Next Article in Journal
Advancing Marine Surveillance: A Hybrid Approach of Physics Infused Neural Network for Enhanced Vessel Tracking Using Automatic Identification System Data
Next Article in Special Issue
An Experimental Study on Vortex-Induced Vibration Suppression of a Long Flexible Catenary Cable by Using Vibration Dampers
Previous Article in Journal
Distributed Improved RILOS Guidance-Based Formation Control of Underactuated ASVs for Cooperative Maritime Search
Previous Article in Special Issue
Experimental Research on the Low-Cycle Fatigue Crack Growth Rate for a Stiffened Plate of EH36 Steel for Use in Ship Structures
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

The Method of the Natural Frequency of the Offshore Wind Turbine System Considering Pile–Soil Interaction

1
School of Electrical Engineering, Guangxi University, Nanning 530004, China
2
School of Civil Engineering and Architecture, Jiangsu University of Science and Technology, Zhenjiang 212003, China
3
Xuzhou City Hongda Civil Engineering Laboratory Co., Ltd., Xuzhou 221000, China
4
School of Civil Engineering, Southeast University, Nanjing 211189, China
*
Author to whom correspondence should be addressed.
J. Mar. Sci. Eng. 2024, 12(11), 1912; https://doi.org/10.3390/jmse12111912
Submission received: 27 September 2024 / Revised: 20 October 2024 / Accepted: 23 October 2024 / Published: 25 October 2024
(This article belongs to the Special Issue Safety and Reliability of Ship and Ocean Engineering Structures)

Abstract

:
Accurately and efficiently evaluating the influence of pile–soil interaction on the overall natural frequency of wind turbines is one of the difficulties in current offshore wind power design. To improve the structural safety and reliability of the offshore wind turbine (OWT) systems, a new closed-form solution method of the overall natural frequency of OWTs considering pile–soil interactions with highly effective calculations is established. In this method, Hamilton’s principle and the equivalent coupled spring model (ECS model) were firstly combined. In Hamilton’s theory, the Timoshenko beam assumption and continuum element theory considering the three-dimensional displacement field of soil were used to simulate the large-diameter monopile–soil interaction under lateral load in multilayer soil. Case studies were used to validate the proposed method’s correctness and efficiency. The results show that when compared with the data of 13 offshore wind projects reported in existing research papers, the difference between the overall natural frequency calculated by the proposed method and that reported in this study is within ±10%. This calculation method achieves the goal of convenient, fast and accurate prediction of the overall natural frequency of offshore wind systems.

1. Introduction

In recent years, offshore wind power development has continued to be a major focus for the expansion of renewable energy, representing a substantial opportunity to advance towards carbon neutrality within the foreseeable future. During the normal operating phase, it is crucial to avoid resonance between the overall natural frequency of the offshore wind turbine (OWT) system and the frequencies associated with environmental loads, such as wind, waves and blade rotation [1,2]. To ensure the structural stability of OWTs, the design process must maintain an eigenfrequency within the safe range of f1P to f3P. On the other hand, monopile foundations are commonly used in offshore wind power projects due to their simplicity in construction, ease of installation, cost-effectiveness and simple design process. In China’s coastal regions, monopiles are employed for over 80% of offshore wind turbines. However, the interaction between monopiles and the underlying soil plays a critical role in the overall modal analysis of the turbine system. Therefore, it is important to study how pile–soil interaction affects the natural frequency of the OWT system.
The simplification methods for the pile–soil interaction model of monopile foundations in offshore wind turbines can be categorized into three main types [3,4,5,6,7,8,9,10,11]: (1) The distributed spring (DS) model, which reduces the complexity of pile–soil interaction to a series of independent springs, aligning with the Winkler foundation model. This includes the m-method and p-y curve method, both prevalent in design codes. Byron and Houlsby [12,13] expanded on this model by proposing the “p-y method with pile end resistance” multi-spring model and corresponding calculation formulas, which have become a hot research topic in this area. (2) The equivalent fixed cantilever (EF) model, which simplifies a single pile to an equivalent cantilever beam with a fixed bottom boundary. It offers a straightforward mechanical interpretation and ease of computation, but its applicability is limited to long, flexible steel pipe piles. Finally, (3) the equivalent coupled spring (ECS) model represents the pile with three types of springs that account for translational (KL), rotational (KR) and coupled stiffness (KLR) at the pile top on the seabed. The ECS model is the best choice among the above three models for its simplicity and clear mechanical interpretation. The model can be used to describe the pile–soil interaction as boundary conditions when calculating the overall natural frequency of offshore wind turbines [14,15,16,17]. For analytical solutions of the ECS model, Randolph [18] and Pender [19] provided formulas for the equivalent head stiffness of flexible piles in linearly elastic homogeneous soil; Gazetas et al. [20] provided solutions for flexible piles in linearly elastic heterogeneous soil; and Poulos et al. [21] and Shadlou et al. [22] derived formulas for flexible and rigid piles in layered soils. If lateral loads and overturning moments are not considered, the ECS model degrades into the uncoupled spring model [23]. Additionally, there are also combined forms of the three models mentioned above. For example, Wang et al. [24] used the “MR-θR” model to simulate ideal rigid piles in sandy soil, with pile–soil interaction equivalent to a bending spring at a certain depth; Zha et al. [25] proposed the “p-y and MR-θR” model, which combines features of the DS model and ECS model; Houlsby et al. [26] introduced the hyperplastic accelerated ratchet model (HARM) based on extensive laboratory test data, which better simulates the hysteretic effect of pile head stiffness; and Arany et al. [14] combined the strengths of the EF and ECS models by treating the wind turbine tower base as a fixed end. By incorporating various correction factors related to pile head displacement and rotation, analytical solutions for natural frequency calculation have been proposed [27,28]. However, there is a notable gap in the literature regarding a comprehensive reflection of the influence of pile–soil interaction on the dynamic response of offshore wind turbines. This gap stems from a lack of studies based on continuum theory that considers the three-dimensional displacement field of the soil.
The variational method based on Hamilton’s principle offers distinctive advantages in the calculation of pile response for large-diameter piles in offshore wind turbines [29,30,31,32]. It can successfully describe the intrinsic mechanisms of pile–soil interaction [33,34]. This is crucial for quickly and accurately adjusting monopile design. By using the variational method, the mechanical performance of the monopile was simplified as a Timoshenko beam. It can efficiently describe the displacement field of the surrounding soil using functions after introducing an appropriate soil constitutive model. If the displacement field of the surrounding soil is determined, the strain distribution in the soil can be obtained using continuous medium theory. Furthermore, by incorporating a suitable soil constitutive model, the stress distribution around the pile during loading can be determined, and thus we can achieve the goal of understanding the pile–soil interaction mechanism. Currently, research progress in incorporating soil constitutive models is relatively slow. Han et al. [29,30,31] tried to characterize the stress–strain relationships in sandy and cohesive soils by employing the modified D-P model and Mises model, respectively. Gupta et al. [32] used a nonlinear elastic stiffness degradation distribution model to describe the deformation of the surrounding soil under normal operating conditions. From principles of energy, the variational method has yielded a fitting expression for the dynamic stiffness of the pile head for wind turbine monopiles in layered soils. Depending on the previous studies, Li et al. [33,34,35,36,37,38,39] extended the soil displacement field into three dimensions, providing an accurate representation of the uplift or subsidence of soil near the seabed and the rotational motion of soil in the vicinity of the pile end. However, the above studies are not extensively applied to natural frequency calculations. To better reflect the influence of pile–soil interaction on the overall frequency of OWT systems, it is important to combine this method with the ECS model.
In this study, based on Hamilton’s principle and the variational method, a mechanical model for pile–soil interaction of offshore wind turbine monopile foundations in multilayered soils was constructed. By using the continuous medium assumption and Timoshenko beam theory, and combining the ECS model for simplified calculations, a rapid calculation method for the natural frequency of wind turbines considering pile–soil interaction was proposed. To verify the correctness and efficiency of the proposed method, the results were then compared with measured data from the literature. This method provides effective guidance for engineering practice, and it also establishes a theoretical foundation for analyzing the intrinsic mechanisms of pile–soil interaction and its impact on the overall dynamic response of wind turbines. Therefore, the novelty of this manuscript is as follows:
(1) To obtain a closed-form solution method of the overall natural frequency of OWTs with highly effective calculation, Hamilton’s principle and the equivalent coupled spring model (ECS model) were combined.
(2) To better describe pile–soil interactions with very large diameters, the pile is modeled as a Timoshenko beam, and soil is modeled as a continuous material considering the three-dimensional displacements of soil (existing studies consider only two-dimensional displacements).

2. Theoretical Analysis

2.1. Fundamental Assumptions in Mechanics

In Figure 1, a schematic diagram of typical frequencies for offshore wind turbines is presented, where the eigenfrequency of OWT systems should be in the range of f1P to f3P. Figure 2 gives several typical models for pile–soil interaction, including the ECS model. Figure 3 shows the mechanical model of the pile–soil interaction for a monopile subjected to lateral loads in an r-θ-z cylindrical coordinate system. In this model, the pile top is subjected to lateral displacement Fa(t) and moment Ma(t). The pile has a cross-sectional radius rp and length Lp. The heights from the ground surface to the bottom of each soil layer are Hi (where i = 1, 2, …,n). The key assumptions for this mechanical model are as follows:
  • The pile body is modeled as a Timoshenko beam, considering the shear deformation of the pile.
  • The soil within each soil layer is treated as a laterally isotropic continuous medium.
  • The pile–soil interface is not separated.

2.2. Governing Equations

2.2.1. Hamilton’s Theory

Depending on the Hamilton theory, the total system energy of the pile and soil is given as follows:
L = T U T + W
where T is the kinetic energy, including T1 and T2; UT is the total potential energy, including UT1 and UT2; and W is the work exerted by external force.
The kinetic energy of the pile (T1) can be given as follows:
T 1 = 1 2 0 L p ρ p w t 2 d z + 1 2 0 L p ρ p I p ϕ t 2 d z
where ρp is the density of the pile (kg/m3); w is the deflection of the pile body; t is time; Ip is the bending moment inertia of a pile unit; and ϕ is the shear deformation of the pile shaft.
The kinetic energy of the soil (T2) can be given as follows:
T 2 = 0 L p 0 2 π 0 r p t ρ s u r t 2 + u θ t 2 + u z t 2 r d r d θ d z + 0 L p 0 2 π r p ρ s u r t 2 + u θ t 2 + u z t 2 r d r d θ d z + L p 0 2 π 0 r p ρ s u r t 2 + u θ t 2 + u z t 2 r d r d θ d z + 1 2 L p 0 2 π r p ρ s u r t 2 + u θ t 2 + u z t 2 r d r d θ d z
where ρs is the soil density; ur is the displacement of soil in the r-direction; uθ is the displacement of soil in the θ-direction; and uz is the displacement of soil in the z-direction.
Depending on assumption (1) in Section 2.1, the total potential energy of pile, UT1, is given as follows:
U T 1 = 0 L p 1 2 E p I p ϕ z 2 d z + 0 L p 1 2 κ G p A p w z ϕ 2 d z
where EpIp is the pile bending stiffness; κ is the cross-section shear coefficient of the pile; and Ap is the cross-sectional area of the pile. If κGp reaches to infinity, then the effect of shear deformations can be neglected; thus, Equation (4) reduces to the Euler–Bernoulli beam theory.
The total potential energy of soil can be given by
U T 2 = 1 2 0 0 2 π r p ε i j T σ i j r d r d θ d z + 1 2 L p 0 2 π 0 r p ε i j T σ i j r d r d θ d z
where σij is the stress tensor of the soil; εij is the strain tensor of the soil; and i,j = r, θ, z.
The work exerted by external forces is expressed as follows:
W = F a t w | z = 0 M a t ϕ | z = 0
where Fa(t) = Fa eiωt; Ma(t) = Ma eiωt.
If the external load is static, then the total energy system Π of the single pile and the soil is
= 1 2 E p I p 0 L p ( d ϕ d z ) 2 d z + 1 2 0 L p κ G p A p ( d w d z ϕ ) 2 d z + 1 2 0 L p 0 2 π r p σ i j ε i j r d r d θ d z + 1 2 L p 0 2 π 0 σ i j ε i j r d r d θ d z F a w z = 0 + M a ϕ z = 0 + W weight
When the constitutive models of the pile and soil are elastic, the work exerted by the self-weight of the pile and soil is zero, i.e., δWweight = 0 if the soil is in an elastic stage.

2.2.2. Three-Dimensional Displacement Expression of the Soil

In the above formula, the displacement of the soil around the laterally loaded pile is assumed to be
u r = w ( z , t ) ϕ r ( r ) cos θ u θ = w ( z , t ) ϕ θ ( r ) sin θ u z = ϕ ( z , t ) r p ϕ z ( r ) cos θ
where ϕr, ϕθ and ϕz represent the displacement distribution functions in the r-, θ- and z-directions of the soil, respectively. All three functions are dimensionless and dependent on the radial distance r. The existence of soil displacement in the z-direction reduces the soil domain stiffness, resulting in relatively larger deformation of the foundation. If uz is not considered, Equation (8) degenerates into the expression proposed by Gupta et al. [32,33,34,35,36,37,38,39,40].
Depending on assumption (3) in Section 2.1 and energy transfer laws, the boundary conditions of ϕr, ϕθ and ϕz can be defined as follows:
ϕ r ( r ) = 1 0 r r p 0 r +
ϕ θ ( r ) = 1 0 r r p 0 r +
ϕ z ( r ) = 0 r p = 0 r r p 0 < r < r p 1 r = r p 0
From Equation (11), within the radial domain inside the pile (0 ≤ r < rp), the vertical displacement of the soil is consistent with the vertical component of the deformation caused by the Timoshenko beam. At r = rp, the interface between the pile and the soil exhibits no interfacial separation, embodying an idealized complete rough contact. At the boundary of infinity, the soil displacement approaches zero.
According to Assumption 2 in Section 2.1, the displacement–strain relationship in the soil domain is as follows:
ε r r ε θ θ ε z z ε r θ ε r z ε θ z = u r r u r r 1 r u θ θ u z z 1 2 1 r u r θ u θ r + u θ r 1 2 u z r u r z 1 2 1 r u z θ u θ z = w ( z , t ) d ϕ r ( r ) d r cos θ w ( z , t ) ϕ r ( r ) r cos θ + 1 r w ( z , t ) ϕ θ ( r ) cos θ ϕ ( z , t ) d z r p ϕ z ( r ) cos θ 1 2 1 r w ( z , t ) ϕ r ( r ) sin θ + w ( z , t ) d ϕ θ ( r ) d r sin θ w ( z , t ) ϕ θ ( r ) r sin θ 1 2 ϕ ( z , t ) r p d ϕ z ( r ) d r cos θ w ( z , t ) z ϕ r ( r ) cos θ 1 2 ϕ ( z , t ) r r p ϕ z ( r ) sin θ + w ( z , t ) z ϕ θ ( r ) sin θ

2.2.3. Constitutive Model of Soil

From the study by Thomas [41], during the standard operation stage of offshore wind turbines, the soil surrounding a monopile foundation is subjected to small-amplitude cyclic loading conditions, and the small-amplitude cyclic loads caused by wind, waves and currents mobilize the soil with minimal to small strains (Figure 4). The soil is also in an elastic state under very small to small strain conditions if the overall natural frequency of wind turbines is mobilized. Therefore, it is reasonable to employ an elastic constitutive model for the soil within the scope of this study.
In this section, two kinds of foundation models were discussed:
(a)
The impact of embedment depth on the elastic modulus of the soil:
E s ( z ) = E s 0 z D p α
where Es represents the elastic modulus of the soil (kPa), Es0 is the initial value of the elastic modulus and α is an exponent, where 0 ≤ α ≤ 1. If α = 0, the soil is homogeneous; if α = 0.5, the distribution of the soil’s elastic modulus with depth follows a parabolic shape; and if α = 1, the soil conforms to the Gibson foundation model.
(b)
Considering the effects of multiple soil layers, the soil in each layer is divided into n segments along the z-direction. The elastic modulus of the soil is assigned within each segment Δz, resulting in a matrix representation:
E s = E s 1 E s 2 E s 3 E s i E s n T 1 × n
where Es represents the global soil matrix of the elastic modulus and Esi denotes the elastic modulus corresponding to the i th soil layer (i = 1,2,3, …,n).
Under dynamic loading, the Lame’s coefficients for elastic soil materials are given by
λ s = λ s 1 + 2 i ς G s = G s 1 + 2 i ς
where λs and Gs represent the compression modulus and shear modulus of soil, respectively; i is the imaginary unit; and ζ is the damping ratio of soil material, which is independent of frequency. Under static loading conditions, λs = Esvs/(1 + vs)/(1 − 2vs) and Gs = Es/2/(1 + vs).
Therefore, the soil’s constitutive relationship can be expressed as follows:
(a)
Dynamic loading condition: σ i j = λ s z δ i j ε i j + 2 G s z ε i j ;
(b)
Static loading condition: σij = λs(z)δijεij + 2Gs(z)εij.
Here, δij is the stress tensor, εij is the strain tensor and z indicates the depth or layer within the soil profile.

2.2.4. The Application of the Variational Method

Substituting Equations (2)–(6) and Equations (8)–(11) into Equation (1) and applying the variational principle, Equation (2) yields the following expression:
M w δ w d z + P ϕ δ ϕ d z + Q ϕ r δ ϕ r d z + R ϕ θ δ ϕ θ d z + S ϕ z δ ϕ z d z = 0
where M(w), P(ϕ), Q(ϕr), R(ϕθ) and S(ϕz) are polynomials that contain δ(w), δ(ϕ), δ(ϕr), δ(ϕθ) and δ(ϕz), respectively. According to the variational method, the equality holds identically if and only if M(w) = 0, P(ϕ) = 0, Q(ϕr) = 0, R(ϕθ) = 0 and S(ϕz) = 0 when δ(ϕ), δ(w), δ(ϕr), δ(ϕθ) and δ(ϕz) are infinitesimal quantities. From this, the relational expressions for ϕ, w, ϕr, ϕθ and ϕz can be further derived.

3. The Governing Equation for Pile–Soil Interaction

3.1. The Governing Equation of Pile

For large-diameter single piles, the relationship between ϕ and w(z) obtained from M(w) and P(ϕ) in Equation (16) is expressed as (0 ≤ zLp)
κ G p A p ( 2 w z 2 ϕ z ) + 2 t 2 w z 2 + k 1 ϕ z k w ρ p A p + n 1 2 w t 2 = 0 E p I p 2 ϕ z 2 + κ G p A p ( w z ϕ ) k 3 ϕ k 1 w z + 2 t 4 2 ϕ z 2 ρ p I p + n 2 2 ϕ t 2 = 0
where t = π 2 r p G s ( ϕ r 2 + ϕ θ 2 ) r d r d r ; k 1 = π r p { G s r p d ϕ z d r ϕ r + G s r p r ϕ z ϕ θ λ s r p ϕ z d ϕ r d r λ s r p ϕ z ϕ r ϕ θ r } r d r ; k = π r p { ( λ s + 2 G s ) ( d ϕ r d r ) 2 + 2 λ s ϕ r ϕ θ r d ϕ r d r + ( λ s + 2 G s ) ( ϕ r ϕ θ r ) 2 + G s ( ϕ r ϕ θ r + d ϕ θ d r ) 2 } r d r ; k 3 = π r p { [ G s d ϕ z d r d ϕ z d r + G s ϕ z r ϕ z r ] r p 2 r d r ; t 4 = π 2 r p ( λ s + 2 G s ) ϕ z 2 r p 2 r d r ; n 1 = π ρ s r p ( λ s + 2 G s ) ϕ z 2 r + G s ( ϕ r 2 + ϕ θ 2 ) r d r ; n 2 = π r p ρ s r p 2 ϕ z 2 r d r .
Equation (17) takes into account the dynamic response of the soil core within the pile. To solve Equation (17), it is necessary to introduce the following intermediate quantities:
w z , t = W z e i ω t ϕ z , t = ψ z e i ω t
where ω represents the frequency of the load and W and ψ are the vibration modes of pile displacement and rotation angle, respectively.
For large-diameter monopiles, the boundary conditions are crucial for accurately modeling their behavior. The boundary conditions typically depend on the specific load and environmental conditions. Here are the common boundary conditions for a large-diameter monopile:
κ G p A p w z ϕ + k 2 ϕ + 2 t w z z = 0 = F a E p I p ϕ z + k 1 + k 2 w + 2 t 4 ϕ z z = 0 = M a z = 0
where k 2 = π r p { G s { ϕ z r ϕ θ + d ϕ z d r ϕ r } r p r d r .
Based on the conditions of deformation and force continuity, the calculation method for the intermediate pile segment in the i-th layer (i = 1, 2, 3, …, n) is
w i , z = H i = w i + 1 , z = H i w z i , z = H i = w z i + 1 , z = H i κ G p A p w z ϕ + { k 2 ϕ + 2 t w z } i , z = H i = κ G p A p w z ϕ + { k 2 ϕ + 2 t w z } i + 1 , z = H i E p I p ϕ z + k 1 + k 2 w + 2 t 4 ϕ z i , z = H i = E p I p ϕ z + k 1 + k 2 w + 2 t 4 ϕ z i + 1 , z = H i 0 < z < L p
At the pile end:
κ G p A p w z ϕ + { k 2 ϕ + 2 t w z } z = L p = k 2 end ϕ + 2 t end w z z = L p E p I p ϕ z + k 1 + k 2 w + 2 t 4 ϕ z z = L p = k 1 end + k 2 end w + 2 t 4 end ϕ z z = L p z = L p
where t end = π 2 r p G s ( ϕ r 2 + ϕ θ 2 ) r d r d r + π 2 G s r p 2 ; t 4 , end = π 2 r p ( λ s + 2 G s ) ϕ z 2 r p 2 r d r + π 2 0 r p ( λ s + 2 G s ) ϕ z 2 r p 2 r d r k 1 , end = π r p { G s r p d ϕ z d r ϕ r + G s r p r ϕ z ϕ θ λ s r p ϕ z d ϕ r d r λ s r p ϕ z ϕ r ϕ θ r } r d r + π 0 r p { λ s r p ϕ z d ϕ r d r + 2 G s r p d ϕ z d r + G s r p r ϕ z } r d r ; k 2 , end = π r p { G s { ϕ z r ϕ θ + d ϕ z d r ϕ r } r p r d r + π 0 r p { G s { ϕ z r + d ϕ z d r } r p r d r .
If the pile end is considered to be fixed:
w z = L p = 0 ϕ z = L p = 0 z = L p
The corresponding governing equation for the soil below the pile bottom (z > Lp) is
2 t end 2 w z 2 + k 1 , end ϕ z k w n 1 , end 2 w t 2 = 0 k 3 , end ϕ k 1 , end w z + 2 t 4 , end 2 ϕ z 2 n 2 , end 2 ϕ t 2 = 0
where k 3 , end = π r p { [ G s d ϕ z d r d ϕ z d r + G s ϕ z r ϕ z r ] r p 2 r d r + π 0 r p { [ G s d ϕ z d r d ϕ z d r + G s ϕ z r ϕ z r ] r p 2 r d r ; t 4 , end = π 2 r p ( λ s + 2 G s ) ϕ z 2 r p 2 r d r + π 2 0 r p ( λ s + 2 G s ) ϕ z 2 r p 2 r d r ; and k 1 , end + k 2 , end = π r p [ λ s d ϕ r d r ϕ z + λ s ϕ r ϕ θ r ϕ z ] r p r d r + π 0 r p [ λ s d ϕ r d r ϕ z + λ s ϕ r ϕ θ r ϕ z ] r p r d r .
Under static load conditions, the control equation for the pile body is the expression mentioned above with the ω term removed, and the calculation method can be found in the literature [33,34,35,36,37,38,39].

3.2. The Governing Equation of Soil

Based on Q(ϕr) from Equation (16), the governing equation of ϕr can be given as follows:
d 2 ϕ r d r 2 + 1 r d ϕ r d r ( γ 1 r 2 + γ 2 r p 2 ) ϕ r = γ 3 2 r d ϕ θ d r γ 1 r 2 ϕ θ + γ 0 2 d ϕ z d r
where γ 0 = 0 G s ϕ r p d W d z r λ s ϕ z r p W r d z 0 ( λ s + 2 G s ) W W d z ; γ 1 = 0 ( λ s + 3 G s ) W W d z 0 ( λ s + 2 G s ) W W d z ; γ 2 = r p 0 G s d W d z d W d z d z ω 2 0 ρ soil W W d z 0 ( λ s + 2 G s ) W W d z ; γ 3 = 0 ( G s + λ s ) W W d z 0 ( λ s + 2 G s ) W W d z .
Based on R(ϕθ) from Equation (16), the governing equation of ϕθ can be given as follows:
d 2 ϕ θ d r 2 + 1 r d ϕ θ d r γ 4 r 2 ϕ θ γ 5 r p 2 ϕ θ = γ 6 2 r d ϕ r d r γ 4 r 2 ϕ r + γ 7 2 r ϕ z
where γ 4 = 0 ( λ s + 3 G s ) W W d z 0 G s W W d z ; γ 5 = r p 0 G s d W d z d W d z d z ω 2 0 ρ soil W W d z 0 G s W W d z ; γ 6 = 0 ( G s + λ s ) W W d z 0 G s W W d z ; γ 7 = 0 G s ψ r p d W d z λ s d ψ d z r p W d z 0 G s W W d z .
Based on S(ϕz) from Equation (16), the governing equation of ϕz can be given as follows:
d 2 ϕ z d r 2 + 1 r d ϕ z d r ( γ 9 2 + 1 r 2 ) ϕ z = γ 8 2 d ϕ r d r 1 r γ 8 2 ϕ r + 1 r γ 8 2 ϕ θ
where γ 8 = 0 λ s d ψ d r r p W G s ψ r p d W d z d z 0 G s ψ r p ψ r p d z ; γ 9 = 0 ( λ s + 2 G s ) d ψ d r r p d ψ d r r p d z ω 2 0 ρ soil ψ r p ψ r p d z 0 G s ψ r p ψ r p d z .
The calculation methods for Equations (24)–(26) are given in Appendix A. Under static load conditions, the control equation for the soil domain is the expression obtained by removing the ω term from the above Equations (24)–(26).

3.3. Programming Method

To solve the pile–soil interaction under dynamic (or static) loading conditions, a series of linear equations need to be calculated. These equations can be translated to solving CA = Λ−1 F, where CA is the global displacement vector, Λ is the global matrix which contains the stiffness matrix and the global mass matrix and F is the global load vector. The details about CA, Λ−1 and F are given in Appendix B. The structured approach to implementing the process in MATLAB (2021b) is presented in Figure 5. Therefore, the problem of pile–soil interaction under static and dynamic loads is essentially a matrix operation problem. Λ is mainly transformed through the following relationships:
(1) Mechanical assumptions of the pile and assumptions of soil displacement distribution (Equation (8));
(2) Constitutive models of the pile and soil (Section 2.2.3);
(3) Principles of force balance, deformation compatibility conditions and boundary conditions (Equations (17)–(23)).

4. Simplified Calculation Method of Natural Frequency

4.1. Pile Head Stiffness

After obtaining pile body displacement and rotation under the dynamic load state using the method introduced in the previous section, the pile head stiffness expression under each mode shape condition can be further obtained using Equation (27):
W 0 ψ 0 = K L K RL K LR K R 1 F a M a
where KL, KR and KLR represent the three spring stiffnesses for translation, rotation and coupling action at the mudline, respectively; W0 and ψ0 are the vibration modes of the pile displacement and rotation angle at z = 0, respectively.
According to the theoretical model of the cantilever beam [26], the methods for solving KL, KR and KLR are as follows:
K L K LR K RL K R = 1 w 1 F a 1 × d w 2 / d z M a 2 w 2 M a 2 × d w 1 / d z H a 1 × d w 2 / d z M a 2 w 2 M a 2 d w 1 / d z H a 1 w 1 F a 1
where w1 represents the horizontal displacement of the pile head when only the horizontal load Fa1 is applied; dw1/dz represents the rotation angle of the pile head when only the horizontal load Fa1 is applied; w2 represents the horizontal displacement of the pile head when only the overturning moment Ma2 is applied; and dw2/dz2 represents the rotation angle of the pile head when only the overturning moment Ma2 is applied. First, assume arbitrary values for the external force Fa1 and Ma1 = 0 and calculate the corresponding pile head displacement w1 and rotation angle dw1/dz according to the calculation process shown in Figure 5; then, reset and calculate the corresponding pile head displacement w2 and rotation angle dw2/dz when the external moment Ma2 is at an arbitrary value and Fa2 = 0, according to the calculation process shown in Figure 5. Finally, substituting the calculation results into Equation (28), the form of the pile head stiffness is obtained.

4.2. The Natural Frequency of Wind Turbines Considering Pile–Soil Interaction

In the analytical solution calculation of the natural frequency of wind turbines, the superstructure of the offshore wind turbine is equivalent to the structure shown in Figure 6, where the wind turbine and its blades are equivalent to a top mass block mRNAJ, the body of the turbine is equivalent to a uniform cross-section rod with equivalent elastic modulus ET, distributed mass mT, outer diameter DT and wall thickness tT; the part from the connection to above the pile mudline is equivalent to a uniform cross-section rod with outer diameter DP and wall thickness tP; and the pile–soil interaction is equivalent to the ECS model. The calculation steps are as follows:
(1)
Calculate the natural vibration frequency fFB when the wind turbine tower base is fixed [27,28]:
f FB = 1 2 π C MP 3 E T I T m RNA + α s m T L T 3
where CMP = (1/(1 + (1 + ρ3)λλ))0.5, λ = Ls/LT, ρ = ETIT/(EpIp) and αs varies depending on the tower section shape; here, αs = 33/140 [27].
(2)
Calculate KL, KRL and KR according to Equation (28). Use the procedure obtained from Figure 5 for the calculation, and then calculate ηL, ηRL and ηR according to the following equations:
η L = K L L p 3 E I η
η R = K R L p E I η
η LR = K LR L p 2 E I η
where EIη represents the equivalent bending stiffness of the superstructure [27,28].
(3)
Calculate the values of the coefficients CR and CL:
C R η L , η LR , η R = 1 1 1 + 0.6 η R η LR 2 η L
C L η L , η LR , η R = 1 1 1 + 0.5 η L η LR 2 η R
(4)
The final first-order natural frequency is given by
f 0 = C L C R f FB
where f0 represents the overall natural frequencies of the OWT system from the method proposed in this study.
Based on the previous procedures, detailed calculations and programming to calculate the pile–soil interaction under dynamic loading conditions are presented in Figure 5. If the geological survey report and wind turbine foundation parameters are known, the program can quickly adjust the design scheme for the single-pile foundation.

5. Results and Discussion

In this section, to verify the correctness of the calculation method presented in this study, 13 offshore wind projects in Europe and China from different references were collected. Based on existing cases from references [13,15,16,42,43,44,45,46,47], the values of the top mass block (mRNAJ), the equivalent elastic modulus of the tower (ET), the distributed mass of the tower (mT), the outer diameter of the tower (DT), the outer diameter of the tower at the end (Ds), the wall thickness of the tower (tT), the wall thickness of the tower at the end (ts), the outer diameter of the monopile (DP), the wall thickness of the monopile (tP) and the measured overall natural frequencies of the OWT system (f) from 13 projects are given in Table 1. The natural frequency of the OWT system with a fixed end (fFB) can also be given by Equation (29) if they are not in the references.
In Table 1, the formulas presented in Section 2.2.3 were used to describe the distribution of soil elastic modulus from 13 projects. The soil type in Kentish Flats and Hua Neng can be described by Equation (14), where multilayered soil was modeled. The elastic modulus Esi from Kentish Flats ranges from 20 MPa to 53.3 MPa, the details of each soil layer can be seen in reference [45]; Esi from Hua Neng ranges from 4 MPa to 60 MPa, the details can be seen in reference [34]. The soil type in Blyth, North Hoyle, Lely (A2) and Irene Vorrink can be described by Equation (13) with α = 0 and an initial value of the elastic modulus Es0, and the soil type in the rest of the projects can be described by Equation (13) with α = 1 and Es0. In these projects, the contact condition of the pile and soil is regarded as completely rough [48,49], which depends on assumption (3) in Section 2.1 and the displacement distribution functions in the r-, θ- and z-directions of the soil (Equations (9)–(11)).
Based on the above parameters presented in Table 1, ηL, ηR and ηRL from 13 projects were given by Equations (30), (31) and (32), respectively, and the results are presented in Table 2. Then, the values of the coefficients CR and CL were also calculated by Equations (33) and (34), respectively, and results are presented in Table 2. Therefore, the overall natural frequencies of the OWT system from the method proposed in this study (f0) were obtained by Equation (35), also presented in Table 2. To verify the correctness of the calculation method presented in this study, the error between the measured value (f) presented in Table 1 and the results from this study (f0) is given in Table 2. The real values of the overall natural frequencies of the OWT system (f) in Kentish Flats, Irene Vorrink, Lely (A2) and Barrow were given as range values in references [42,45]. To better compare the errors in different projects, Figure 7 presents the values of the error from 13 offshore wind projects in Table 2 with the error bar, where the horizontal ordinate presents the serial number of the projects, and the longitudinal ordinate presents the measured overall natural frequencies (which can also be regarded as the real value) of the OWT system. Therefore, the results from this study fall into the range shown with statistics in Figure 7.
During calculations, we found that the computation time using the proposed method was less than 10 s (on a workstation with Intel(R) Core(TM) i7-7700HQ CPU @2.80GHz (California, United States)), demonstrating the efficiency of the algorithm in this study. As shown in Table 1 and Table 2 and Figure 7, compared with the measured overall natural frequencies of the OWT system from references, the results obtained from this study are generally smaller. This means the pile–soil interaction described by the method in this study is slightly softer than the real situation. Compared with the condition that the pile–soil interaction is regarded as an ideal fixed boundary condition, the method in this study is much safer in the design process. The statistical errors in Table 2 and Figure 7 indicate that the error in the computed results can be controlled between −5% and 5%. Specifically, for NREL and Lely (A2), the errors between the calculated results and measured values were −11.2% to −8.4% and −4.5% to 10.7%, respectively. This conclusion further confirms the accuracy and reliability of the method proposed in this study.

6. Conclusions

In the design of offshore wind turbines, accurately and efficiently quantifying the impact of pile−soil interaction on the overall natural frequency of wind turbines is important. Few studies have comprehensively reflected the influence of pile−soil interaction on the dynamic response of OWTs based on continuum theory considering the three-dimensional displacement field of soil. Therefore, by combining Hamilton’s principle and the ECS model, an efficient analytical approach for the overall natural frequencies of wind turbines was obtained in this study. Thirteen different offshore wind projects were used to verify the correctness and efficiency of the proposed method. Compared with measured data, the proposed method for analyzing the overall natural frequencies of wind turbines achieves errors generally within the range of 10%. Although the pile−soil interaction described by the method in this study is slightly softer than the real situation, it is still much safer in the situation if the pile−soil interaction is regarded as an ideal rigid foundation. The results prove that a Timoshenko beam, which considers shear deformation of the monopile, and a proper soil constitutive relationship are necessary in the method, and soil is modeled as continuous element considering that the three-dimensional displacement of soil is also important.
This computational method provides a convenient, rapid and accurate means for predicting the overall natural frequencies of offshore wind turbines. However, the current derivation does not account for the separation between the soil and the pile, and soil mechanical properties may change under long-term cyclic loading. Therefore, a factor reflecting the separation of the pile−soil interface under large lateral loads and a new soil constitutive model reflecting stiffness degradation under long-term cyclic loading is one of the future directions for this research.

Author Contributions

Conceptualization, X.L.; Methodology, X.L.; Software, Q.Y.; Validation, M.Z.; Investigation, W.L.; Resources, X.L., X.Z. and L.Y.; Data curation, W.L.; Writing—original draft, W.L.; Writing—review and editing, X.Z. and L.Y.; Visualization, W.L. and L.Y. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Natural Science Foundation of China (NO.52201324) and the Natural Science Foundation of the Jiangsu Higher Education Institution of China (22KJB560015). The authors are grateful for their support.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are available upon request due to restrictions. The data presented in this study are available upon request from the corresponding author.

Conflicts of Interest

Authors Xufeng Zhao and Le Yang were employed by the company Xuzhou City Hongda Civil Engineering Laboratory Co., Ltd. The remaining authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

Abbreviations

The following symbols are used in this paper:
CR and CLCorrection factors considering rocking and sway effects of the pile–soil interaction, respectively;
DpThe outer diameter of the monopile;
DsThe outer diameter of the tower at the end;
DTThe outer diameter of the tower;
w/∂zThe rotation of the beam section;
EpThe Young’s modulus of the pile;
ETThe Young’s modulus of the tower;
Es0The initial elastic modulus of the soil at z = 0 m;
EsiThe elastic modulus of the ith layer of soil;
EThe equivalent bending stiffness of the superstructure;
fThe measured overall natural frequency of the wind turbine considering pile–soil interaction;
f0The overall natural frequencies of the OWT system from the method proposed in this study;
fFBThe overall natural frequency of the wind turbine when the tower base is fixed;
FaThe lateral force at the head of the beam;
IpThe second moment of inertia of the cross-section of the monopile;
KLThe lateral stiffness;
KRThe rocking stiffness;
KLRThe cross-coupling stiffness;
LpThe embedment depth of the beam;
mRNAThe mass of the rotor–nacelle assembly (kg);
MaThe moment at the head of the beam;
rpThe radius of the pile;
tThe wall thickness of the beam;
tpThe wall thickness of the tower;
tTThe wall thickness of the tower;
tsThe wall thickness of the tower at the end;
ur, uθ, uzThe displacement of soil at r-direction, θ-direction and z-direction;
UTThe total potential energy;
wThe lateral displacement of the beam’s central line;
WThe work exerted by external force;
αThe index of the function;
ρp, ρsThe density of the pile and soil (kg/m3), respectively;
γThe relative stiffness of the pile and soil;
The soil domain that participates in the structure–soil interaction;
ϕThe shear rotation of the plane section;
κThe shear correction factor;
σpqThe stress in the soil domain;
εpqThe strain in the soil domain;
λsi, GsiThe Lame constants of the ith layer of the multilayered continuum;
vsiThe Poisson ratio of the ith layer of soil;
ϕrDimensionless decay functions of the displacement components in the r-direction;
ϕθDimensionless decay functions of the displacement components in the θ-direction;
ϕzDimensionless decay functions of the displacement components in the z-direction;
ηL, ηR and ηLRCorrection factors of KL, KR and KLR, respectively.

Appendix A

The calculation methods of ϕr, ϕθ and ϕz are given as follows:
ϕ r = K ϕ r C K ϕ z 1 M ( B + C K ϕ z 1 N ) ( K ϕ θ F K ϕ z 1 N ) 1 ( E + F K ϕ z 1 M ) 1 A + ( B + C K ϕ z 1 N ) ( K ϕ θ F K ϕ z 1 N ) 1 D ϕ θ = ( K ϕ θ F K ϕ z 1 N ) 1 D + ( E + F K ϕ z 1 M ) ϕ r ϕ z = K ϕ z 1 M ϕ r + K ϕ z 1 N ϕ θ
where K ϕ r = 1 0 . . . . . . . . 0 0 1 0 . . . . . . . 0 . . . . . . . . . . . 0 . . 1 0 . . . . . . 0 . . 1 Δ r 2 1 r 1 2 Δ r 2 Δ r 2 + γ 1 r 2 + γ 2 r p 2 1 Δ r 2 + 1 r 1 2 Δ r                   1 Δ r 2 1 r 1 2 Δ r 2 Δ r 2 + γ 1 r 2 + γ 2 r p 2 1 Δ r 2 + 1 r 1 2 Δ r                                                                                           1 Δ r 2 1 r 1 2 Δ r 2 Δ r 2 + γ 1 r 2 + γ 2 r p 2 1 Δ r 2 + 1 r 1 2 Δ r             . . 0 0 1 ;
A = 1 1 . 1 0 0 . . . . 0 ; B = 0 0 . . . . . . . . 0 0 0 0 . . . . . . . 0 . . . . . . . . . . . 0 . . 0 0 . . . . . . 0 . . γ 3 2 r 1 2 Δ r γ 1 r 2 γ 3 2 r 1 2 Δ r                   γ 3 2 r 1 2 Δ r γ 1 r 2 γ 3 2 r 1 2 Δ r                                                                                           γ 3 2 r 1 2 Δ r γ 1 r 2 γ 3 2 r 1 2 Δ r             . . 0 0 0 ; C = 0 0 . . . . . . . . 0 0 0 0 . . . . . . . 0 . . . . . . . . . . . 0 . . 0 0 . . . . . . 0 . . γ 0 2 2 Δ r 0 γ 0 2 2 Δ r                   γ 0 2 2 Δ r 0 γ 0 2 2 Δ r                                                                                           γ 0 2 2 Δ r 0 + γ 0 2 2 Δ r             . . 0 0 0 ; D = 1 1 . 1 0 0 . . . . 0 ;
K ϕ θ = 1 0 . . . . . . . . 0 0 1 0 . . . . . . . 0 . . . . . . . . . . . 0 . . 1 0 . . . . . . 0 . . 1 Δ r 2 1 r 1 2 Δ r 2 Δ r 2 + γ 4 r 2 + γ 5 r p 2 1 Δ r 2 + 1 r 1 2 Δ r                   1 Δ r 2 1 r 1 2 Δ r 2 Δ r 2 + γ 4 r 2 + γ 5 r p 2 1 Δ r 2 + 1 r 1 2 Δ r                                                                                           1 Δ r 2 1 r 1 2 Δ r 2 Δ r 2 + γ 4 r 2 + γ 5 r p 2 1 Δ r 2 + 1 r 1 2 Δ r             . . 0 0 1 ; E = 0 0 . . . . . . . . 0 0 0 0 . . . . . . . 0 . . . . . . . . . . . 0 . . 0 0 . . . . . . 0 . . γ 6 2 r 1 2 Δ r γ 4 r 2 γ 6 2 r 1 2 Δ r                   γ 6 2 r 1 2 Δ r γ 4 r 2 γ 6 2 r 1 2 Δ r                                                                                           γ 6 2 r 1 2 Δ r γ 4 r 2 γ 6 2 r 1 2 Δ r             . . 0 0 0 ; K ϕ z = 1 0 0 . . . . . . 0 1 0 0 . . . . . 0 0 1 0 0 . . . . . . 0 1 0 . . . . . . 1 Δ r 2 1 r 1 2 Δ r 1 Δ r 2 + γ 9 2 + 1 r 2 1 Δ r 2 + 1 r 1 2 Δ r 0   . . . . . . . . . . . . . . . 0 1 Δ r 2 1 r 1 2 Δ r 1 Δ r 2 + γ 9 2 + 1 r 2 1 Δ r 2 + 1 r 1 2 Δ r 0 . . . . . 0 1 Δ r 2 1 r 1 2 Δ r 1 Δ r 2 + γ 9 2 + 1 r 2 1 Δ r 2 + 1 r 1 2 Δ r 0 . . 0 0 0 0 0 1 4 n × 4 n ;
M = 0 0 0 . . . . . . 0 0 0 0 . . . . . 0 0 0 0 0 . . . . 0 0 0 γ 8 2 2 Δ r γ 8 2 r γ 8 2 2 Δ r . . .   . . . . . . . . . . . . . . . . . . . . . 0 γ 8 2 2 Δ r γ 8 2 r γ 8 2 2 Δ r 0 . . . . . 0 γ 8 2 2 Δ r γ 8 2 r γ 8 2 2 Δ r 0 . . 0 0 0 0 0 1 4 n × 4 n ; N = 0 0 0 . . . . . . 0 γ 8 2 r 0 0 . . . . . 0 0 γ 8 2 r 0 0 . . . . 0 0 0 γ 8 2 r 0 . . . .   . . .   . . . . . . . . . . . . . . . . .   0 γ 8 2 r 0 0 . . . . .   0 γ 8 2 r 0 0 . . 0 0 0 0 0 0 ; G = 0 r 1 r p r 2 r p 1 0 . . . . 0 0 F = 0 0 . . . . . . . . 0 0 0 0 . . . . . . . 0 . . . . . . . . . . . 0 . . 0 0 . . . . . . 0 . . 0 γ 7 2 r 0                   0 γ 7 2 r 0                                                                                           0 γ 7 2 r 0             . . 0 0 0 .

Appendix B

Λ is given as follows:
Λ = Φ | z = 0 ( 1 ) 0 0 0 0 0 0 0 Φ | z = Δ z ( 1 ) Φ | z = Δ z ( 2 ) 0 0 0 0 0 0 0 Φ | z = 2 Δ z ( 2 ) Φ | z = 2 Δ z ( 3 ) 0 0 0 0 0 0 0 Φ | z = 3 Δ z ( 3 ) Φ | z = 3 Δ z ( 4 ) 0 0 0 0 0 0 0 Φ | z = 4 Δ z ( 4 ) Φ | z = 4 Δ z ( 5 ) 0 0 0 . . . . . . . . . . . . . Φ | z = L p 2 Δ z ( n 2 ) Φ | z = L p 2 Δ z ( n 1 ) 0 . . . . . 0 Φ | z = L p Δ z ( n 1 ) Φ | z = L p Δ z ( n ) 0 . . . . 0 0 Φ | z = L p ( n ) 4 n × 4 n
where Φ | z = 0 ( 1 ) = η 14 Φ 1 + η 15 Φ 1 η 17 Φ 1 + η 18 Φ 1 η 14 Φ 2 + η 15 Φ 2 η 17 Φ 2 + η 18 Φ 2 η 14 Φ 3 + η 15 Φ 3 η 17 Φ 3 + η 18 Φ 3 η 14 Φ 4 + η 15 Φ 4 η 17 Φ 4 + η 18 Φ 4 T z = 0 ; Φ | z = L p ( n ) = η 3 Φ 1 + η 4 Φ 1 2 t e n d η 1 Φ 1 η 7 Φ 1 + η 8 Φ 1 + η 9 Φ 1 + η 10 Φ 1 η 3 Φ 2 + η 4 Φ 2 2 t e n d η 1 Φ 2 η 7 Φ 2 + η 8 Φ 2 + η 9 Φ 2 + η 10 Φ 2 η 3 Φ 3 + η 4 Φ 3 2 t e n d η 1 Φ 3 η 7 Φ 3 + η 8 Φ 3 + η 9 Φ 3 + η 10 Φ 3 η 3 Φ 4 + η 4 Φ 4 2 t e n d η 1 Φ 4 η 7 Φ 4 + η 8 Φ 4 + η 9 Φ 4 + η 10 Φ 4 T ; and Φ | z = H i ( i ) = Φ 1 Φ 2 Φ 3 Φ 4 Φ 1 Φ 2 Φ 3 Φ 4 η 14 Φ 1 + η 15 Φ 1 η 14 Φ 2 + η 15 Φ 2 η 14 Φ 3 + η 15 Φ 3 η 14 Φ 4 + η 15 Φ 4 η 17 Φ 1 + η 18 Φ 1 η 17 Φ 2 + η 18 Φ 2 η 17 Φ 3 + η 18 Φ 3 η 17 Φ 4 + η 18 Φ 4 i ; Φ | z = H i ( i + 1 ) = Φ 1 Φ 2 Φ 3 Φ 4 Φ 1 Φ 2 Φ 3 Φ 4 η 14 Φ 1 + η 15 Φ 1 η 14 Φ 2 + η 15 Φ 2 η 14 Φ 3 + η 15 Φ 3 η 14 Φ 4 + η 15 Φ 4 η 17 Φ 1 + η 18 Φ 1 η 17 Φ 2 + η 18 Φ 2 η 17 Φ 3 + η 18 Φ 3 η 17 Φ 4 + η 18 Φ 4 i + 1 ;
η 2 = b a 1 2 t 4 , end k 3 , end 2 t end k 1 , end b 3 a 3 k 1 , end k 3 , end + 2 t 4 , end k 3 , end k k 1 , end b a ; η 3 = κ G p A p + k 2 k 2 , end 2 t end η 2 E p I p + 2 t 4 κ G p A p + 2 t κ G p A p k 1 κ G p A p + k 3 ρ p I p + n 2 ω 2 ; η 4 = κ G p A p + k 2 k 2 , end 2 t end η 2 E p I p + 2 t 4 k + ρ p A p + n 1 ω 2 + κ G p A p k 1 2 κ G p A p k 1 κ G p A p + k 3 ρ p I p + n 2 ω 2 + κ G p A p + 2 t ; η 6 = a 2 + b 2 2 t 4 , end k 3 , end ω 2 n 2 , end 2 t end k 1 , end b 3 a 3 k 2 t 4 , end k 1 , end 2 n 1 , end 2 t 4 , end ω 2 k 1 , end k 3 , end ω 2 n 2 , end b a ; η 7 = 2 t end 2 t 4 , end k 1 , end η 6 E p I p + 2 t 4 κ G p A p + 2 t κ G p A p k 1 κ G p A p + k 3 ρ p I p + n 2 ω 2 ; η 8 = E p I p + 2 t 4 κ G p A p k 1 κ G p A p + 2 t ; η 9 = 2 t end 2 t 4 , end k 1 , end η 6 E p I p + 2 t 4 k + ρ p A p + n 1 ω 2 + κ G p A p k 1 2 κ G p A p k 1 κ G p A p + k 3 ρ p I p + n 2 ω 2 ; η 10 = E p I p + 2 t 4 κ G p A p k 1 k ρ p A p + n 1 ω 2 + k 1 + k 2 k 1 , end + k 2 , end + 2 t end 2 t 4 , end k 1 , end η 5 2 k t 4 , end k 1 , end ; η 14 = κ G p A p + k 2 E p I p + 2 t 4 κ G p A p + 2 t κ G p A p k 1 κ G p A p + k 3 ρ p I p + n 2 ω 2 ; η 15 = κ G p A p + 2 t + κ G p A p + k 2 E p I p + 2 t 4 k + ρ p A p + n 1 ω 2 + κ G p A p + k 2 κ G p A p k 1 2 κ G p A p k 1 κ G p A p + k 3 ρ p I p + n 2 ω 2 ; η 17 = E p I p + 2 t 4 κ G p A p + 2 t κ G p A p k 1 ; η 18 = k 1 + k 2 + k + ρ p A p + n 1 ω 2 E p I p + 2 t 4 κ G p A p k 1 ; a = t ˜ + t ˜ 2 k ˜ , b = t ˜ t ˜ 2 k ˜ ; t ˜ = 2 t end k 3 , end k 2 1 , end + 2 t 4 , end k 2 t 4 , end n 1 , end + 2 t end n 2 , end ω 2 2 * 2 t 4 , end 2 t end (z > Lp); k ˜ = k k 3 , end + n 1 , end n 2 , end ω 4 k 3 , end n 1 , end + k n 2 , end ω 2 2 t 4 , end 2 t end (z > Lp).

References

  1. Yang, H.S.; Alkhabbaz, A.; Tongphong, W.; Tongphong, W.; Lee, Y.H. Cross-comparison analysis of environmental load components in extreme conditions for pontoon-connected semi-submersible FOWT using CFD and potential-based tools. Ocean Eng. 2024, 304, 117248. [Google Scholar] [CrossRef]
  2. Alkhabbaz, A.; Hamza, H.; Daabo, A.M.; Yang, H.S.; Yoon, M.; Koprulu, A.; Lee, Y.H. The aero-hydrodynamic interference impact on the NREL 5-MW floating wind turbine experiencing surge motion. Ocean Eng. 2024, 295, 116970. [Google Scholar] [CrossRef]
  3. Abdollah, M.; Soroosh, J.; Paul, D.; David, I. Foundation damping for monopile supported offshore wind turbines: A review. Mar. Struct. 2021, 77, 102937. [Google Scholar]
  4. Qin, W.; Dai, G.; Zhao, X.; Shu, G.; Gong, W. Experimental Investigation of CFFP-Soil Interaction in Sand under Cyclic Lateral Loading. Geotech. Test. J. 2019, 42, 1055–1074. [Google Scholar] [CrossRef]
  5. Qin, W.; Dai, G.; Gong, W.; Zhang, C. SH Experiments of Kaolin Clay Dynamic Responding Under Repeated Impact. China J. Highw. Transp. 2020, 33, 41–50. (In Chinese) [Google Scholar]
  6. Qin, W.; Li, X.; Dai, G.; Hu, P. Analytical Penetration Solutions of Large-Diameter Open-Ended Piles Subjected to Hammering Loads. J. Mar. Sci. Eng. 2022, 10, 885. [Google Scholar] [CrossRef]
  7. Qin, W.; Cai, S.; Dai, G.; Wang, D.; Chang, K. Soil Resistance during Driving of Offshore Large-Diameter Open-Ended Thin-Wall Pipe Piles Driven into Clay by Impact Hammers. Comp. Geotech. 2023, 153, 105085. [Google Scholar] [CrossRef]
  8. Qin, W.; Cai, S.; Dai, G.; Wei, H. Analytical solutions of soil plug behaviors in open-ended pile driven by impact load. Acta Geotech. 2023, 18, 4183–4194. [Google Scholar] [CrossRef]
  9. Wang, J.; Jin, T.; Qin, W.; Zhang, F. Performance evaluation of post-grouting for bored piles installed in inhomogeneous gravel pebble stratum overlaid by deep and soft soil layers. Geomech. Geoeng. 2024, 19, 77–95. [Google Scholar] [CrossRef]
  10. Qin, Q.; Gao, J.; Chang, K.; Dai, G.; Wei, H. Set-up effect of large-diameter open-ended thin-walled pipe piles driven in clay. Comp. Geotech. 2023, 159, 105459. [Google Scholar] [CrossRef]
  11. Wang, J.; Wu, X.; Qin, W.; Chang, K. Negative skin friction of piles installed in dredged slurry after being reinforced by the vacuum preloading method. Acta Geotech. 2023, 18, 6159–6174. [Google Scholar] [CrossRef]
  12. Byrne, B.W.; Burd, H.J.; Zdravković, L.; McAdam, R.A.; Taborda, D.M.; Houlsby, G.T.; Jardine, R.J.; Martin, C.M.; Potts, D.M.; Gavin, K.G. PISA: New design methods for offshore wind turbine monopiles. In Proceedings of the 8th International Conference for Offshore Site Investig. and Geotech, London, UK, 12–14 September 2017; pp. 11–17. [Google Scholar]
  13. Harvey, B.; Chreistelle, N.A.; Byron, W.B.; Houlsby, G.T. Application of the PISA Design Model to Monopiles Embedded in Layered Soils. Géotechnique 2017, 70, 1067–1082. [Google Scholar]
  14. Aranya, L.; Bhattacharyab, S.; Macdonalda, 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]
  15. Bouzid, D.A.; Bhattacharya, S.; Ostsmane, L. Assessment of natural frequency of installed offshore wind turbines using nonlinear finite element model considering soil-monopile interaction. J. Rock Mech. Geotech. Eng. 2018, 10, 333–346. [Google Scholar] [CrossRef]
  16. Bouzida, D.A.; Bhattacharyab, S.; Otsmanec, L. Natural frequency assessment of some installed OWTs using an efficient FE nonlinear model including soil/monopile interaction. J. Rock Mech. Geotech. Eng. 2017, 1–23. Available online: https://www.researchgate.net/publication/321109894 (accessed on 26 September 2024).
  17. Otsmanea, L.; Bouzid, D.A. An Efficient FE model for SSI: Theoretical background and assessment by predicting the response of large diameter monopiles supporting OWECs. Comp. Geotech. 2018, 97, 155–166. [Google Scholar] [CrossRef]
  18. Randolph, M.F. The response of flexible piles to lateral loading. Géotechnique 1981, 31, 247–259. [Google Scholar] [CrossRef]
  19. Pender, M.J. Aseismic pile foundation design analysis. Bull. N. Zealand Soc. Earthq. Eng. 1993, 26, 49–161. [Google Scholar] [CrossRef]
  20. Gazetas, G. Seismic response of end-bearing single piles. Int. J. Soil Dyn. Earthq. Eng. 1984, 3, 82–93. [Google Scholar] [CrossRef]
  21. Poulos, H.; Davis, E. Pile Foundation Analysis and Design; Rainbow-Bridge Book Co.: Palo Alto, CA, USA, 1980. [Google Scholar]
  22. Shadlou, M.; Bhattacharya, S. Dynamic stiffness of monopiles supporting offshore wind turbine generators. Int. J. Soil Dyn. Earthq. Eng. 2016, 88, 15–32. [Google Scholar] [CrossRef]
  23. Bhattacharya, S.; Adhikari, S. Experimental validation of soil-structure interaction of offshore wind turbines. Int. J. Soil Dyn. Earthq. Eng. 2011, 31, 805–816. [Google Scholar] [CrossRef]
  24. Wang, H.; Lehance, B.M.; Bransby, M.F.; Askarinejad, A.; Wang, L.Z.; Hong, Y. A simple rotational spring model for laterally loaded rigid piles in sand. Mar. Struct. 2022, 84, 103225. [Google Scholar] [CrossRef]
  25. Zha, X.; Guo, Z.; Wang, L.; Rui, S. A simplified model for predicting the accumulated displacement of monopile under horizontal cyclic loadings. Appl. Ocean Res. 2022, 129, 103389. [Google Scholar] [CrossRef]
  26. Houlsby, G.T.; Abadie, C.N.; Beuckelaers, W.J.A.P.; Byrne, B.W. A model for nonlinear hysteretic and ratcheting behaviour. Int. J. Solids Struct. 2017, 120, 67–80. [Google Scholar] [CrossRef]
  27. Trevor, M.L. Wind Energy Engineering; Elsevier Inc.: New York, NY, USA, 2017. [Google Scholar]
  28. Bhattacharya, S. Design of Foundations for Offshore Wind Turbines; Wiley: Hoboken, NJ, USA, 2019. [Google Scholar]
  29. Han, F.; Salgado, R.; Prezzi, M. Nonlinear analyses of laterally loaded piles-A semi-analytical approach. Comput. Geotech. 2015, 70, 116–129. [Google Scholar] [CrossRef]
  30. Han, F.; Prezzi, M.; Salgado, R. Energy-Based Solutions for Nondisplacement Piles Subjected to Lateral Loads. Int. J. Geomech. 2017, 17, 04017104. [Google Scholar] [CrossRef]
  31. Han, F.; Salgado, R.; Prezzi, M. A semi-analytical method for analysis of laterally loaded piles in elasto-plastic soil. In Proceedings of the 19th International Conference on Soil Mechanics and Geotechnical Engineering (Seoul), Seoul, Republic of Korea, 17–21 September 2017; pp. 2771–2774. [Google Scholar]
  32. Gupta, B.K.; Basu, D. Offshore wind turbine monopile foundations: Design perspectives. Ocean Eng. 2020, 213, 107514. [Google Scholar] [CrossRef]
  33. Li, X.; Dai, G.; Zhang, F.; Gong, W. Energy-based Analysis of Laterally Loaded Caissons with Large Diameters under Small-strain Conditions. Int. J. Geomech. 2022, 22, 05022005. [Google Scholar] [CrossRef]
  34. Li, X. Research on the Pile-Soil Interaction of Offshore Wind Farm Foundation Under Lateral Load; Southeast University: Nanjing, China, 2021. (In Chinese) [Google Scholar]
  35. Li, X.; Dai, G.; Zhu, M.; Gong, W. Application of static loading tests to steel pipe piles with large diameters in Chinese offshore wind farms. Ocean Eng. 2019, 186, 106041. [Google Scholar] [CrossRef]
  36. Li, X.; Dai, G.; Zhu, M.; Zhu, W.; Zhang, F. Investigation of the soil deformation around laterally loaded deep foundations with large diameters. Acta Geotech. 2023, 19, 2293–2314. [Google Scholar] [CrossRef]
  37. Li, X.; Dai, G. Closure to Energy-Based Analysis of Laterally Loaded Caissons with Large Diameters under Small-Strain Conditions. Int. J. Geomech. 2023, 8, 07023008. [Google Scholar] [CrossRef]
  38. Li, X.; Dai, G.; Zhu, M.; Wang, L.; Liu, H. A Simplified Method for Estimating the Initial Stiffness of Monopile—Soil Interaction Under Lateral Loads in Offshore Wind Turbine Systems. China Ocean Eng. 2023, 37, 165–174. [Google Scholar] [CrossRef]
  39. Li, X.; Zhu, M.; Dai, G.; Wang, L.; Liu, J. Interface Mechanical Behavior of Flexible Piles Under Lateral Loads in OWT Systems. China Ocean Eng. 2023, 3, 484–494. [Google Scholar] [CrossRef]
  40. Gupta, B.K.; Basu, D. Applicability of Timoshenko, Euler–Bernoulli and rigid beam theories in analysis of laterally loaded monopiles and piles. Géotechnique 2018, 68, 772–785. [Google Scholar] [CrossRef]
  41. Thomas, B. Small-Strain Stiffness of Soils and Its Numerical Consequences; Universität Stuttgart: Stuttgart, Germany, 2007. [Google Scholar]
  42. Arany, L.; Bhattacharya, S.; Macdonald, J.H.; Hogan, S.J. 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]
  43. Bhattacharya, S.; Cox, J.A.; Lombardi, D.; Wood, D.M. Dynamics of offshore wind turbines supported on two foundations. Proc. ICE-Geotech. Eng. 2013, 166, 159–169. [Google Scholar] [CrossRef]
  44. Júnior, C.V.; de Alencar Araújo, R.C.; de Souza, C.M.; Ferreira, A.C.; Ribeiro, P.M. A collocation method for bending, torsional and axial vibrations of offshore wind turbines on monopile foundations. Ocean Eng. 2020, 217, 107735. [Google Scholar] [CrossRef]
  45. Damgaard, M.; Ibsen, L.B.; Andersen, L.V.; Andersen, J.K.F. Cross-wind modal properties of offshore wind turbines identified by full scale testing. J. Wind Eng. Ind. Aerodyn. 2013, 116, 94–108. [Google Scholar] [CrossRef]
  46. Ko, Y.Y. A simplified structural model for monopile-supported offshore wind turbines with tapered towers. Ren. Energy 2020, 156, 777–790. [Google Scholar] [CrossRef]
  47. Adhikari, S.; Bhattacharya, S. Dynamic analysis of wind turbine towers on flexible foundations. Shock. Vib. 2012, 19, 37–56. [Google Scholar] [CrossRef]
  48. Johari, A.; Kalantari, A.R. System reliability analysis of soldier-piled excavation in unsaturated soil by combining random finite element and sequential compounding methods. Bull. Eng. Geol. Environ. 2021, 80, 2485–2507. [Google Scholar] [CrossRef]
  49. Johari, A.; Talebi, A. Stochastic analysis of piled-raft foundations using the random finite-element method. Int. J. Geomech. 2021, 4, 04021020. [Google Scholar] [CrossRef]
Figure 1. Schematic diagram of typical frequencies for offshore wind turbines (the overall natural frequency of the offshore wind turbine (OWT) system should be in the range of f1P to f3P).
Figure 1. Schematic diagram of typical frequencies for offshore wind turbines (the overall natural frequency of the offshore wind turbine (OWT) system should be in the range of f1P to f3P).
Jmse 12 01912 g001
Figure 2. Several typical models for pile–soil interaction ((a) the distributed spring model; (b) the equivalent fixed cantilever (EF) model; (c) the equivalent coupled spring model; (d) the “MRR” model; (e) the “p-y and MRR” model).
Figure 2. Several typical models for pile–soil interaction ((a) the distributed spring model; (b) the equivalent fixed cantilever (EF) model; (c) the equivalent coupled spring model; (d) the “MRR” model; (e) the “p-y and MRR” model).
Jmse 12 01912 g002
Figure 3. The mechanical model of pile–soil interaction for a monopile subjected to lateral loads in r-θ-z cylindrical coordinate system.
Figure 3. The mechanical model of pile–soil interaction for a monopile subjected to lateral loads in r-θ-z cylindrical coordinate system.
Jmse 12 01912 g003
Figure 4. Relationship between the normalized shear stiffness (G/G0) and shear strain (ε) of soil.
Figure 4. Relationship between the normalized shear stiffness (G/G0) and shear strain (ε) of soil.
Jmse 12 01912 g004
Figure 5. The flow chat to implement the pile–soil interaction under dynamic loading conditions in MATLAB (2021b).
Figure 5. The flow chat to implement the pile–soil interaction under dynamic loading conditions in MATLAB (2021b).
Jmse 12 01912 g005
Figure 6. The offshore wind turbine model with the pile–soil interaction being equivalent to the ECS model.
Figure 6. The offshore wind turbine model with the pile–soil interaction being equivalent to the ECS model.
Jmse 12 01912 g006
Figure 7. Statistical analysis of 13 offshore wind projects in Table 2 with error bars.
Figure 7. Statistical analysis of 13 offshore wind projects in Table 2 with error bars.
Jmse 12 01912 g007
Table 1. The related parameters of offshore wind turbines and the measured overall natural frequencies of the OWT system (f) from 13 projects.
Table 1. The related parameters of offshore wind turbines and the measured overall natural frequencies of the OWT system (f) from 13 projects.
No.12345678910111213
LiteratureBlyth
[44]
Belwind
[44]
Walney
[44]
NREL
[44]
Kentish Flats
[45]
North Hoyle
[47]
Lely (A2)
[42]
Irene Vorrink [42]Barrow
[42]
Thanet
[42]
BurboBank
[42]
Gunfleet
[42]
Hua Neng [34]
Wind power (MW)233.6520.50.6
mRNAJ (t)80130.8234.53501321003235.7130.8130.8234.5234.5243
LT (m)54.55367.387.660.0667.337.944.55854.1666086
Ds (m)4.254.3564.4543.23.54.454.3555.5
DT (m)2.752.333.871.92.31.91.72.32.3333.1
ts (mm)34284027263513143236283360
tT (mm)342840191535 83236283360
Ep (GPa)210210210210210210210210210210210210210
mT (t)159120260347.4610813031.4435.7153160180193269.75
Ls (m)16.53737.32016712.15.23341.122.82844
Pile diameter Dp (m) 3.55664.343.23.54.754.74.756.1
Wall thickness tp (mm)506080604550352845–806545-7535-5070
Lp (m)153523.54029.53313.51930.2–40.725-30243844.3
Es0 (MPa)42.53.7217.437.1Multilayer soils
20~53.3
382.542.542.53.723.723.7217.4Multilayer soils
4~60
α01110001111
fFB (HZ) 0.5140.4010.380.5920.380.3640.7130.583~0.5860.3870.4020.3220.3520.388
f (HZ)0.4880.3720.350.546/0.5630.3390.350.634~0.7350.553~0.5600.3690.370.2920.3140.325
Table 2. The overall natural frequencies of OWT system from the method proposed in this study (f0) and their errors compared with measured data.
Table 2. The overall natural frequencies of OWT system from the method proposed in this study (f0) and their errors compared with measured data.
No.12345678910111213
ηL1983.792437.03283.683635.194019.5913739.26324.08644.322594.142037.402743.561757.82442.38
ηR199.5325.92148.681814.1533.2949.2722.4299.6823.0420.8018.8214.5922.04
ηLR−118.46−149.64−0.51183−135.05−217.81−472.66−0.77−8.55−145.56−122.56−135.28−95.34−75.60
CR0.9720.90940.9889150.84570.92800.95190.91810.99740.89920.88960.87930.84960.8456
CL0.9650.97120.9929990.99950.99950.99990.99490.98030.99920.99900.99930.99890.9955
f0
(HZ)
0.4820.3540.3410.5000.3520.3460.7020.5400.3480.3570.2830.2990.327
Error (%)−1.2−4.8−2.6−11.2~ −8.43.8−1.1−4.5~10.7−3.6~ −2.4−5.7−3.5−3.1−4.80.6
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Li, W.; Li, X.; Zhao, X.; Yin, Q.; Zhu, M.; Yang, L. The Method of the Natural Frequency of the Offshore Wind Turbine System Considering Pile–Soil Interaction. J. Mar. Sci. Eng. 2024, 12, 1912. https://doi.org/10.3390/jmse12111912

AMA Style

Li W, Li X, Zhao X, Yin Q, Zhu M, Yang L. The Method of the Natural Frequency of the Offshore Wind Turbine System Considering Pile–Soil Interaction. Journal of Marine Science and Engineering. 2024; 12(11):1912. https://doi.org/10.3390/jmse12111912

Chicago/Turabian Style

Li, Wei, Xiaojuan Li, Xufeng Zhao, Qian Yin, Mingxing Zhu, and Le Yang. 2024. "The Method of the Natural Frequency of the Offshore Wind Turbine System Considering Pile–Soil Interaction" Journal of Marine Science and Engineering 12, no. 11: 1912. https://doi.org/10.3390/jmse12111912

APA Style

Li, W., Li, X., Zhao, X., Yin, Q., Zhu, M., & Yang, L. (2024). The Method of the Natural Frequency of the Offshore Wind Turbine System Considering Pile–Soil Interaction. Journal of Marine Science and Engineering, 12(11), 1912. https://doi.org/10.3390/jmse12111912

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