytical Model for GaN HEMT under the Influence of Nonlinear Parameters
Previous Article in Special Issue
Performance Comparisons of Three-Phase/Four-Wire Model Predictive Control-Based DC/AC Inverters Capable of Asymmetric Operation for Wave Energy Converters
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A Compendium of Formulae for Natural Frequencies of Offshore Wind Turbine Structures

by
Ramon Varghese
1,2,
Vikram Pakrashi
1,2,* and
Subhamoy Bhattacharya
3
1
UCD Centre for Mechanics, Dynamical Systems and Risk Laboratory, School of Mechanical and Materials Engineering, University College Dublin, D04 V1W8 Dublin, Ireland
2
SFI MaREI Centre UCD, UCD Energy Institute, University College Dublin, D04 V1W8 Dublin, Ireland
3
Department of Civil and Environmental Engineering, University of Surrey, Guildford GU2 7XH, UK
*
Author to whom correspondence should be addressed.
Submission received: 1 March 2022 / Revised: 23 March 2022 / Accepted: 30 March 2022 / Published: 18 April 2022
(This article belongs to the Special Issue Marine Renewable Energy Technology)

Abstract

:
The design of an offshore wind turbine system varies with the turbine capacity, water depth, and environmental loads. The natural frequency of the structure, considering foundation flexibility, forms an important factor in structural design, lifetime performance estimates, and cost estimates. Although nonlinear numerical analysis in the time domain is widely used in the offshore industry for detailed design, it becomes necessary for project planners to estimate the natural frequency at an earlier stage and rapidly within reasonable accuracy. This paper presents a compendium of mathematical expressions to compute the natural frequencies of offshore wind turbine (OWT) structures on various foundation types by assimilating analytical solutions for each type of OWT, obtained by a range of authors over the past decade. The calculations presented can be easily made using spreadsheets. Example calculations are also presented where the compiled solutions are compared against publicly available sources.

1. Introduction

Offshore wind power is a potential sustainable solution to the growing energy needs of the industrialized world. With rapid expansion in Europe and Asia, and reducing installation and operating costs, wind power is speculated to become the cheapest form of green energy. The huge initial capital investment cost is, however, one of the primary challenges faced by the offshore wind market. Cost estimation in offshore wind hugely relies on the design of the turbine structure, which is highly site-specific and involves multiple iterative design and analysis steps. In most cases, the turbine system and the tower structure can be identified early but the foundation system retains a degree of uncertainty owing to the site conditions and potentially unreliable design guidelines. Typically costing 16–34% of the total cost of a wind farm project, the foundation system becomes a controlling factor [1]. The choice of a foundation in turn depends on factors such as the water depth, fabrication, installation, operation and maintenance, decommissioning, and economics.
The offshore wind turbine (OWT), due to its shape and form, is a dynamically sensitive system [2]. An offshore wind turbine essentially consists of the Rotor Nacelle Assembly (RNA), tower, substructure, and the foundation. The size of OWTs has been continuously increasing since the 2000s, and larger turbines tend to have lower natural frequencies. In most cases, the dominant wave frequencies are close to the natural frequencies, where dynamic design becomes important. The target natural frequency of the chosen turbine system at a location depends on the operating range of the turbine as well as the mass and stiffness of the turbine-foundation system.
The foundation stiffness, which may have group effects, forms an important input parameter for natural frequency estimation. Several researchers have proposed closed-form expressions for the natural frequency of OWTs considering soil–structure interaction (SSI) effects. Considerable attention has been given to the vibration characteristics of wind turbines on monopile foundations [3,4,5,6,7,8,9,10]. Closed-form solutions of Eigen frequencies and design steps of monopile-supported OWTs have been reported by Arany et al. [11,12,13]. The dynamic stiffness of rigid surface foundations available from previous works [14,15,16,17,18,19] can be employed to estimate the flexible-base natural frequency of gravity-based foundation wind turbines. A solution for the first natural frequency of OWT, supported by multiple jacket foundations, was discussed by Jalbi et al. [20]. This method can be further enhanced by adopting the group-effect correction factors reported by [21,22]. Ryu et al. [23] presented a simple closed-form solution for OWTs with a tripod suction caisson foundation. For floating wind turbines, the structural responses, such as nacelle surge motion and bending moments at the tower base, are dominated by rigid body motions rather than elastic deformations [24]. Numerical methods have been largely used in studies on floating wind turbines [25,26,27,28,29,30,31,32,33,34]. Simplified expressions for spreadsheet calculation of rigid body natural frequencies of tension-leg-platform-supported OWT have been reported by [35]. For the case of spar-buoy-type wind turbines, the analytical procedure presented by Ye and Ji [36] can be used to estimate the natural frequency. Studies on semisubmersible-type wind turbines have been largely based on experimental or numerical methods [37,38,39,40].
It is in the best interest of project planners and designers that a single point reference of formulae for the foundation stiffness and natural frequencies of the various types of OWTs is readily available. Simplified analytical solutions help in choosing the tower stiffness or foundation stiffness. Such estimates can also help in understanding the output from a numerical analysis. The main objective of the present paper is to provide a compendium of formulae to estimate the undamped natural frequency of OWTs in various modes, considering SSI wherever applicable. The formulae for each type of offshore wind turbines are compiled. Example calculations are also added for clarity.

2. Offshore Wind Turbine Structure and Vibration Characteristics

2.1. The Structure of an Offshore Wind Turbine

The main components of a wind turbine include the tower, foundation, nacelle, rotor blades, and the hub as shown in Figure 1. Most offshore wind turbines are characterized by a three-bladed turbine driving a horizontally mounted generator. The nacelle houses the generator and the gear box. The tower structure supports the weight of the RNA as well as the heavy lateral wind loads. Tubular steel towers are widely used in the offshore wind power industry. The operational rotor frequency range of a turbine is called the 1P range. For example, the NREL 5 MW turbine has a 1P range of 6.9 to 12.1 rpm, essentially meaning that the turbine operates at this frequency during its entire lifecycle. The foundation of an OWT is an important design consideration that often controls the financial viability of an offshore wind farm. The foundation engineer considers factors such as seabed conditions, installation related aspects, and environmental regulations before selecting a foundation type. The foundation design can also influence the target natural frequency of the whole system, which needs to avoid the 1P range. Simplified models often idealize the tower as a beam, and the RNA as a lumped mass with or without a rotational inertia. The foundation is often idealized as translational and rotational springs at the base. Table 1 presents the typical parameters of the tower–turbine system that are required for natural frequency estimates.
Typically, the rotor-nacelle assembly (RNA) and the tower are supplied by the turbine manufacturers as standard units, while the foundation design is site-specific. Design is often a collaborative effort between the turbine manufacturer and the substructure designer. Table 2 presents the characteristics of the NREL 5 MW, LW 8 MW, DTU 10 MW, and Haliade-X 12 MW turbines.

2.2. OWT As a Dynamical System

The OWT system can be idealized using simple structural models to estimate its dynamical characteristics. The tower–RNA system can be, for all practical purposes, idealized as a vertical beam with a lumped mass at its end. Often, an OWT falls under the category of a high-slenderness–low-stiffness dynamical system [9,41]. Simplified models simulate the flexibility of the foundation–soil system using springs representing the various degrees of freedom. The design ensures that the system frequency lies outside of the wind and wave frequencies as well as the operating range of the turbine.
The coordinate system and general rigid-body degrees of freedom used in this article are shown in Figure 2. The modes of vibration of a wind turbine structure essentially depend on the combination of foundation system and the superstructure’s stiffness and mass distribution [1]. The fundamental modes of vibration for grounded OWT systems can be mainly of two types: sway-bending mode and rocking mode. The sway-bending mode consists of flexible modes of the tower–RNA system and is dominant in cases where the foundation has sufficient axial stiffness in comparison to the tower. The rocking mode is dominant when the foundation is axially deformable (less stiff), such as in the case of an OWT supported on multiple shallow foundations. Schematic diagrams showing the modes of vibrations in shallow-embedded foundations and deep-embedded foundations are presented in Figure 3a,b, respectively. Floating OWTs, on the other hand, exhibit surge, sway, heave, roll/bending, and pitch/bending modes.

2.3. Classification of OWT Systems

Offshore wind turbines can be broadly classified based on their substructures as grounded systems and floating systems. Grounded systems, for which the structure is anchored to the seabed, are planned for water depths below 60 m. For water depths below 30 m, monopiles, gravity-based foundations, and suction caissons are currently being used. For water depths between 30 m and 60 m, jackets or seabed frame structures supported on piles or caissons are either used or planned. Suction caissons or piles can also be used in groups with the support of tripod or jacket structures. These scenarios involve group effects in addition to soil–structure interaction (SSI). Typically for water depths more than 60 m, floating systems like tension-leg platforms (TLP) and spar-buoy floating concepts are employed where the tower–RNA system is allowed to float and is anchored to the seabed by a mooring system. Figure 4 illustrates the various types of foundations used today for different depths of water.
For water depths below 30 m, monopiles, gravity-based foundations, and suction caissons are currently being used. For water depths between 30 m and 60 m, jackets or seabed frame structures supported on piles or caissons are either used or planned. Monopiles are so far the most popular foundation choice for grounded systems. A floating system is usually considered for deeper waters, typically more than 60 m. Among floating systems, the semisubmersible type has been by far the most widely adopted worldwide. Despite its poorer fundamental stability in comparison to spar-buoy and TLP concepts, the semisubmersible greatly simplifies the transportation and installation procedure, thereby reducing costs.

3. Formulae for Stiffness and Natural Frequency

3.1. Wind Turbine Supported by Gravity Based Foundation

The gravity-based foundations for OWTs are designed to avoid uplift or overturning by providing an adequate dead load. Simple macro element models of the wind turbine structure can be employed to estimate the natural frequency, provided that the foundation is modelled by three springs: lateral, rotational, and cross-coupling springs. Arany et al. [11,13] presented an analytical method based on the Timoshenko beam model to compute the flexible-base natural frequency of OWT structures. The method represents the flexible-base natural frequency in terms of the fixed-base natural frequency multiplied by flexibility factors representing the substructure and foundation. This procedure can essentially be used for any foundation types that can be represented by the three-spring model as shown in Figure 5. In this paper, we adopt the recent work equivalent framework for foundation stiffness by [19] for layered soil stratum. The parameters required for the natural frequency calculation are presented in Table 3.
The steps involved in the estimation of natural frequency are:
  • Step 1. Calculate the average tower diameter, the average wall thickness, and the average tower diameter.
The average area moment of inertia is calculated as:
I T = 1 8 ( D T t T ) 3 t T π
where D b is the tower bottom diameter, and D t is the tower top diameter. The average wall thickness and the average tower diameter are calculated as:
t T = m T ρ T L T D T π D T = D b + D t 2
where ρ T is the density of the tower material (steel).
  • Step 2. Calculate the non-dimensional foundation stiffness values:
The non-dimensional foundation stiffness values are computed as:
η L = K L L T 3 E I η η L R = K L R L T 2 E I η   η R = K R L T E I η
where K L , K L R , K R are the stiffness parameters, and E I η is the equivalent bending stiffness of the tower, calculated as:
E I η = E I t · f q
where:
f q = 1 3 · 2 q 2 q 1 3 2 q 2 l n q 3 q 2 + 4 q 1 q = D b D t
The foundation stiffnesses are characteristics of a soil–foundation system. Simplified methods exist to estimate the foundation parameters in homogeneous and layered soils. For foundations mounted on soil surface, the cross-coupling stiffness term can be ignored for practical purposes. For strata that can be idealized as homogeneous, the formulae from various classical works collated in [42] can be adopted. For circular foundations in layered soil profiles, the recent method proposed by [19] can be adopted for a spreadsheet-based estimation. The procedure is described below:
  • Identify and tabulate the shear modulus and Poisson’s ratio for a depth not less than 20 times the foundation diameter/width.
  • Compute the weight distribution function as follows:
    w ^ = a b z ^ b a 1 e x p z ^ b a
    where z ^ is the depth normalized with the foundation diameter, and a and b are parameters defined in Table 4.
  • Evaluate the equivalent shear modulus using the expression:
    1 G e q = 0 1 G w ^ d z ^
  • Compute the foundation stiffnesses using the expressions:
    K L = 16 1 ν D 7 8 ν G e q
    K R = D 3 3 1 ν G e q
  • Step 3. Compute the foundation flexibility factors.
The foundation flexibility coefficients are given as:
C R η L , η R , η L R = 1 1 1 + a η R η L R 2 η L C L η L , η R , η L R = 1 1 1 + b η L η L R 2 η R
where a 0.5 and b 0.6 are empirical coefficients [9,10]. The substructure flexibility coefficient represents the flexibility of the substructure. The distance between the mudline and the bottom of the tower is L S , and E s I s is the bending stiffness of the substructure. The coefficient is expressed in terms of two dimensionless parameters, the bending stiffness ratio χ and the length ratio ψ :
C S = 1 1 + 1 + ψ 3 χ χ
χ = E T I T E s I s ψ = L S L T
  • Step 4. Compute the fixed-base natural frequency.
The fixed-base natural frequency of the tower is expressed simply with the equivalent stiffness k 0 and equivalent mass m 0 of the first mode of vibration as:
f F B = 1 2 π k 0 m 0 = 1 2 π 3 E T I T L T 3 m R N A + 33 140 m T
where E T is the Young’s modulus of the tower material, I T is the average area moment of inertia of the tower, m T is the mass of the tower, m R N A is the mass of the rotor-nacelle assembly, and L T is the length of the tower.
  • Step 5. Compute the flexible-base natural frequency.
The sway-bending mode natural frequency can be expressed as:
f 0 = C L C R C S f F B
where C L and C R are the lateral and rotational foundation flexibility coefficients, C S is the substructure flexibility coefficient, and f F B is the fixed-base (cantilever) natural frequency of the tower.

3.2. Wind Turbine Supported by Single-Suction Caisson

Large diameter caissons are regarded as alternatives to monopiles for water depths below 30 m. A suction caisson essentially consists of a large-diameter rigid circular lid with thin skirts as shown in Figure 6. The procedure proposed by Arany et al. [9,10], presented in Section 3.1., can also be used for the single-suction caisson that can be represented by the three-spring model. The input parameters required for computing the natural frequency of a suction-caisson-supported OWT are listed in Table 5.
A list of parameters required for the estimation of natural frequency is presented in Table 5.
The steps involved in the estimation of natural frequency are the same as those described in Section 3.1. with the exception of the foundation stiffness calculations. Skirted foundations trap soil below the lid and their stiffness can resemble that of rigid embedded foundations [43]. Jalbi et al. [44] carried out numerical analysis to study the stiffness of suction caissons. The aforementioned research also found that when the aspect ratio, L/D, was greater than 2, available closed-form solutions for deep foundations reported by [45], can be used. The stiffness functions for three different soil profiles are presented in Appendix A.1: Table A1. For the aspect ratio in the range 0.2 < L/D < 2, solutions presented by [44] as presented in Appendix A.1: Table A2, are recommended.

3.3. Wind Turbine Supported by Monopile

The mathematical modeling of the tower–monopile system is presented in Figure 7. The natural frequency of the system can be estimated following the procedure developed in Arany et al. [11,13], discussed in the preceding sections. However, it is important to identify the pile-head stiffness based on the soil profile at the site. The calculation steps remain the same, with the exception that pile stiffness needs to be calculated in Step 2. The data required to estimate the natural frequency of a monopile-supported wind turbine are listed in Table 6.
Over the years, several authors have proposed expressions to compute the dynamic stiffness of pile foundations. The first step in the calculation of the pile-head stiffness is the classification of pile behavior, i.e., whether the monopile will behave as a long flexible pile or a short one. Rigid piles undergo rigid body rotation in soil owing to their short length, whereas slender piles undergo deflection through the formation of plastic hinge. Two commonly adopted criteria to determine whether a pile can be slender or rigid are presented in Table 7. Formulae proposed by different researchers for KL, KR, and KLR for slender piles in various soil profiles are presented in Table A3. Table A4 presents formulae proposed by different researchers for pile-head stiffness for rigid piles in various soil profiles.

3.4. Wind Turbine Supported by Jackets on Suction Buckets

Jackets or seabed frames have been used to support OWTs in water depths ranging from 23 m to 60 m. Jackets are typically three or four legged and are supported on either piles or suction caissons. In such systems with multiple foundations, the influence of soil structure interaction and group effects becomes important [49]. A closed-form solution to estimate the natural frequency of a jacket was developed by Jalbi and Bhattacharya [20]. This approach combined with closed-form correction factors to incorporate group effects, proposed by [21] can provide a quick estimate of the natural frequency. The idealized tower jacket system is shown in Figure 8. The information required to estimate the natural frequency is presented in Table 8.
The steps involved in the calculation of natural frequency are as follows:
  • Step 1. Calculate the equivalent stiffness of the jacket-truss and tower,
The equivalent flexural rigidity of the jacket and tower can be obtained using the expressions (refer to Figure 9):
E I J = E I J t o p · f m I J t o p = A c L t o p 2 2 f m = 1 3 · m m 1 3 m 2 2 m l n m 1
where m = L b o t t o m L t o p
E I T   = E I T t o p · f m I T t o p = π 8 D t o p 3 t T f q = 1 3 · 2 q 2 q 1 3 q 2 2 l n q 1 + 4 q 1
where
D T = D t o p + D b o t t o m 2 t T = m T ρ π h T D T q = D b o t t o m D t o p
  • Step 2. Calculate the equivalent bending stiffness of the tower–jacket system.
The equivalent bending stiffness, E I T J , of the tower–jacket system is given in Equation (13):
E I T J = E T I T 1 1 + 1 + Ψ 3 Χ Χ h J + h T h T 3 Χ = E T I T E J I J Ψ = h J h T
  • Step 3. Calculate the equivalent mass of the tower–jacket system.
The equivalent mass can be computed by the following equation:
m e q = m z φ 1 2 d z φ 1 2 d z = i = 1 n m i z i 1 z i φ 1 2 d z 0 z i φ 1 2 d z = m J 0 h J φ 1 2 d z + m T h J h J + h T φ 1 2 d z 0 h J + h T φ 1 2 d z
The numerical integral can be easily computed as [20]:
φ 1 2 d z = z + 1 β 1 2 4 λ 1 L × sin 2 λ 1 L z β 1 2 λ 1 L × cos 2 λ 1 L z + 1 + β 1 2 4 λ 1 L × sin h 2 λ 1 L z + β 1 2 2 λ 1 L × cos h 2 λ 1 L z 2 β 1 λ 1 L × sin λ 1 L z × sin h λ 1 L z 1 + β 1 2 λ 1 L × sin λ 1 L z × cos h λ 1 L z 1 β 1 2 λ 1 L × cos λ 1 L z × sin h λ 1 L z
The values of λ1 and β1 for a fixed-end cantilever beam can be taken as 1.8751 and −0.7341 respectively [50].
  • Step 4. Calculate the fixed-base natural frequency:
    f F B = 1 2 π 3 E I T J 0.243 m e q h t o t a l + m R N A h t o t a l 3
    where h t o t a l = h t + h J
  • Step 5. Calculate the rotational stiffness of the foundation group.
Assuming homogenous soil, and solid embedded cylindrical foundations as illustrated in Figure 10, the rotational stiffness of the foundation group, K R g can be computed using a group interaction factor as given in Equation (17):
K R g = K R g s · 1 1 + f 1 R s D + f 2 R s D 2
f 1 R = 0.67 1 0.13 N 1 0.53 ν 1 + 0.35 L D 0.49
f 2 R = 0.29 1 0.04 N 1 0.12 ν 1 + 2.87 L D
where s represents spacing between foundations and D represents the diameter of the cylindrical foundation element as shown in Figure 10. The rocking stiffness parameter for the foundation group, K R g s is given by:
K R g s = N K R f + 1 2 r 2 K V f
K R f = G D 3 3 1 ν 1 + 7.5 9 ν L D + 10.5 7.7 ν L D 2.5
K V f = 2 G D l n 3 4 ν 1 2 ν 1 + 1.12 1 0.84 ν L D 0.84
  • Step 6. Calculate the flexibility factor C J .
The flexibility factor C J can be computed using the expression:
C J = τ τ + 3 τ = K R g h t o t a l E I T J
  • Step 7. Calculate the flexible-base natural frequency.
The flexible-base natural frequency can be expressed as the multiple of a flexibility parameter C J and the fixed-base natural frequency f F B as:
f 0 = C J f F B
where C J is a function of the rotational stiffness of the foundation and f F B is the fixed-base natural frequency.

3.5. Wind Turbine Supported by Jacket on Piles

The case of jacket supported on a deep foundation can be analyzed following a similar procedure as that of a jacket on suction caissons, using the appropriate flexibility factor. The rotational stiffness of the foundation group in Equation (19) can be computed using structural models, considering both the lateral and vertical stiffness of individual piles. An approximate method proposed by Jalbi and Bhattacharya [20] leads to the following equation for rotational stiffness, K R g for square configuration of the foundation:
K R g = K V p s 2
The vertical stiffness of a single pile, K V p , can be estimated using an appropriate numerical analysis or estimation methods assuming idealized soil profiles. An approximate expression proposed by [39] is:
K V p = 2 π L p G s ζ
where ζ lies between 3 and 5.
Another expression proposed for the seismic design of bridges by [51] is as follows:
K V p = 1.25 E p A L p
In these expressions, L p refers to the length of the pile, G s refers to the shear modulus of homogenous soil, and A represents the cross-sectional area of the pile.

3.6. Wind Turbine Supported on Tension-Leg Platform

The tension-leg platform wind turbine (TLPWT) concept design has promising application in intermediate water depths. Floating wind turbine system modeling requires simulation of complex interactions between wind- and wave-driven responses. However, following reasonable engineering assumptions and physical principles, first predictions of TLPWT behavior can be computed using simple spreadsheet calculations. Following simple linear approximations to find the added mass matrix [A], hydrostatic stiffness matrix [C], and mooring system stiffness matrix [K] discussed in previous literature [52,53], Bachynski and Moan [35] compiled the equations for quick estimation of the natural frequencies. Figure 11 presents the geometry parameters for rectangular and cylindrical pontoon designs and Table 9 presents the input parameters required for the calculation.
The step-by-step procedure to arrive at the surge/sway and heave natural frequencies is as follows:
  • Step 1. Calculate the mooring system stiffness factors.
The mooring system stiffness factors are calculated as:
K 11 j = 1 n t k 11 k 11 = T o l o
K 33 j = 1 n t k 33 k 33 = E t A t l o
where T o represents the pretension in each mooring cable, l o represents the unstretched length of cable, nt represents the number of tendons, Et represents the elastic modulus, and At represents the cross-sectional area of the tendons.
  • Step 2. Calculate the added-mass terms.
The added-mass terms can be calculated as:
A 11 a t D 1 h 1 b t + a t D 2 h 2 + j = 1 n p l p a t [ h p   o r   d p ] cos 2 θ i
A 33 ρ π 1 12 D 2 3 + j = 1 n p l p a t [ w p   o r   d p ]
In these equations, h 1 , h 2 , D 1 , D 2 , d p , and w p are geometry parameters as defined in Figure 5, and b t represents the distance from freeboard to tower base ( b t = h 1 + h 2 T ). The transverse added-mass per-unit lengths for a cylinder with diameter D, and a square section with side length h are given by Equations (24) and (25) respectively.
a t D = ρ π D 2 / 4
a t h = 4.75 ρ h 2 2
  • Step 3. Calculate the hydrostatic stiffness terms.
The hydrostatic stiffness term can be computed using the equation:
C 33 = ρ g A w p
  • Step 4. Compute the natural frequencies.
    f n 1 = 1 2 π K 11 M 11 + A 11
    f n 3 = 1 2 π C 33 + K 33 M 33 + A 33
The parameters M 11 and M 33 represent the mass in surge and heave movements.

3.7. Wind Turbine Supported by Floating Spar Buoy

A spar-buoy-type OWT system essentially consists of a relatively deep cylindrical base providing the ballast, with a heavier lower part raising the centre of buoyancy about the centre of gravity of the system. Ye and Ji [36] idealized the spar-type OWT system into a free–free beam with two large mass components attached at its ends, which represent the RNA and the floating platform. A schematic of the idealized model of the spar-buoy OWT is presented in Figure 12. Torsion springs k1 and k2 are imposed at both ends of the tower beam to model the connections at the tower–floating platform and tower–nacelle joints. The angle between the longitudinal axes of the platform and the undeformed tower is represented by θ 1 and the rotating angle at this location is represented by α 1 . The angle between the longitudinal axes of the nacelle and the undeformed tower is represented by θ 2 , and the rotating angle at this location is represented by α 2 . The parameters required to estimate the natural frequency is presented in Table 10.
The steps to estimate the natural frequency are as follows:
  • Step 1. Compute the matrix elements.
The matrix elements can be formed keeping λ as a variable.
R 11 = R 13 = I 01 ω 2 λ
R 12 = E I λ 2   1 I 01 ω 2 k 1 + g 1 M 1 ω 2 cos θ 1
R 14 = E I λ 2   1 I 01 ω 2 k 1 + g 1 M 1 ω 2 cos θ 1
R 21 = E I λ 3 + m r 2 ω 2 λ g 1 M 1 ω 2 λ cos θ 1
R 22 = M 1 ω 2 g 1 M 1 ω 2 cos θ 1 E I λ 2 k 1
R 23 = E I λ 3 + m r 2 ω 2 λ g 1 M 1 ω 2 λ cos θ 1
R 24 = M 1 ω 2 + g 1 M 1 ω 2 cos θ 1 E I λ 2 k 1
R 31 = E I λ 2 S I 01 ω 2 λ C + I 02 ω 2 E I λ 2 k 2 S g 2 M 2 ω 2 cos θ 2 S
R 32 = E I λ 2 C + I 01 ω 2 λ C + I 02 ω 2 E I λ 2 k 2 C g 2 M 2 ω 2 cos θ 2 C
R 33 = E I λ 2 S H I 01 ω 2 λ C H I 02 ω 2 E I λ 2 k 2 S H g 2 M 2 ω 2 cos θ 2 C H
R 41 = E I λ 3 C + m r 2 ω 2 λ C + M 2 ω 2 S + g 2 M 2 ω 2 cos θ 2 λ C E I λ 2 k 2 S
R 42 = E I λ 3 S m r 2 ω 2 λ S + M 2 ω 2 C + g 2 M 2 ω 2 cos θ 2 λ S + E I λ 2 k 2 C
R 43 = E I λ 3 C H + m r 2 ω 2 λ C H + M 2 ω 2 S H + g 2 M 2 ω 2 cos θ 2 λ C H + E I λ 2 k 2 S H R 44 = E I λ 3 S H + m r 2 ω 2 λ S H + M 2 ω 2 C H + g 2 M 2 ω 2 cos θ 2 λ S H + E I λ 2 k 2 C H
where:
S = sin λ L T ,   C = cos λ L T ,   S H = sin h λ L T   a n d   C H = cos h λ L T
In these equations, M1 is the mass of the platform, M2 is the mass of the nacelle, I01 is the moment of inertia of the platform about the bottom end of the tower, I02 is the moment of inertia of the nacelle about the top end of the tower, g1 is the location of centre of gravity of the platform, g2 is the location of centre of gravity of the superstructure, and r is the radius of gyration of the tower section.
  • Step 2. Solve the transcendental equation for λ .
The transcendental equation presented in Equation (39) can now be set up. The equation can be solved numerically for λ defined as λ = ω m t E I where ω represents the angular frequency, m t is the mass per unit length of tower, and EI is the flexural rigidity of the tower section.
d e t R 11 R 12 R 13 R 14 R 21 R 22 R 23 R 24 R 31 R 32 R 33 R 34 R 41 R 42 R 43 R 44 = 0

4. Conclusions

Estimating the flexible-base natural frequency of OWTs is an essential step in the cost estimation and design stage of offshore wind projects. This article forms a single point of reference for formulae to estimate the natural frequencies of OWTs. A collection of step-by-step analytical procedures to determine the natural frequencies of OWTs considering soil–structure interaction effects is presented.
Seven different foundation options for OWTs are discussed, with procedures to estimate the foundation stiffness and natural frequency, making this paper a go-to document for rapid decision making and analyses of such turbines, along with comparison of their relative feasibility in terms of structural performance versus power production [54]. It also allows for a wide range of researchers to access and compare their results in public domain, through a common and easily interpretable set of results.
There are several additional advantages offered by these drastically simple, yet effective estimates of natural frequencies. Development of models and subsequent testing of them in scaled scenarios [55] are better understood this way, along with their variabilities and uncertainties, as evidenced from existing reviews [56, 57] and experiments [58]. On the other hand, such estimates allow for easier handling of probabilistic results, like extreme value estimates [59] and related fragility [60]. Time-dependent degradation, like scour aspects, can be considered more easily, and future re-use of sites through assessment of geotechnical fatigue can also be understood in a more direct fashion. The sensitivity of the natural frequencies with respect to various conditions make it easier for a rapid sensitivity analysis, along with better control over sensor placement strategies and non-contact measurements [61]. It is expected that these initial estimates also create baselines for system identification, control [62], model updating, and eventual digital twinning.

Author Contributions

Conceptualization, S.B. and V.P.; methodology, S.B.; software, R.V.; writing—original draft preparation, R.V.; writing—review and editing, V.P. and S.B.; All authors have read and agreed to the published version of the manuscript.

Funding

This research received funding from Enterprise Ireland SEMPRE project DT-2020-0243-A.

Data Availability Statement

No new data were created or analyzed in this study. Data sharing is not applicable to this article.

Acknowledgments

R.V. and V.P. would like to acknowledge SFI MaREI centre R2302_2 and UCD Energy Institute.

Conflicts of Interest

The authors declare no conflict of interest.

Appendix A

Appendix A.1

Expressions for estimating the stiffness of suction caissons are presented below.
Table A1. Impedance functions for deep foundations exhibiting rigid behavior L/D > 2.
Table A1. Impedance functions for deep foundations exhibiting rigid behavior L/D > 2.
Ground profile K L D E S O f υ s K L R D 2 E S O f υ s K R D 3 E S O f υ s
Homogeneous 3.2 L D 0.62 1.8 L D 1.56 1.65 L D 2.5
Parabolic 2.65 L D 1.07 1.8 L D 2 1.63 L D 3
Linear 2.35 L D 1.53 1.8 L D 2.5 1.59 L D 3.45
f υ s = 1 + 0.6 0.25 υ s .
Table A2. Impedance functions for shallow skirted foundations exhibiting rigid behavior 0.5 < L/D < 2.
Table A2. Impedance functions for shallow skirted foundations exhibiting rigid behavior 0.5 < L/D < 2.
Ground profile K L D E S O f υ s K L R D 2 E S O f υ s K R D 3 E S O f υ s
Homogeneous 2.91 L D 0.56 1.87 L D 1.47 2.7 L D 1.92
Parabolic 2.7 L D 0.96 1.99 L D 1.89 2.54 L D 2.44
Linear 2.53 L D 1.33 2.02 L D 2.29 2.46 L D 2.9
f υ s = 1.1 × 0.096 L D + 0.6 υ s 2 0.7 υ s + 1.06 .

Appendix A.2

Expressions for estimating the stiffness of monopile foundations are represented below:
Table A3. Formulae for stiffness by different researchers for slender piles.
Table A3. Formulae for stiffness by different researchers for slender piles.
Lateral stiffness K L Cross-coupling stiffness K L R Rotational stiffness K R
Randolph (1981), slender piles, both for homogeneous and linear inhomogeneous soils
1.67 E S 0 D P f ( ν S ) ( E e q E S 0 ) 0.14 0.3475 E S 0 D P 2 f ( ν S ) ( E e q E S 0 ) 0.42 0.1975 E S 0 D P 3 f ( ν S ) ( E e q E S 0 ) 0.7
Pender (1993), slender piles, homogeneous soil
1.285 E S 0 D P ( E e q E S 0 ) 0.188 0.3075 E S 0 D P 2 ( E e q E S 0 ) 0.47 0.18125 E S 0 D P 3 ( E e q E S 0 ) 0.738
Pender (1993), slender piles, linear inhomogeneous soil
0.85 E S 0 D P ( E e q E S 0 ) 0.29 0.24 E S 0 D P 2 ( E e q E S 0 ) 0.53 0.15 E S 0 D P 3 ( E e q E S 0 ) 0.77
Pender (1993), slender piles, parabolic inhomogeneous soil
0.735 E S 0 D P ( E e q E S 0 ) 0.33 0.27 E S 0 D P 2 ( E e q E S 0 ) 0.55 0.1725 E S 0 D P 3 ( E e q E S 0 ) 0.776
Poulos and Davis (1980) following Barber (1953), slender pile, homogeneous soil
k h D P β k h D P β 2 k h D P 2 β 3
Poulos and Davis (1980) following Barber (1953), slender pile, linear inhomogeneous soil
1.074 n h 3 5 ( E P I P ) 2 5 0.99 n h 2 5 ( E P I P ) 3 5 1.48 n h 1 5 ( E P I P ) 4 5
Gazetas (1984) and Eurocode 8 Part 5 (2003), slender pile, homogeneous soil
1.08 D P E S 0 ( E e q E S 0 ) 0.21 0.22 D P 2 E S 0 ( E e q E S 0 ) 0.50 0.16 D P 3 E S 0 ( E e q E S 0 ) 0.75
Gazetas (1984) and Eurocode 8 Part 5 (2003), slender pile, linear inhomogeneous soil
0.60 D P E S 0 ( E e q E S 0 ) 0.35 0.17 D P 2 E S 0 ( E e q E S 0 ) 0.60 0.14 D P 3 E S 0 ( E e q E S 0 ) 0.80
Gazetas (1984) and Eurocode 8 Part 5 (2003), slender pile, parabolic inhomogeneous soil
0.79 D P E S 0 ( E e q E S 0 ) 0.28 0.24 D P 2 E S 0 ( E e q E S 0 ) 0.53 0.15 D P 3 E S 0 ( E e q E S 0 ) 0.77
Shadlou and Bhattacharya (2016), slender pile, homogeneous soil
1.45 E S 0 D P f ( ν s ) ( E e q E s 0 ) 0.186 0.30 E S 0 D P 2 f ( ν s ) ( E e q E s 0 ) 0.50 0.18 E S 0 D P 3 f ( ν s ) ( E e q E S 0 ) 0.73
Shadlou and Bhattacharya (2016), slender pile, linear inhomogeneous soil
0.79 E S 0 D P f ( ν s ) ( E e q E s 0 ) 0.34 0.26 E S 0 D P 2 f ( ν s ) ( E e q E S 0 ) 0.567 0.17 E S 0 D P 3 f ( ν s ) ( E e q E S 0 ) 0.78
Shadlou and Bhattacharya (2016), slender pile, parabolic inhomogeneous soil
1.02 E S 0 D P f ( ν s ) ( E e q E S 0 ) 0.27 0.29 E S 0 D P 2 f ( ν s ) ( E e q E S 0 ) 0.52 0.17 E S 0 D P 3 f ( ν s ) ( E e q E S 0 ) 0.76
Parameter definitions:
E e q = E P I P D P 4 π 64
f ( ν S ) = 1 + ν S 1 + 0.75 ν S for Randolph (1981) and
f ( ν s ) = 1 + | ν s 0.25 | for Shadlou and Bhattacharya (2016)
β = k h D P E P I P 4
Table A4. Formulae for stiffness by different researchers for rigid piles.
Table A4. Formulae for stiffness by different researchers for rigid piles.
K L K L R K R
Poulos and Davis (1980) following Barber (1953), rigid pile, homogeneous soil ( n = 0 )
k h D P L P k h D P L P 2 2 k h D P L P 3 3
Poulos and Davis (1980) following Barber (1953), rigid pile, linear inhomogeneous soil ( n = 1 )
1 2 L P 2 n h 1 3 L P 3 n h 1 4 L P 4 n h
Carter and Kulhawy (1992), rigid pile, rock
3.15 G * D P 2 3 L P 1 3 1 0.28 ( 2 L P D P ) 1 4 2 G * D P 7 8 L P 9 8 1 0.28 ( 2 L P D P ) 1 4 4 G * D P 4 3 L P 5 3 1 0.28 ( 2 L P D P ) 1 4
Shadlou and Bhattacharya (2016), rigid pile, homogeneous soil ( n = 0 )
3.2 E S 0 D P f ( ν s ) ( L P D P ) 0.62 1.7 E S 0 D P 2 f ( ν s ) ( L P D P ) 1.56 1.65 E S 0 D P 3 f ( ν s ) ( L P D P ) 2.5
Shadlou and Bhattacharya (2016), rigid pile, linear inhomogeneous soil ( n = 1 )
2.35 E S 0 D P f ( ν s ) ( L P D P ) 1.53 1.775 E S 0 D P 2 f ( ν s ) ( L P D P ) 2.5 1.58 E S 0 D P 3 f ( ν s ) ( L P D P ) 3.45
Shadlou and Bhattacharya (2016), rigid pile, parabolic inhomogeneous soil ( n = 1 / 2 )
2.66 E S 0 D P f ( ν s ) ( L P D P ) 1.07 1.8 E S 0 D P 2 f ( ν s ) ( L P D P ) 2.0 1.63 E S 0 D P 3 f ( ν s ) ( L P D P ) 3.0
Parameter definitions:
E e q = E P I P D P 4 π 64 f ( ν s ) = 1 + | ν s 0.25 |
β = k h D P E P I P 4

Appendix B

Appendix B.1

The natural frequency of a 2.5 MW wind turbine in the Qidong sea area, described in [63], is evaluated. The instrumented turbine rested on a pre-stressed concrete bucket foundation with a diameter of 30 m. The details of the structure are presented in Table A5 and the soil profile is presented in Table A6.
Table A5. Input data for the wind turbine on GBS.
Table A5. Input data for the wind turbine on GBS.
#Input ParameterValueUnit
1Mass of the rotor-nacelle assembly 130 tons
2Tower height80 m
3Tower top diameter2.8 m
4Tower bottom diameter4.3 m
5Average tower wall thickness0.035 m
6Tower Young’s modulus210 GPa
7Tower mass247 tons
10Foundation diameter30 m
Table A6. Assumed soil profile at the site.
Table A6. Assumed soil profile at the site.
Soil LayerThickness (m)Unit Weight (kN/m3)Elastic Modulus (MPa)Poisson’s Ratio
1100.9714.20.35
2180.9210.60.35
3220.9511.50.35
The calculations are as follows:
D T = D b + D t 2 = 2.8 + 4.3 2 = 3.55 t T = 247000 × 7850 × 80 × 3.55 × π = 0.035
I T = 1 8 × 3.55 3 × 0.035 × π = 0.615
E T I T = 129.131 × 10 9   Nm 2
The lateral and rotational stiffness of the foundation is calculated using the method proposed by Suryasentana and Mayne using Equations (7) to (10).
The Geq computed for lateral and rotational modes using Equation (8) are 5.07 and 5.03 MPa respectively.
K l = 16 1 ν D 7 8 ν G e q = 376.84   MPa
K r = D 3 3 1 ν G e q = 69593.17   MPa
q = D b D t = 1.535
f q = 1 3 · 2 q 2 q 1 3 2 q 2 l n q 3 q 2 + 4 q 1 = 2.653
E I η = E I t · f q = 342.628 × 10 9 Nm 2
The nondimensional stiffness factors are:
η l = K l L 3 E I = 563.363
η r = K r L E I = 16.247
The foundation flexibility coefficients are:
C R η L , η R , η L R = 1 1 1 + a η R η L R 2 η L = 0.890
C L η L , η R , η L R = 1 1 1 + b η L η L R 2 η R = 0.997
χ = E T I T E s I s = 0 C S = 1 1 + 1 + ψ 3 χ χ = 1
f F B = 1 2 π 3 E T I T L T 3 m R N A + 33 140 m T = 0.319   Hz
f 0 = C L C R C S f F B = 0.283   Hz
The calculation yields a first natural frequency estimate of 0.283 Hz. The influence of boundary conditions on the natural frequency of this structure can be understood with a simple parametric study using this example problem. In the first analysis, the ratio P L 2 / E T I T was varied, where P represents the weight of the RNA. In the second analysis, the lateral and rotational stiffness of the foundation was determined for a range of foundation diameters, represented by the ratio D/LT. The variation of natural frequency with varying dimensionless mass and foundation stiffnesses is presented in Figure A1.
Figure A1. Variation of natural frequency of the gravity base wind turbine with varying (a) mass of RNA and (b) foundation diameter.
Figure A1. Variation of natural frequency of the gravity base wind turbine with varying (a) mass of RNA and (b) foundation diameter.
Energies 15 02967 g0a1

Appendix B.2

A solved example of an OWT on monopile is discussed here. Data from the Horns Rev Wind Farm [64], presented in Table A7, is used in this example.
Table A7. Input data from Horns Rev 1 OWT.
Table A7. Input data from Horns Rev 1 OWT.
#Input ParameterValueUnit
1Mass of the rotor-nacelle assembly 100 tons
2Tower height70 m
3Tower top diameter2.3 m
4Tower bottom diameter4 m
5Average tower wall thickness0.035 m
6Tower Young’s modulus210 GPa
7Tower mass130 tons
8Length of the substructure0 m
9Monopile length21.9 m
10Monopile diameter4 m
11Monopile wall thickness0.05 m
12Monopile Young’s modulus210 N m 2
The calculation of natural frequency is as follows:
D T = D b + D t 2 4 + 2.3 2 = 3.15 I T = 1 8 3.15 3 × 0.035 × π = 0.4 15   m 4
f F B = 1 2 π 3 × 210 E 9 × 0.415 ( 100000 + 33 × 130000 140 ) 70 3 = 0.385   Hz
The fixed-base frequency is therefore 0.385 Hz.
q = 4 2.3 = 1.74 f ( q ) = 1 3 × 2 × 1.74 2 ( 1.74 1 ) 3 2 × 1.74 2 l n 1.74 3 × 1.74 2 + 4 × 1.74 1 = 3.56
E I η = 210 × 1 8 2.3 3 × 0.035 × π × 3.56 = 119.4   GNm 2
The nondimensional groups are:
η L = 0.8941 × 70 3 119.4 = 2568 η L R = 4.45 × 70 2 119.4 = 182.6 η R = 46.25 × 70 119.4 = 27.1
The foundation flexibility coefficients are given as follows:
C R η L , η L R , η R = 1 1 1 + 0.6 27.1 182.6 2 2568 = 0.894 C L η L , η L R , η R = 1 1 1 + 0.5 2568 182.6 2 27.1 = 0.999
The natural frequency is therefore given by:
f 0 = 0.894 × 0.999 × 0.385 = 0.344 Hz

Appendix B.3

A solved example of an OWT on mono-caisson is discussed here. A 5 MW reference wind turbine on a mono-caisson is considered in this example. The input data is presented in Table A8.
Table A8. Input data [65].
Table A8. Input data [65].
#Input ParameterValueUnit
1Mass of the rotor-nacelle assembly350 tons
2Mass of tower347.5 tons
3Height of tower87.6 m
4Platform height30 m
5Tower bottom diameter3.87 m
6Wall thickness of tower27 mm
7Diameter of caisson12 m
8Height of caisson6 m
9Elastic modulus of soil40 MPa
10Poisson’s ratio of soil0.35
The natural frequency is calculated as follows:
Numerical calculation:
D T = 3.87 + 6 2 = 4.935   m
I T = 1 2 4.935 3 × 0.027 × π = 1.25   m 4
χ = 210 × π × 4.935 3 × 0.027 × 1 8 210 × π × 6 3 × 0.027 × 1 8 = 0.55
ψ = 30 87.6 = 0.34
C M P = 1 1 + 1 + 0.34 3 × 0.55 0.55 = 0.75
f F B = 1 2 π × 0.75 × 3 × 210 × 10 9 × 1.25 350000 + 33 × 347460 140 × 87.6 3 = 0.26   Hz
The fixed-base frequency is therefore 0.26 Hz.
q = 6 3.87 = 1.55
f q = 1 3 × 2 × 1.55 2 1.55 1 3 2 × 1.55 2 l n 1.55 3 × 1.55 2 + 4 × 1.55 1 = 2.71
E I η = 210 × 1 8 3.87 0.027 3 × 0.027 × π × 2.71 = 342   GNm 2
The nondimensional groups are:
η L = 0.86 × 87.6 3 342 = 1690
η L R = 3.5 × 87.6 2 342 = 1690
η R = 44 × 87.6 342 = 11
The foundation flexibility coefficients are given as follows:
C R η L , η L R , η R = 1 1 1 + 0.6 11 78 2 1690 = 0.82
C L η L , η L R , η R = 1 1 1 + 0.5 1690 78 2 11 = 0.999
In this case CS is 1. The natural frequency is given by:
f 0 = C L C R f F B
f 0 = 0.82 × 0.999 × 0.26 = 0.213   Hz

Appendix B.4

The case of an OWT supported by a tower–jacket–foundation system described in [20] is adopted in this example.
Table A9. Input data for the tower–jacket–foundation system.
Table A9. Input data for the tower–jacket–foundation system.
#Input ParameterSymbolUnit
Tower
1Mass of RNA350 tons
2Height of tower70 m
3Top diameter4 m
4Bottom diameter5.6 m
5Distributed mass of tower3730 m
6Density of material7850 kg m 3
7Young’s modulus of material210 GPa
Jacket
9Top width9.5 m
10Bottom width12 m
11Height of jacket70 m
12Area of jacket leg chords 0.1281 m 2
13Distributed mass of jacket8150 m
Foundation
15Number of footings4
16Distance, s12 m
17Distance, r8.48 m
18Diameter of foundation4 kg
19Depth of foundation
dpe
4 kg
The calculations for natural frequency are as follows:
The equivalent flexural rigidity of the jacket calculated first.
m = L b o t t o m L t o p = 1.263
f m = 1 3 · m m 1 3 m 2 2 m l n m 1 = 1.425
I J t o p = A c L t o p 2 2 = 11.561 m 4
E I J = E I J t o p · f m = 3.46082 × 10 12   Nm 2
Similarly for the tower,
q = D b o t t o m D t o p = 1.4
D T = D t o p + D b o t t o m 2 = 4.8   m
t T = m T ρ π h T D T = 0.032   m
f q = 1 3 · 2 q 2 q 1 3 q 2 2 l n q 1 + 4 q 1 = 2.146
I T t o p = π 8 D t o p 3 t T = 0.792
E I T = E I T t o p · f m = 3.56869 × 10 11   Nm 2
The equivalent bending stiffness, E I T J of the tower–jacket system is calculated next.
Ψ = h J h T = 1
Χ = E T I T E J I J = 0.103
E I T J = E T I T 1 1 + 1 + Ψ 3 Χ Χ h J + h T h T 3 = 1.6581 × 10 12   Nm 2
The equivalent mass is calculated as:
L = 140 λ 1 L = 0.0134 β 1 = 0.7341 φ 1 2 d z @ z = 0   = 47.523 φ 1 2 d z @ z = 0   = 47.523 φ 1 2 d z @ z = 70   = 165.709 φ 1 2 d z @ z = 140   = 1151.122 0 70 φ 1 2 d z   = 118.187 70 140 φ 1 2 d z   = 985.412 0 140 φ 1 2 d z   = 1103.599 m e q = 4203.347
The fixed-base natural frequency is:
f F B = 1 2 π 3 E I T J 0.243 m e q h t o t a l + m R N A h t o t a l 3 = 0.305
The foundation stiffnesses are computed as:
K V f = 2 G D l n 3 4 ν 1 2 ν 1 + 1.12 1 0.84 ν L D 0.84 = 8.31 × 10 7   N m
K R f = G D 3 3 1 ν 1 + 7.5 9 ν L D + 10.5 7.7 ν L D 2.5 = 1.66 × 10 9   N m rad
K R g s = N K R f + 1 2 r 2 K V f = 1.86 × 10 10   N m rad
K R g = K R g s · 1 1 + f 1 R s D + f 2 R s D 2 = 1.90 × 10 10   N m rad
The flexibility factor:
τ = K R g h t o t a l E I T J = 1.605
C J = τ τ + 3 = 0.590
The flexible-base natural frequency is then calculated as:
f 0 = C J f F B = 0.179   Hz

Appendix B.5

An example of an OWT supported by a tension-leg platform is presented here. Data from the NREL-MIT design [28,65] with rectangular pontoons is considered.
Table A10. Input data.
Table A10. Input data.
#Input ParameterSymbolUnit
1Overall draft40 m
2Number of tendons 4
3Pretension in each cable6868 m
4Unstretched length of tendon140 m
5Young’s modulus 200 m
6Outer diameter1.4 m
7Thickness 46.2 m
8Number of pontoons4
9Height of rectangular pontoon2.4 m
10Width of rectangular pontoon2.4 m
11Radius of pontoon from cylinder center27 m
12Vertical location of pontoon−43.8 m
13Diameter of main cylinder of hull18 m
14Diameter of base node of hull18 m
15Height of main cylinder of hull52.6 m
16Height of base node of hull2.4 m
17Center of gravity (full system)−32.7957 m
18Center of buoyancy−23.945 m
19Total steel mass2322 tons
Individual surge stiffness:
k 11 = T o l o = 49.06   kN / m
Total surge stiffness K 11 = 196.23   kN / m
Individual heave stiffness:
k 33 = E t A t l o = 290283.2   kN / m
Total heave stiffness K 33 = 1161132.65   kN / m
Freeboard to the tower base (bt = h1 + h2T) = 10 m
a t D = ρ π D 2 / 4
a t h = 4.75 ρ h 2 2
Added mass is computed as:
A 11 = 1000 π 18 2 4 52.6 10 + 1000 π 18 2 4 2.4 + 2 × 27 × 4.754 × 1000 × 2.4 2 2 = 11820776.26   kg
A 33 = 1000 π 1 12 18 3 + 4 × 27 × 4.754 × 1000 × 2.4 2 2 = 2266156.11   kg
Hydrostatic stiffness term for heave:
C 33 = 1000 × 9.81 × π × 18 2 4 = 2496340.94   kg / s 2
Natural frequencies are:
f n 1 = 1 2 π K 11 M 11 + A 11 = 0.018   Hz
f n 3 = 1 2 π C 33 + K 33 M 33 + A 33 = 4.493   Hz

Appendix B.6

The example of a spar-buoy-supported OWT is presented here. The case study of a NREL designed 5 MW direct-drive wind turbine is considered in this example [32,36]. The input data are summarized in Table A11.
Table A11. Input data.
Table A11. Input data.
#Input ParameterSymbolUnit
Tower
1Mass of nacelle + blades293.22 tons
2Height of tower90 m
3Average diameter5 m
4Thickness0.03 m
5Location of CG of superstructure64 m
6Angle between longitudinal axes of nacelle and tower90 degrees
7Stiffness of torsion spring between nacelle and tower 1.00 × 10 15 Nm rad
8Moment of inertia of the nacelle about the top end of the tower698720 m 4
Platform
9Mass of platform7593 tons
10Location of CG of platform−92.6 m
11Height of jacket70 m
12Angle between longitudinal axes of platform and tower180 degrees
13Stiffness of torsion spring between platform and tower 1.00 × 10 15 Nm rad rad
14Moment of inertia of the platform about the bottom end of the tower65108152680 m 4
The calculated geometry properties of the tower are:
Area moment of inertia of tower section, I = π 64 d 4 d t 4 = 1.446 m4
Cross-sectional area of tower section, A = 0.468 m2
Radius of gyration, r = I A = 1.757   m
Mass per unit length of tower, m t = ρ × A = 3653.609 kg/m
The equation:
d e t R 11 R 12 R 13 R 14 R 21 R 22 R 23 R 24 R 31 R 32 R 33 R 34 R 41 R 42 R 43 R 44 = 0
is solved numerically using an MS Excel macro.
The elements R 11 to R 44 are calculated for each value of ω in the iteration.
The natural frequency computed from the analysis is 2.908 Hz.

References

  1. Bhattacharya, S. Design of Foundations for Offshore Wind Turbines; John Wiley & Sons: Hoboken, NJ, USA, 2019. [Google Scholar]
  2. Bashetty, S.; Ozcelik, S. Review on Dynamics of Offshore Floating Wind Turbine Platforms. Energies 2021, 14, 6026. [Google Scholar] [CrossRef]
  3. Zania, V. Natural vibration frequency and damping of slender structures founded on monopiles. Soil Dyn. Earthq. Eng. 2014, 59, 8–20. [Google Scholar] [CrossRef]
  4. Andersen, L.; Vahdatirad, M.; Sichani, M.; Sørensen, J. Natural frequencies of wind turbines on monopile foundations in clayey soils—A probabilistic approach. Comput. Geotech. 2012, 43, 1–11. [Google Scholar] [CrossRef]
  5. Harte, M.; Basu, B.; Nielsen, S.R. Dynamic analysis of wind turbines including soil-structure interaction. Eng. Struct. 2012, 45, 509–518. [Google Scholar] [CrossRef]
  6. Lombardi, D.; Bhattacharya, S.; Wood, D.M. Dynamic soil–structure interaction of monopile supported wind turbines in cohesive soil. Soil Dyn. Earthq. Eng. 2013, 49, 165–180. [Google Scholar] [CrossRef]
  7. Bisoi, S.; Haldar, S. Dynamic analysis of offshore wind turbine in clay considering soil–monopile–tower interaction. Soil Dyn. Earthq. Eng. 2014, 63, 19–35. [Google Scholar] [CrossRef]
  8. Bhattacharya, S.; Adhikari, S. Experimental validation of soil–structure interaction of offshore wind turbines. Soil Dyn. Earthq. Eng. 2011, 31, 805–816. [Google Scholar] [CrossRef]
  9. Adhikari, S.; Bhattacharya, S. Vibrations of wind-turbines considering soil-structure interaction. Wind Struct. 2011, 14, 85. [Google Scholar] [CrossRef] [Green Version]
  10. Dezvareh, R. Evaluation of turbulence on the dynamics of monopile offshore wind turbine under the wave and wind excitations. J. Appl. Comput. Mech. 2019, 5, 704–716. [Google Scholar]
  11. Arany, L.; Bhattacharya, S.; Adhikari, S.; Hogan, S.; Macdonald, J. An analytical model to predict the natural frequency of offshore wind turbines on three-spring flexible foundations using two different beam models. Soil Dyn. Earthq. Eng. 2015, 74, 40–45. [Google Scholar] [CrossRef] [Green Version]
  12. Arany, L.; Bhattacharya, S.; Macdonald, J.; Hogan, S.J. Design of monopiles for offshore wind turbines in 10 steps. Soil Dyn. Earthq. Eng. 2017, 92, 126–152. [Google Scholar] [CrossRef] [Green Version]
  13. Arany, L.; Bhattacharya, S.; Macdonald, J.; Hogan, S. Closed form solution of Eigen frequency of monopile supported offshore wind turbines in deeper waters incorporating stiffness of substructure and SSI. Soil Dyn. Earthq. Eng. 2016, 83, 18–32. [Google Scholar] [CrossRef] [Green Version]
  14. Gazetas, G. Formulas and charts for impedances of surface and embedded foundations. J. Geotech. Eng. 1991, 117, 1363–1381. [Google Scholar] [CrossRef]
  15. Gazetas, G. Analysis of machine foundation vibrations: State of the art. Int. J. Soil Dyn. Earthq. Eng. 1983, 2, 2–42. [Google Scholar] [CrossRef]
  16. Veletsos, A.S.; Verbič, B. Vibration of viscoelastic foundations. Earthq. Eng. Struct. Dyn. 1973, 2, 87–102. [Google Scholar] [CrossRef]
  17. Veletsos, A.S.; Wei, Y.T. Lateral and rocking vibration of footings. J. Soil Mech. Found. Div. 1971, 97, 1227–1248. [Google Scholar] [CrossRef]
  18. Mayne, P.W.; Poulos, H.G. Approximate displacement influence factors for elastic shallow foundations. J. Geotech. Geoenviron. Eng. 1999, 125, 453–460. [Google Scholar] [CrossRef]
  19. Suryasentana, S.K.; Mayne, P.W. Simplified method for the lateral, rotational, and torsional static stiffness of circular footings on a nonhomogeneous elastic half-space based on a work-equivalent framework. J. Geotech. Geoenviron. Eng. 2022, 148, 04021182. [Google Scholar] [CrossRef]
  20. Jalbi, S.; Bhattacharya, S. Closed form solution for the first natural frequency of offshore wind turbine jackets supported on multiple foundations incorporating soil-structure interaction. Soil Dyn. Earthq. Eng. 2018, 113, 593–613. [Google Scholar] [CrossRef] [Green Version]
  21. Bordón, J.; Aznárez, J.J.; Padrón, L.; Maeso, O.; Bhattacharya, S. Closed-form stiffnesses of multi-bucket foundations for OWT including group effect correction factors. Mar. Struct. 2019, 65, 326–342. [Google Scholar] [CrossRef]
  22. Bordón, J.D.R.; Aznárez, J.J.; Maeso, O.; Bhattacharya, S. Simple approach for including foundation–soil–foundation interaction in the static stiffnesses of multi-element shallow foundations. Géotechnique 2021, 71, 686–699. [Google Scholar] [CrossRef]
  23. Ryu, M.-S.; Jung, M.-U.; Lee, J.-S.; Kim, D.-S. Closed Form Solutions for Predicting Lateral Response of Tripod Suction Pile for Offshore Wind Turbine Foundation. Energies 2020, 13, 6176. [Google Scholar] [CrossRef]
  24. Karimirad, M.; Moan, T. A simplified method for coupled analysis of floating offshore wind turbines. Mar. Struct. 2012, 27, 45–63. [Google Scholar] [CrossRef]
  25. Pegalajar-Jurado, A.; Bredmose, H.; Borg, M. Multi-level hydrodynamic modelling of a scaled 10 MW TLP wind turbine. Energy Procedia 2016, 94, 124–132. [Google Scholar] [CrossRef] [Green Version]
  26. Nematbakhsh, A.; Bachynski, E.E.; Gao, Z.; Moan, T. Comparison of wave load effects on a TLP wind turbine by using computational fluid dynamics and potential flow theory approaches. Appl. Ocean Res. 2015, 53, 142–154. [Google Scholar] [CrossRef] [Green Version]
  27. Oguz, E.; Clelland, D.; Day, A.; Incecik, A.; López, J.A.; Sánchez, G.; Almeria, G.G. Experimental and numerical analysis of a TLP floating offshore wind turbine. Ocean Eng. 2018, 147, 591–605. [Google Scholar] [CrossRef] [Green Version]
  28. Matha, D. Model Development and Loads Analysis of an Offshore Wind Turbine on a Tension Leg Platform with a Comparison to Other Floating Turbine Concepts: April 2009; National Renewable Energy Lab. (NREL): Golden, CO, USA, 2010.
  29. Jonkman, J.M.; Matha, D. Dynamics of offshore floating wind turbines—analysis of three concepts. Wind Energy 2011, 14, 557–569. [Google Scholar] [CrossRef]
  30. Lackner, M.A.; Rotea, M.A. Structural control of floating wind turbines. Mechatronics 2011, 21, 704–719. [Google Scholar] [CrossRef]
  31. Robertson, A.N.; Jonkman, J.M. Loads analysis of several offshore floating wind turbine concepts. In Proceedings of the 21st International Offshore and Polar Engineering Conference, Maui, HI, USA, 19–24 June 2011. [Google Scholar]
  32. Si, Y.; Karimi, H.R.; Gao, H. Modelling and optimization of a passive structural control design for a spar-type floating wind turbine. Eng. Struct. 2014, 69, 168–182. [Google Scholar] [CrossRef]
  33. Cottura, L.; CaraDonna, R.; Ghigo, A.; Novo, R.; Bracco, G.; Mattiazzo, G. Dynamic Modeling of an Offshore Floating Wind Turbine for Application in the Mediterranean Sea. Energies 2021, 14, 248. [Google Scholar] [CrossRef]
  34. Liu, Y.; Ishihara, T. Numerical Study on Sectional Loads and Structural Optimization of an Elastic Semi-Submersible Floating Platform. Energies 2021, 14, 182. [Google Scholar] [CrossRef]
  35. Bachynski, E.E.; Moan, T. Design considerations for tension leg platform wind turbines. Mar. Struct. 2012, 29, 89–114. [Google Scholar] [CrossRef]
  36. Ye, K.; Ji, J. Natural Frequency Analysis of a Spar-Type Offshore Wind Turbine Tower With End Mass Components. J. Offshore Mech. Arct. Eng. 2018, 140, 064501. [Google Scholar] [CrossRef]
  37. Chen, J.; Hu, Z.; Wan, D.; Xiao, Q. Comparisons of the dynamical characteristics of a semi-submersible floating offshore wind turbine based on two different blade concepts. Ocean Eng. 2018, 153, 305–318. [Google Scholar] [CrossRef] [Green Version]
  38. Liu, Y.; Li, S.; Yi, Q.; Chen, D. Developments in semi-submersible floating foundations supporting wind turbines: A comprehensive review. Renew. Sustain. Energy Rev. 2016, 60, 433–449. [Google Scholar] [CrossRef]
  39. Galera-Calero, L.; Blanco, J.M.; Iglesias, G. Numerical Modelling of a Floating Wind Turbine Semi-Submersible Platform. Appl. Sci. 2021, 11, 11270. [Google Scholar] [CrossRef]
  40. Zhao, Z.; Li, X.; Wang, W.; Shi, W.; Zhao, L.; Wang, S. Analysis of dynamic characteristics of an ultra-large semi-submersible floating wind turbine. J. Mar. Sci. Eng. 2019, 7, 169. [Google Scholar] [CrossRef] [Green Version]
  41. Katsanos, E.I.; Thöns, S.; Georgakis, C.Τ. Wind turbines and seismic hazard: A state-of-the-art review. Wind Energy 2016, 19, 2113–2133. [Google Scholar] [CrossRef] [Green Version]
  42. Veritas, D.N. DNV-OS-J101-Design of Offshore Wind Turbine Structures; DNV: Copenhagen, Denmark, 2014. [Google Scholar]
  43. Gelagoti, F.M.; Lekkakis, P.P.; Kourkoulis, R.S.; Gazetas, G. Estimation of elastic and non-linear stiffness coefficients for suction caisson foundations. In Geotechnical Engineering for Infrastructure and Development; ICE Publishing: London, UK, 2015; pp. 943–948. [Google Scholar]
  44. Jalbi, S.; Shadlou, M.; Bhattacharya, S. Impedance functions for rigid skirted caissons supporting offshore wind turbines. Ocean Eng. 2018, 150, 21–35. [Google Scholar] [CrossRef]
  45. Shadlou, M.; Bhattacharya, S. Dynamic stiffness of monopiles supporting offshore wind turbine generators. Soil Dyn. Earthq. Eng. 2016, 88, 15–32. [Google Scholar] [CrossRef]
  46. Poulos, H.G.; Davis, E.H. Pile Foundation Analysis and Design; Wiley: New York, NY, USA, 1980; Volume 397. [Google Scholar]
  47. Randolph, M.F. The response of flexible piles to lateral loading. Geotechnique 1981, 31, 247–259. [Google Scholar] [CrossRef]
  48. Carter, J.P.; Kulhawy, F.H. Analysis of laterally loaded shafts in rock. J. Geotech. Eng. 1992, 118, 839–855. [Google Scholar] [CrossRef]
  49. Abhinav, K.; Saha, N. Coupled hydrodynamic and geotechnical analysis of jacket offshore wind turbine. Soil Dyn. Earthq. Eng. 2015, 73, 66–79. [Google Scholar] [CrossRef]
  50. Blevins, R.D.; Plunkett, R. Formulas for natural frequency and mode shape. J. Appl. Mech. 1980, 47, 461. [Google Scholar] [CrossRef]
  51. Shama, A.; El Naggar, H. Bridge Foundations. Encycl. Earthq. Eng. 2015, 1, 298–317. [Google Scholar]
  52. Chakrabarti, S.; Hanna, S. Added mass and damping of a TLP column model. In Proceedings of the Offshore Technology Conference, Houston, TX, USA, 7–10 May 1990. [Google Scholar]
  53. Jain, A. Nonlinear coupled response of offshore tension leg platforms to regular wave forces. Ocean Eng. 1997, 24, 577–592. [Google Scholar] [CrossRef]
  54. O’Kelly-Lynch, P.; Long, C.; McAuliffe, F.D.; Murphy, J.; Pakrashi, V. Structural design implications of combining a point absorber with a wind turbine monopile for the east and west coast of Ireland. Renew. Sustain. Energy Rev. 2020, 119, 109583. [Google Scholar] [CrossRef]
  55. Buckley, T.; Watson, P.; Cahill, P.; Jaksic, V.; Pakrashi, V. Mitigating the structural vibrations of wind turbines using tuned liquid column damper considering soil-structure interaction. Renew. Energy 2018, 120, 322–341. [Google Scholar] [CrossRef]
  56. Bhattacharya, S.; Lombardi, D.; Amani, S.; Aleem, M.; Prakhya, G.; Adhikari, S.; Aliyu, A.; Alexander, N.; Wang, Y.; Cui, L.; et al. Physical Modelling of Offshore Wind Turbine Foundations for TRL (Technology Readiness Level) Studies. J. Mar. Sci. Eng. 2021, 9, 589. [Google Scholar] [CrossRef]
  57. Otter, A.; Murphy, J.; Pakrashi, V.; Robertson, A.; Desmond, C. A review of modelling techniques for floating offshore wind turbines. Wind Energy 2021. [Google Scholar] [CrossRef]
  58. Jaksic, V.; O’Shea, R.; Cahill, P.; Murphy, J.; Mandic, D.P.; Pakrashi, V. Dynamic response signatures of a scaled model platform for floating wind turbines in an ocean wave basin. Philos. Trans. A Math Phys. Eng. Sci. 2015, 373, 20140078. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  59. O’Leary, K.; Pakrashi, V.; Kelliher, D. Optimization of composite material tower for offshore wind turbine structures. Renew. Energy 2019, 140, 928–942. [Google Scholar] [CrossRef]
  60. Quilligan, A.; O’Connor, A.; Pakrashi, V. Fragility analysis of steel and concrete wind turbine towers. Eng. Struct. 2012, 36, 270–282. [Google Scholar] [CrossRef]
  61. Jaksic, V.; Wright, C.S.; Murphy, J.; Afeef, C.; Ali, S.F.; Mandic, D.P.; Pakrashi, V. Dynamic response mitigation of floating wind turbine platforms using tuned liquid column dampers. Philos. Trans. A Math Phys. Eng. Sci. 2015, 373, 20140079. [Google Scholar] [CrossRef]
  62. Arrigan, J.; Pakrashi, V.; Basu, B.; Nagarajaiah, S. Control of flapwise vibrations in wind turbine blades using semi-active tuned mass dampers. Struct. Control Health Monit. 2011, 18, 840–851. [Google Scholar] [CrossRef]
  63. Zhang, P.Y.; Guo, Y.H.; Ding, H.Y.; Xiong, K.P.; Yang, Y.F. Study on vibration of offshore wind turbine supporting system under the wind-wave coupling effect. J. Vibroengineering 2015, 17, 2655–2668. [Google Scholar]
  64. Augustesen, A.H.; Brødbæk, K.T.; Møller, M.; Sørensen, S.P.H.; Ibsen, L.B.; Pedersen, T.S.; Andersen, L. Numerical Modelling of Large-Diameter Steel Piles at Horns Rev. In Proceedings of the 12th International Conference on Civil, Structural and Environmental Engineering Computing, Funchal, Madeira, 1–4 September 2009. [Google Scholar]
  65. 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.
Figure 1. Simplified diagram of a monopole-supported OWT showing its components.
Figure 1. Simplified diagram of a monopole-supported OWT showing its components.
Energies 15 02967 g001
Figure 2. Rigid-body degrees of freedom.
Figure 2. Rigid-body degrees of freedom.
Energies 15 02967 g002
Figure 3. Modes of vibration for (a) jacket structures supported on piles and (b) jacket structures supported on shallow foundations [20].
Figure 3. Modes of vibration for (a) jacket structures supported on piles and (b) jacket structures supported on shallow foundations [20].
Energies 15 02967 g003aEnergies 15 02967 g003b
Figure 4. Illustration showing (a) bucket/suction caisson; (b) gravity-based; (c) monopile; (d) jacket/lattice structure; (e) tension-leg platform; (f) spar-buoy floating concept and (g) semisubmersible-type offshore wind turbine systems (Modified from [1]).
Figure 4. Illustration showing (a) bucket/suction caisson; (b) gravity-based; (c) monopile; (d) jacket/lattice structure; (e) tension-leg platform; (f) spar-buoy floating concept and (g) semisubmersible-type offshore wind turbine systems (Modified from [1]).
Energies 15 02967 g004
Figure 5. Mathematical model of the OWT on a gravity-based foundation.
Figure 5. Mathematical model of the OWT on a gravity-based foundation.
Energies 15 02967 g005
Figure 6. Mathematical model of the OWT on single-suction caisson.
Figure 6. Mathematical model of the OWT on single-suction caisson.
Energies 15 02967 g006
Figure 7. Illustrations showing (a) the various parts of a monopile supported OWT, (b) dimensions for design and (c) simplified mathematical model.
Figure 7. Illustrations showing (a) the various parts of a monopile supported OWT, (b) dimensions for design and (c) simplified mathematical model.
Energies 15 02967 g007
Figure 8. Assumed distribution of mass and stiffness of the tower–jacket system.
Figure 8. Assumed distribution of mass and stiffness of the tower–jacket system.
Energies 15 02967 g008
Figure 9. Assumed distribution of mass and stiffness of the tower–jacket system.
Figure 9. Assumed distribution of mass and stiffness of the tower–jacket system.
Energies 15 02967 g009
Figure 10. Commonly adopted foundation group layouts.
Figure 10. Commonly adopted foundation group layouts.
Energies 15 02967 g010
Figure 11. Definition of the design parameters of a TLP.
Figure 11. Definition of the design parameters of a TLP.
Energies 15 02967 g011
Figure 12. Schematic of the idealized spar-type OWT system [36].
Figure 12. Schematic of the idealized spar-type OWT system [36].
Energies 15 02967 g012
Table 1. Parameters defining the tower–RNA system.
Table 1. Parameters defining the tower–RNA system.
#Input ParameterSymbolUnit
1Mass of the rotor-nacelle assembly m R N A kg
2Tower height L T m
3Tower top diameter D t m
4Tower bottom diameter D b m
5Average tower wall thickness t T m
6Tower Young’s modulus E T N/m2
7Tower mass m T kg
Table 2. Characteristics of typical wind turbine systems.
Table 2. Characteristics of typical wind turbine systems.
TurbineNRELLWDTUHaliade-XUnit
Power rating581012MW
Rotor diameter126164178.3218.2m
Hub height90110119135m
Rotor speed range6.9–12.16.3–10.56–9.67.81rpm
Cut-in, rated3, 11.44, 12.54, 11.43.5m/s
Cut-out wind speed25252528m/s
Nacelle mass296.78375551.56600tonne
Blade mass17.743541.7255tonne
Tower mass347.465586052500tonne
Tower height87.6106.3115.6129.1m
Tower top diameter3.8755.55.5m
Tower bottom diameter67.788m
Table 3. Information required for estimating the natural frequency of OWT supported on a gravity-based foundation.
Table 3. Information required for estimating the natural frequency of OWT supported on a gravity-based foundation.
#Input ParameterSymbolUnit
1Foundation diameter D m
2Total tower height L t m
3Substructure Young’s modulus E s N m 2
4Substructure moment of inertia I s m 4
5Lateral stiffness of foundation K L N m
6Rocking stiffness of foundation K R Nm rad
Table 4. Values of parameters a and b.
Table 4. Values of parameters a and b.
Stiffnessab
Lateral K L 1.27 0.237 0.049 ν
Rotational K R 1.35 0.17 + 5 ν 4
Here, ν is the Poisson’s ratio of the soil layer.
Table 5. Information required for estimating the natural frequency of OWT supported by single suction caisson.
Table 5. Information required for estimating the natural frequency of OWT supported by single suction caisson.
#Input ParameterSymbolUnit
1Platform height above mudline L S m
2Substructure Young’s modulus E s N m 2
3Substructure moment of inertia I s m 4
4Caisson diameter D m
5Caisson depth L m
6Initial soil Young’s modulus at 1D depth E S 0 N m 2
7Soil Poisson’s ratio ν s
8Lateral stiffness of foundation K L N m
9Cross-stiffness of foundation K L R N
10Rocking stiffness of foundation K R Nm rad
Table 6. Information required for estimating the natural frequency of an OWT supported on a monopile.
Table 6. Information required for estimating the natural frequency of an OWT supported on a monopile.
#Input ParameterSymbolUnit
1Platform height above mudline L S m
2Substructure Young’s modulus E s N m 2
3Substructure moment of inertia I s m 4
4Pile diameter D p m
5Pile depth L p m
6Pile Young’s modulus E p N m 2
7Pile moment of inertia I p m 4
8Initial soil Young’s modulus at 1D depth E S 0 N / m 2
9Soil Poisson’s ratio ν s
10Lateral stiffness of foundation K L N m
11Cross-stiffness of foundation K L R N
12Rocking stiffness of foundation K R Nm rad
Table 7. Criteria to determine relative pile rigidity.
Table 7. Criteria to determine relative pile rigidity.
ReferenceParameters RequiredCriteria
Poulos and Davis [46] β = k h D p 4 E p I p 4 If β L p > 2.5 pile is slender
If β L p < 1.5 pile is rigid
Randolph [47], Carter, and Kulhawy [48] E e = E p I p D p 4 π 64
G * = G s 1 + 3 4 ν s
If L p D p E e G * 2 7 pile is slender
If L p D p 0.05 E e G * 1 2 pile is rigid
Table 8. Information required for estimating the natural frequency of OWT supported by jacket on suction buckets.
Table 8. Information required for estimating the natural frequency of OWT supported by jacket on suction buckets.
#Input ParameterSymbolUnit
1Tower Young’s modulus E GPa
2Height of jacket h J m
3Top spacing of jacket leg chords L t o p m
4Bottom spacing of jacket leg chords L b o t t o m m
5Area of jacket leg chords A c m 2
6Distributed mass of jacket m J kg m
7Equivalent distributed mass of tower–jacket system m e q kg m
8Jacket bending stiffness E I J Nm 2
9Tower bending stiffness E I T Nm 2
10Tower–jacket system bending stiffness E I J T Nm 2
11Number of foundations N
12Vertical stiffness of individual foundation K R f N m
13Rocking stiffness of individual foundation K v f Nm rad
14Rocking stiffness of foundation group K R g s Nm rad
Table 9. Information required for estimating the natural frequency of an OWT supported on a tension-leg platform.
Table 9. Information required for estimating the natural frequency of an OWT supported on a tension-leg platform.
#Input ParameterSymbolUnit
1Overall draft T m
2Number of tendons n t
3Pretension in each cable T o N
4Unstretched length of tendon I o m
5Young’s modulus E t N / m 2
6Outer diameter of tendon d t m
7Thickness of tendon t t m
8Number of pontoons n p
9Diameter of cylindrical pontoon d p m
11Height of rectangular pontoon h p m
12Width of rectangular pontoon w p m
13Radius of pontoon from cylinder center r p m
14Vertical location of pontoon z s m
15Diameter of main cylinder of hull D 1 m
16Diameter of base node of hull D 2 m
17Height of main cylinder of hull h 1 m
18Height of base node of hull h 2 m
19Total steel mass M kg
Table 10. Information required for estimating the natural frequency of the spar-buoy-type OWT.
Table 10. Information required for estimating the natural frequency of the spar-buoy-type OWT.
#Input ParameterSymbolUnit
1Mass of the platform M 1 tonne
2Mass of the rotor-nacelle assembly M 2 tonne
3Tower height h T m
4Tower top diameter D t m
Tower bottom diameter D b m
5Average tower diameter D T m
6Average tower wall thickness t T m
7Tower Young’s modulus E T GPa
8Mass per unit length of tower m T tonne
9Location of CG of platformG1 m
10Location of CG of superstructureg2 m
11Angle between longitudinal axes of platform and tower θ 1 degrees
12Angle between longitudinal axes of nacelle and tower θ 2 degrees
13Radius of gyration of tower sectionr m
14Moment of inertia of the platform about the bottom end of the tower I 01 m 4
15Moment of inertia of the nacelle about the top end of the tower I 02 m 4
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Varghese, R.; Pakrashi, V.; Bhattacharya, S. A Compendium of Formulae for Natural Frequencies of Offshore Wind Turbine Structures. Energies 2022, 15, 2967. https://0-doi-org.brum.beds.ac.uk/10.3390/en15082967

AMA Style

Varghese R, Pakrashi V, Bhattacharya S. A Compendium of Formulae for Natural Frequencies of Offshore Wind Turbine Structures. Energies. 2022; 15(8):2967. https://0-doi-org.brum.beds.ac.uk/10.3390/en15082967

Chicago/Turabian Style

Varghese, Ramon, Vikram Pakrashi, and Subhamoy Bhattacharya. 2022. "A Compendium of Formulae for Natural Frequencies of Offshore Wind Turbine Structures" Energies 15, no. 8: 2967. https://0-doi-org.brum.beds.ac.uk/10.3390/en15082967

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