Next Article in Journal
Safety Leadership: A Catalyst for Positive Safety Climate on Construction Sites
Next Article in Special Issue
Effect of Drying and Wetting Cycles on the Surface Cracking and Hydro-Mechanical Behavior of Expansive Clays
Previous Article in Journal
Role of the Subgrade Reaction Modulus in the Design of Foundations for Adjacent Buildings
Previous Article in Special Issue
Experimental Study on the Horizontal Bearing Characteristic of a Strip-Walled Underground Diaphragm Wall
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

The Simplified Method of Head Stiffness Considering Semi-Rigid Behaviors of Deep Foundations in OWT Systems

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
School of Civil Engineering, Southeast University, Nanjing 211189, China
*
Author to whom correspondence should be addressed.
Buildings 2024, 14(6), 1803; https://doi.org/10.3390/buildings14061803
Submission received: 19 April 2024 / Revised: 25 May 2024 / Accepted: 7 June 2024 / Published: 14 June 2024

Abstract

:
Simplified methods of static free head stiffness of the semi-rigid foundation under lateral loads were limited to flexible or rigid behavior by the critical length of piles. This would lead to errors when predicting the static or dynamic performance of their upper structures in OWT Systems. This paper presents a comprehensive analysis of the head static stiffness of the semi-rigid pile without considering the critical length. Firstly, case studies using the energy-based variational method encompassing nearly twenty thousand cases were conducted. These cases involved different types of foundations, including steel pipe piles and concrete caissons, in three types of soil: homogeneous soil, linearly inhomogeneous soil, and heterogeneous soil. Through the analysis of these cases, a series of polynomial equations of three kinds of head static stiffness, containing the relative stiffness of the pile and soil, the slenderness ratio, and Poisson’s ratio, were developed to capture the semi-rigid behavior of the foundations. Furthermore, the lateral deflection, the rotation for concrete caissons in the bridge projects, and several natural frequencies of three cases about the OWT system considering the SSI effect were carried out. the error of high-order frequency of the OWT system reached 13% after considering the semi-rigid effect of the foundation.

1. Introduction

Monopiles and well foundations (also known as caissons), with large diameters (2 m ≤ D ≤ 10 m) and smaller slender ratios (2 ≤ Lp/D ≤ 10, Lp is the length of the foundation), are extensively employed in offshore structures or bridges (Figure 1). The circular cross-sections of these foundations are designed as massive hollow bodies to comply with mechanics and reduce costs [1,2,3]. These foundations carry mechanisms of lateral loads to the surrounding soils caused by wind, waves, and earthquakes, and bring great challenges to predicting the supporting performance. It is important to note that the laterally loaded foundation belongs to a semi-rigid beam, which both covers rigid and flexible behavior (Figure 1). The soil-structure interaction (SSI) effect can lead to changes in the static or dynamic mechanical behavior of the upper structures, and the effect would be much more obvious when involving high-order derivative operations [4,5,6,7]. On the other hand, Structural monitoring or real-time analysis of dynamics leads to an urgent need for a more accurate expression of the SSI effect. Therefore, to maintain the stability and serviceability of a structure, it is imperative to accurately predict the foundation’s free-head stiffness under lateral loads. This precision is particularly critical for preserving the accuracy of upper structure results during high-order derivative mechanical calculations, especially when accounting for the SSI effect.
The free head stiffness of a foundation under lateral loads is commonly characterized by three parameters: KL (lateral stiffness), KR (rocking stiffness), and KLR (cross-coupling stiffness). The critical length of piles, typically determined as a function of the relative stiffness between the pile and soil, plays a pivotal role in influencing foundation performance [8,9,10]. Depending on the specified threshold value for the critical length, various beam theories are employed to model the foundation’s bending behavior. Consequently, different equations for KL, KR, and KLR are utilized. These stiffness parameters, particularly their initial values, serve two key purposes: firstly, they are instrumental in predicting the foundation’s head deflection and rotation under static loading conditions; secondly, they are essential for factoring in the Soil-Structure Interaction (SSI) effect during dynamic analysis, particularly within the framework of a Serviceability Limit State (SLS).
To characterize the head stiffness of laterally loaded piles, researchers have explored several approaches over the decades, including closed-form solutions based on Winkler-based theory, semi-analytical analyses grounded in continuum theory, and experimental testing. Table 1 and Table 2 provide a compilation of key findings from these studies. For example, Bergfelt [11] proposed a closed function of KL based on the long flexible beam in a Winkler foundation. Building upon this work, Thomas [12] provided a solution considering the relative stiffness of the pile and soil. Randolph [13] developed the finite element method to modify the 3-D continuum behavior of the soil and put forward equations for KL, KR, and KLR in flexible and rigid beams based on the relative stiffness of the pile and soil (Eeq/Es). Additionally, Poulos and Davis [14], following Barber [15], presented a series of functions for KL, KR, and KLR, considering flexible and rigid beams with the threshold value of β. Other researchers, such as Budhu and Davies [16], Pender [17], Gazetas [18], Higgins et al. [19], Shadlou and Bhattacharya [20], Aissa et al. [21]. and Anoyatis et al. [22], have proposed expressions for flexible piles and rigid piles. These studies collectively yield dimensionless expressions for KL, KLR, and KR, which are denoted as KL/(EsD), KLR/(EsD2), and KR/(EsD3), respectively. Notably, these investigations reveal that the dimensionless head stiffness of long, flexible piles is primarily influenced by the parameter Eeq/Es, while for short piles, the slenderness ratio (Lp/D) assumes a significant role. However, the foundations examined in this study primarily exhibit semi-rigid beam characteristics, demonstrating aspects of both rigid and flexible behavior. Specifically, when predicting the static or dynamic performance of upper structures—particularly those involving high-order derivative operations like high-order self-resonant frequencies—traditional methods may yield results with significant, non-negligible errors [23]. Therefore, there is a compelling need to develop a more precise equation for calculating the dimensionless head stiffness—those that consider both the rigid and flexible behavior of the foundation while remaining as practical to calculate as the methods employed in previous studies.
The energy-based variational method, characterized by its energy-based theory and virtual work principles, has emerged as a valuable approach for predicting the response of piles subjected to lateral loads. This method offers notable advantages, combining the efficiency of quick computations with the mathematical rigor comparable to that of three-dimensional finite-difference analyses [30,31,32]. Previously, Shadlou et al. [20] conducted an energy-based study that involved the modification of pile-head spring stiffness (KL, KR, and KLR) using the Euler-Bernoulli beam theory in layered soils. As highlighted in previous research (Byron & Houlsby [33,34]; Gupta [35,36,37], Cao [38]), the Timoshenko beam model has gained recognition for its appropriateness in characterizing the response of laterally loaded piles with substantial diameters. More intricate models, such as shells or solid bodies, have found application in commercial software like ABAQUS, FLAC3D, or PLAXIS. A consensus has emerged within the research community, affirming that the Timoshenko beam strikes a commendable balance between accuracy and simplicity. Gupta [37] and Gupta and Basu [36] introduced the Timoshenko beam theory within the energy-based variational method, discussing pile-head spring stiffness under both dynamic and static lateral loads with no specific calculation equations provided in this work; Taking the method further, Li [39,40,41,42,43] extended its application by considering both vertical and horizontal soil displacement in the analysis of laterally loaded deep foundations.
The objective of this research is to investigate the head static stiffness of the semi-rigid pile without the critical length. To achieve this, a comprehensive study using the energy-based variational method with Timoshenko beam encompassing nearly twenty thousand cases was conducted. By using the efficiency of the variational method, different types of foundations (steel pipe piles and concrete caissons) with diameters and slenderness ratios (Lp/D) ranging from 2 m to 10 m, and three types of soil (homogeneous soil, linearly inhomogeneous soil, and heterogeneous soil) were studied, each characterized by varying elastic moduli and Poisson’s ratio. Through the analysis of these cases, the three kinds of static stiffness of the pile head, covering the relative stiffness of the pile and soil, the slenderness ratio as well and Poisson’s ratio without consideration of the critical length, were developed to accurately capture the semi-rigid behavior exhibited by these foundations. These new equations can ensure the accuracy of mechanical calculation involving high-order derivative operations of the upper structures.

2. Methodology

2.1. Energy-Based Variational Method

The energy-based variational method has the advantage of computational efficiency and open accessibility. To investigate the head stiffness of a semi-rigid pile, we examine a single circular beam, which may represent either a pile or a caisson. This beam has specific geometric parameters, including a radius (rp), wall thickness (t), and length (Lp), and it is embedded within a three-dimensional continuous medium. The beam experiences lateral forces (Fa) and/or moments (Ma) at its head (Figure 1). In this analysis, we treat the pile as a vertical Timoshenko beam, characterized by a lateral deflection denoted as w(z), which is associated with the depth parameter z. and the shear rotation of the plane section is ϕ(z), the three-dimensional soil displacement distribution can be simplified as below in terms of r-θ-z:
u r = w z ϕ r r cos θ u θ = w z ϕ θ r sin θ u r = ϕ z r p ϕ z r cos θ
The boundary conditions of ϕr, ϕθ, and ϕz are given as:
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 r
where ϕz(rp) is the value of ϕz(r) when r = rp. In Equation (2), the functions ϕr and ϕθ both have a value of 1 when 0 ≤ rrp and 0 when r approaches infinity. As for ϕz, it equals 0 when r is 0 and r/rp when 0 < r < rp, ϕz becomes 0 again as r approaches infinity. These functions are related to the variables r, θ, and the pile radius rp. In Equation (2), it is further assumed that ϕr, ϕθ, and ϕz are mutually independent in the r-direction. The boundary condition of ϕz at 0 ≤ rrp is from the static Timoshenko beam theory (Timoshenko, 1932; https://ccrma.stanford.edu/~bilbao/master/node163.html, accessed on 16 February 2024.), where the displacements of the beam are ur = w(z)cosθ, uθ = −w(z)sinθ, and uz = ϕ(z)cosθ.
For Timoshenko beam theory, depending on the total energy of the system and the principle of virtual work, the following equation is obtained:
δ = E p I p 0 L p ( d ϕ d z ) δ ( d ϕ d z ) d z + 0 L p κ G p A p ( d w d z ϕ ) δ ( d w d z ϕ ) d z + 0 L p 0 2 π r p σ p q δ ε p q r d r d θ d z + L p 0 2 π 0 σ p q δ ε p q r d r d θ d z δ W
where Ep is Young’s modulus of the beam; Ip is the second moment of inertia of the cross-section; κ is the shear correction factor; w is the lateral displacement of the beam central line; W is the work by outer force (lateral force Fa or moment Ma). The soil potential energy is in the region of r > rp when Lpz ≥ 0 and r > 0 when zLp; σpq is the stress in soil domain; εpq is the strain in soil domain.

2.2. Soil Conditions

To obtain the initial stiffness, the soil is in an elastic state under very small to small strain conditions, and the stress-strain relationship is elastic. In this section, four main categories are considered from the point of view of the current analysis:
(1) Homogenous soil conditions (Es = Es0(z/D)α, α = 0): Es0 is the initial value of soil elastic modulus, α is an index of the function. The elastic modulus is considered constant with a depth of the soil, often used in cohesive soils, weathered bedrock, and very dense sand for typical North Sea soils [7,44].
(2) Lineally inhomogeneous soil conditions (Es = Es0(z/D)α, α = 1): Gs increases linearly with depth from zero value at the ground surface, also called Gibson’s soil, often used to describe normally consolidated cohesive soils [7], London clay reported by Skempton and Henkel [45], see also, Ward et al. [46], Burland and Lord [47], Butler [48], and Hobbs [49].
(3) Heterogeneous soil conditions (Es = Es0(z/D)α, 0 < α < 1): Gs increases nonlinearly with depth from zero value at the ground surface; when α = 0.5, it is described as a parabolic soil profile, which has been used to describe soil profiles in several Europe offshore wind projects [7,50].
(4) Complex layered soil conditions: Es is obviously not included in the above categories, different layers with obviously different Es are often observed, and often used in real sites, the multilayered sedimentation in river or coast areas, i.e., interlaced layers with soft clay and sand. There are other soil conditions characterized by heterogeneity, such as G = G1 + G2 cos(z/D). These conditions often arise in geomaterials formed through periodic deposition, such as varved clays and sedimentary rocks [51].
This paper primarily centers its attention on deep foundation scenarios that pertain to bridges and offshore structures. These scenarios are categorized within groups (1) to (3). In elastic media, the constitutive model is σpq = λs(z)δpqεpp +2Gs(z)εpq (p, q = 1, 2, 3), where λs(z) and Gs(z) are Lame’s constants; λs = Esvs/(1 + vs)/(1 − 2vs); Gs = Es/2/(1 + vs); Es is the elastic modulus; vs is Poisson’s ratio. Therefore, λs in categories (1) to (3) can be given by:
λ s = E s 0 ν s z / D α 1 + ν s 1 2 ν s

2.3. Governing Differential Equations of the Beam and Soils

The governing differential equations of the foundation and soils have been derived in the Appendix A and Appendix B, the governing differential equations of soils have been derived in the Appendix C. The stiffness of a beam under lateral loading is expressed in matrix form as follows:
K L K L R K R L K R w z = 0 d w d z z = 0 = F M
where KL is the swaying dynamic head stiffness, KLR is the coupled swaying-rocking head stiffness, and KR is the rocking dynamic head stiffness KLR = KRL.
To calculate KL, KLR, and KR, three steps are needed:
(1) A response analysis with applied lateral force Fa and restricted pile head rotation θ to derive the lateral stiffness coefficient (KL = Fa/w0), the boundary conditions at the pile head are Fa = constant, θ0 = 0, w0 is unknown, and
κ G p A p d w d z ϕ + k 2 ϕ + 2 t d w d z = F a ( z   =   0   m ) θ 0 = d w d z = 0
where t and k2 are shown in Appendix A.
(2) A response analysis for calculating to moment Ma that leads to zero rotation θ for a given pile head displacement w0. This defines the off-diagonal term (KLR = −Ma/w0), the boundary conditions at the pile head are w0 = constant, θ0 = 0, Ma is unknown, and
w 0 = constant ( z   =   0   m ) θ 0 = 0
(3) A response analysis with zero pile head deflection w0 and an applied bending moment Ma. The rotational stiffness coefficient can be derived from the observed pile head rotation (KR = −Ma/θ) the boundary conditions at the pile head are Ma = constant, w0 = 0, θ0 is unknown, and
w 0 = 0 ( z   =   0   m ) E p I p d ϕ d z + k 1 + k 2 w + 2 t 4 d ϕ d z = M a
where k1, k2 and t4 are shown in Appendix A.
From Equations (1)–(8), the values of KL, KLR, and KR can be obtained, respectively. The derivation details are given in Appendix B.

2.4. Modeling Cases

The values of KL, KLR, and KR are influenced by a multitude of factors, including the diameter of the pile (D), the slenderness ratio (Lp/D), the thickness of the pile wall (t), the relative stiffness of the pile and soil, the elastic modulus of the soil (Es), and Poisson’s ratio (vs). To comprehensively explore the impact of these parameters, a wide range of values has been considered in the analyses conducted for this paper, as summarized in Table 3. For instance, the wall thickness for monopiles typically falls within the range of D/80 to D/120, equivalent to 0.005D to 0.015D [14]. In the case of concrete caissons, the wall thickness is in the range of 0.05D to 0.15D. Additionally, the elastic modulus of the soil (Es0) has been studied within the range of 2 MPa to 300 MPa, Poisson’s ratio (vs) spans from 0.20 to 0.45. In fact, this study has examined over twenty thousand cases to comprehensively investigate the relationships and effects of these parameters on KL, KLR, and KR. These cases encompass a wide array of deep foundation scenarios pertinent to bridges and offshore structures, ensuring a thorough exploration of the subject matter.

3. Results and Discussion

3.1. Validation of the Analysis Compared against Different Methods

Homogenous soil conditions (α = 0) To investigate the influence of ln(Eeq/Es) and Lp/D on the head stiffness of laterally loaded foundations in homogeneous soil conditions, we present Figure 2, Figure 3 and Figure 4 as illustrative examples. These figures depict the effects of ln(Eeq/Es) and Lp/D on KL/(EsD), KLR/(EsD2), and KLR/(EsD3) for steel monopiles, considering t = 0.01D and vs = 0.30 in 3-D version. The results are compared with closed-form solutions from Randolph [13], Shadlou et al. [20], Higgins et al. [19], and Abed et al. [29] presented in Table 1 and Table 2.
As presented in previous studies, Randolph [13] and Carter et al. [27] respectively proposed the critical pile length of slender piles and rigid piles as:
L p D E e q E s 1 + 0.75 v s / 1 + v s 2 / 7 flexible   piles 0.05 E e q E s 1 + 0.75 v s / 1 + v s 1 / 2 rigid   piles
Then Higgins et al. [19] proposed a method for determining flexible pile and rigid piles with the consideration of:
L p D 40 flexible   piles 1 412.8 E e q E s 1 + 0.75 v s / 1 + v s 1 / 3.23 rigid   piles ( α   =   0 )
and
L p D 1 650 E e q d E s dz D 1 + 0.5 v s / 1 + v s 1 / 3.45 rigid   piles ( α   =   1 )
Figure 3a illustrates a similar relationship between KLR/(EsD2) and both (Lp/D) and ln(Eeq/Es) in the 3-D context. Figure 3b highlights the trend of decreasing values of KLR/(EsD2) with increasing Lp/D for short piles (2 ≤ Lp/D ≤ 6, ln(Eeq/Es) > 4.5). These values are consistent with those from Randolph [13] and Shadlou et al. [20] for rigid piles, falling within the range of KLR/(EsD2) values obtained in the present study. However, the rate of increase differs: the former exhibits a positive increasing rate concerning (Lp/D), while the results from this study indicate that KLR/(EsD2) becomes insensitive to Lp/D when Lp/D ≥ 6, suggesting a flexible pile behavior. The successful prediction of the threshold zone where KLR/(EsD2) becomes insensitive to Lp/D by Higgins et al. [19] further supports this observation. In Figure 3c, the equations provided by Shadlou et al. [20] and Randolph [13] for flexible piles in Table 1 also demonstrate good performance on KLR/(EsD2) when Lp/D ≥ 6.
A similar phenomenon is also evident in Figure 4, where the relationship between KR/(EsD3) and (Lp/D) and ln(Eeq/Es) in the 3-D context has been analyzed. This phenomenon has also been observed in the case of concrete caissons.
Based on the findings presented in Figure 2, Figure 3 and Figure 4, it can be concluded that while traditional methods demonstrate good performance for KL/(EsD), KLR/(EsD2), and KR/(EsD3), it is essential for the functions of KL/(EsD), KLR/(EsD2), and KR/(EsD3) to incorporate the parameters Eeq/Es and Lp/D. Therefore, the development of more accurate equations that consider both parameters is warranted. These observations highlight the need for more comprehensive and refined equations that account for a broader range of influencing factors, ultimately improving our ability to predict the head stiffness of laterally loaded foundations in various scenarios.
Linearly inhomogeneous soil conditions (α = 1). To comprehend the impact of ln(Eeq/Es) and Lp/D on the head stiffness of laterally loaded foundations in linearly inhomogeneous soil conditions, Figure 5, Figure 6 and Figure 7 provide insights into the effects of these parameters on KL/(EsD), KLR/(EsD2), and KLR/(EsD3) for steel monopiles with t = 0.01D and vs = 0.30. In Figure 5a, it is evident that the results presented in this study are influenced by ln(Eeq/Es) and Lp/D. Figure 5b demonstrates that KL/(EsD) for short piles (2 ≤ Lp/D ≤ 4, ln(Eeq/Es) ≥ 6) exhibits an increasing trend with rising Lp/D values, aligning well with the results from Shadlou et al. [20], Higgins et al. [19], Abed et al. [29] for rigid pile conditions in Table 2. Conversely, results from Randolph [13] tend to underpredict KL/(EsD) at smaller Lp/D values, while those from Higgins et al. [19], Shadlou et al. [20], and Abed et al. [29] tend to overpredict it at larger Lp/D values. In Figure 5c, it becomes evident that KL/(EsD) becomes less sensitive to Lp/D as ln(Eeq/Es) decreases and Lp/D increases. This behavior indicates a flexible pile response. The threshold zone where KL/(EsD) becomes insensitive to Lp/D is successfully predicted by Higgins et al. [19], and equations from Pender [17] and Shadlou et al. [20] for flexible piles perform better in terms of prediction accuracy compared to the equation from Randolph [13].
Figure 6 demonstrates a similar relationship between KLR/(EsD2) and both (Lp/D) and ln(Eeq/Es) in the 3-D context. In Figure 6a, it’s evident that the results presented in this study are influenced by ln(Eeq/Es) and Lp/D. Figure 6b illustrates that for short piles (2 ≤ Lp/D ≤ 4, ln(Eeq/Es) > 6), the values of KLR/(EsD2) exhibit a curvilinear decrease as Lp/D increases. The range of Lp/D values (2 ≤ Lp/D ≤ 4) covered by the results from the present study for KLR/(EsD2) includes the results from Randolph [13], Higgins et al. [19], Shadlou et al. [20], and Abed et al. [29] for rigid pile conditions. Similar to the previous case, as Lp/D increases beyond 6, KLR/(EsD2) becomes insensitive to Lp/D, with the threshold zone falling between the predictions of Randolph [13] and Higgins et al. [19]. In Figure 6c, the results from this paper at Lp/D = 3 align more closely with the predictions for flexible piles by Randolph [13] (as given in Table 1), while the results from the present study for Lp/D ≥ 6 better match the equations from Shadlou et al. [20] and Pender [17] for flexible piles (as given in Table 1). A similar trend is observed in Figure 6, where KR/(EsD3) versus (Lp/D, ln (Eeq/Es)) in the 3-D version was analyzed, and the same phenomenon has also been observed in concrete well foundations (or caissons).
Based on the insights provided in Figure 5, Figure 6 and Figure 7, it becomes evident that the relationships governing KL/(EsD), KLR/(EsD2), and KR/(EsD3) should be expressed as functions of both Eeq/Es and Lp/D when dealing with linearly inhomogeneous soils. This underscores the need for more precise equations that take into account the combined influence of these two parameters. It’s worth noting that the threshold zones for long flexible piles and short rigid piles in linearly inhomogeneous soils differ from those observed in homogeneous soils, indicating the importance of considering soil heterogeneity in foundation analysis.
Heterogeneous soil conditions (α = 0.5). Figure 8, Figure 9 and Figure 10 provide insights into the impact of ln (Eeq/Es) and Lp/D on KL/(EsD), KLR/(EsD2), and KLR/(EsD3) for steel monopiles with t = 0.01D and vs = 0.30. In Figure 8a, it is evident that the results presented in this study are influenced by ln(Eeq/Es) and Lp/D. Figure 8b shows that the values of KL/(EsD) for short piles exhibit an increasing trend with rising Lp/D values, and the results at Lp/D = 2 and ln(Eeq/Es) = 9.08 align well with those of Shadlou et al. [20] for rigid pile conditions, as presented in Table 1 and Table 2. In Figure 8c, KL/(EsD) becomes less sensitive to Lp/D as ln(Eeq/Es) decreases and Lp/D increases, indicating a foundation behavior more akin to a flexible pile. The values from the present study at Lp/D = 2 are closer to those from Syngros [26], and the values at Lp/D > 5 fall within the range between the equations from Gazetas [18] and the equations from Shadlou et al. [20] used for flexible piles.
Figure 9 illustrates a similar relationship between KLR/(EsD2) and both (Lp/D) and ln(Eeq/Es) in the 3-D context. In Figure 9b, it is evident that the values of KLR/(EsD2) decrease with increasing Lp/D, and the values from the present study at Lp/D = 2 align well with the results of Shadlou et al. [20] for rigid pile conditions, as presented in Table 1 and Table 2. Furthermore, KLR/(EsD2) from the present study becomes insensitive to Lp/D as Lp/D increases. In Figure 9c, the results from the present study, where Lp/D ≥ 6, better align with the equations from Shadlou et al. [20] for flexible piles, indicating that KLR/(EsD2) exhibits behavior characteristic of flexible piles in this regime. A similar phenomenon has been observed in Figure 10, where KR/(EsD3) versus (Lp/D, ln(Eeq/Es)) in the 3-D context were analyzed, further emphasizing the importance of considering these parameters for accurate predictions in laterally loaded foundation analysis.
The observations made in Figure 8, Figure 9 and Figure 10 highlight the limitations of traditional methods in covering all calculating conditions for foundations with D ≥ 2 m and 2 ≤ Lp/D ≤ 10 in heterogeneous soil conditions. It becomes evident that the functions of KL/(EsD), KLR/(EsD2), and KR/(EsD3) should indeed be expressed as functions of Eeq/Es and Lp/D in heterogeneous soil conditions, necessitating the development of more accurate equations that account for both of these parameters. Moreover, the threshold zones for long flexible piles and short rigid piles in heterogeneous soil conditions differ from those in inhomogeneous and linearly inhomogeneous soils, underlining the importance of considering specific soil characteristics when analyzing foundation behavior.
Heterogeneous soil conditions (α = 0.25 and 0.75). Additionally, a similar phenomenon observed in KL/(EsD), KLR/(EsD2), and KR/(EsD3) versus (Lp/D, ln(Eeq/Es)) when α = 0.25 and 0.75 in the 3-D context further emphasizes the need for more accurate equations that can account for a wider range of scenarios without relying on the critical length parameter. This suggests the importance of developing equations that consider both the relative stiffness of the pile and soil and the slenderness ratio in future studies.

3.2. Parameter Analysis

The effect of materials and thickness. Figure 11 and Figure 12 provide valuable insights into the behavior of foundations in heterogeneous soil conditions with varying materials. In Figure 11, which considers different materials (steel piles with t = 0.01D and concrete piles with 0.1D) and α = 0.75, the results for KL/(EsD) show that the difference between these materials is within 3%. This suggests that using the function Eeq/Es is an effective way to account for different materials of hollow piles when analyzing the head stiffness of foundations in heterogeneous soil conditions. Similarly, in Figure 12, which focuses on concrete piles with varying wall thicknesses (t = 0.1D and 0.15D) and α = 0.25, vs = 0.30, the results for KL/(EsD) also indicate that the difference between these material variations is within 3%. This further supports the notion that the function Eeq/Es is a robust approach for considering material variations in the analysis of foundation behavior in heterogeneous soil conditions. These findings suggest that, compared with Poulos and Davis [14] and Aissa et al. [21], the function Eeq/Es can efficiently incorporate different materials in the analysis, simplifying the modeling process while maintaining accuracy in predicting the behavior of foundations with varying materials.
The effect of Poisson’s ratio. The consideration of Poisson’s ratio in the analysis of KL/(EsD), KLR/(EsD2), and KR/(EsD3) is relatively important. In previous research, the effect of Poisson’s ratio on KL/(EsD), KLR/(EsD2), and KR/(EsD3) is commonly neglected in some studies. Randolph [13] considered the effect by f(vs) = (1 + 0.75 vs)/(1+ vs); Shadlou and Bhattacharya [20] gave the function f ( ν s ) = 1 + ν s 0.25 ; Jabli et al. [52] presented solutions for stiffness of suction caissons having rigid skirted for 0.5 < L/D < 2 in three types of ground profiles with f ν s = 1 + 0.6 ν s 0.25 and by f(vs) = 1.1(0.096Lp/D + 0.6)vs2 − 0.7vs + 1.06. In this paper, the effect of Poisson’s ratio is observed, and it’s noted that the effect is influenced by various factors, including ln(Eeq/Es), Lp/D, and α. Comparing the values obtained for different Poisson’s ratios (vs = 0.2 to vs = 0.45) with those for vs = 0.30 under the same pile and soil conditions, it’s observed that the error ranges from −2% to 14%. This suggests that accounting for varying Poisson’s ratios in soils is essential for accurate analysis. Compared with Randolph [13] and Jabli et al. [52], a more efficient way is proposed to take into account different Poisson’s ratios in soils, and the following equation can be used:
f 1 ν s = f 2 ν s = 0.7146 α + 2.837 ν s 2 0.2666 α + 1.4381 ν s + 1.17
f 3 ν s = 1 + 0.4 ν s 0.3
The proposed equation for considering different Poisson’s ratios (vs) in soils has limited the errors in KL/(EsD), KLR/(EsD2), and KR/(EsD3) to a range of −1.5% to 5% across all cases in Table 3. Figure 13 gives an example of the modification effect of Equations (12) and (13) for KL/(EsD) in different conditions (α = 0 to 1, concrete and steel material), where the values of KL at vs = 0.2 to 0.45 were divided by the value of KL when vs = 0.3. results show that the values are in the range of 0.985 to 1.035, and the average values are nearly equal to 1.0. This suggests that the equation effectively accounts for variations in Poisson’s ratio and provides accurate predictions for the behavior of foundations under different soil conditions.

3.3. The Equations of Stiffness of Pile-Head Springs

From Section 3.1, the dimensionless function of the stiffness of pile-head springs, g(n), can be described by a serial of polynomial equations containing ln(Eeq/Es0) and Lp/D:
g n ln E e q E s 0 , L p D = i = 0 i = 3 j = 0 j = 4 P i j ( n ) ln E e q E s 0 i L p D j
where g(n) is the polynomial function related to KL/Es0D, KLR/Es0D2 and KR/Es0D3 at vs = 0.3, respectively; n = 1, 2, 3, g(1) is the constant of KL, g(2) is the constant of KLR, g(3) is the constant of KR; Pij is the dimensionless constant (i = 0, 1, 2, 3; j = 0, 1, 2, 3, 4); the value of P(1)ij is shown in Table 4, the value of P(2)ij is shown in Table 5, the value of P(3)ij is shown in Table 6.
From Section 3.2, the effect of Poisson’s ratio is given by Equations (12) and (13). Put Equations (12) and (13) into Equation (14), the semi-experimental equations of KL, KLR, and KR considering the effects of Poisson’s ratio are given as:
K L E s 0 D = g 1 f 1 ν s
K L R E s 0 D 2 = g 2 f 2 ν s
K R E s 0 D 3 = g 3 f 3 ν s
The polynomial equations proposed in Equations (15)–(17) exhibit a strong fit with the outcomes obtained through the energy-based variational method in Table 3, with a coefficient of determination (R2) exceeding 0.997.

4. Application of the Methodology

4.1. The Lateral Deflection and Rotation of Caissons for Serviceability Limit State Calculations

In the design process, caissons cannot be decidedly classified as slender or rigid beams. By using Equations (15)–(17), the head stiffness of the caissons was given, and the values can be put into Equation (18) to predict the lateral deflection and rotation at the caisson’s head. The results are given in Table 7.
w = K R K L K R K L R 2 F a K L R K L K R K L R 2 M a θ = K L R K L K R K L R 2 F a + K L K L K R K L R 2 M a
Table 7 and Figure 14 present the head stiffness of the caissons obtained from Equations (15)–(17) and Table 4, Table 5 and Table 6. Results show that these values differ significantly from those provided by Shadlou et al. [20], who categorized caissons as either rigid or flexible beams. This suggests that the proposed method in this paper takes into account the caissons’ behavior more accurately by not categorizing them as strictly rigid or flexible. Table 8 outlines the parameters for comparing lateral deflection and rotation between the method in this paper, Shadlou [20], and FEM analysis performed by ABAQUS [53]. The relative parameters are presented in Table 8, where different Ma/Fa/Lp were covered. In the FEM, half models were built, both the method in this paper and the FEM analysis use consistent characteristics of the caisson and soil properties, grid size, boundary conditions, and method of force application, which is essential for a meaningful comparison. However, the method in this paper employs the Timoshenko beam to construct the caisson, whereas FEM analysis employs the solid element.
Figure 15a illustrates a comparison of lateral deflection errors between this study, FEM, and Shadlou et al. [20] for different values of Lp/D, specifically Lp/D = 2.4 and Lp/D = 6. The errors in lateral deflection range from 8% to 12% for Lp/D = 2.4 and from −2% to 6% for Lp/D = 6 when compared to FEA results. In contrast, the errors from Shadlou et al. [20] and an unspecified FEM show significantly higher deviations, with errors ranging from −26% to −21% for Lp/D = 2.4 and −54% to 36% for Lp/D = 6. Similarly, Figure 15b presents a comparison of the rotation errors obtained from the head stiffness equations (Equations (15)–(17)) in this study with FEM results and those from Shadlou et al. [20]. The rotation errors from this study are found to be more consistent with FEM results when compared to those from Shadlou et al. [20].
Therefore, in the case of a semi-rigid caisson, the prediction of head deflection and rotation under static loading conditions is significantly contingent on the accurate determination of its head stiffness. Hence, there is a clear imperative to develop a more precise method for forecasting head stiffness.

4.2. The Natural Frequency of OWT Considering the SSI Effect

Three cases from Belwind [54], Walney [36] and Kentish Flats Offshore Wind Farm [54] were adopted to verify the accuracy of Equations (15)–(17) in predicting the natural frequencies of Offshore Wind Turbine (OWT) models while considering the Soil-Structure Interaction (SSI) effect. The relative parameters of the OWT system are presented in Figure 16 and detailed in Table 9. For the modification of the OWT system, FEM was conducted using ABAQUS [53], and a concentrated mass block was placed on top of a series of Timoshenko beams. To account for the monopile-soil interaction, three types of springs with initial stiffness values of KL, KLR, and KR, which were calculated using Equations (15)–(17), were used to account for the SSI effect.
From Table 9, it is observed that in the second case, where the critical length of the pile aligns more closely with a slender pile (as indicated in Table 1), the results of KL, KLR, and KR closely match those from the reference [54]. Conversely, in the first and the third cases, where the classification of the pile as slender or rigid is less clear based on Table 1 and Table 2, the results from the first case are lower than those expected for a typical rigid pile according to the reference [54], on the other hand, the results from the third case are higher than those expected for a typical flexible pile according to the reference [54].
The proposed method in this study has also been verified to enhance the accuracy of natural frequency predictions for OWT models considering the SSI effect. The natural frequencies of wind turbine structures, after accounting for the SSI effect, are provided in Table 10. Among the studied cases, the first natural frequency, as evaluated by both the references and this study, closely aligns with field testing results in the rotor-stop condition. However, discrepancies become more apparent in higher natural frequencies, with the largest error reaching 13%. This underscores the need for a more accurate equation for calculating dimensionless head stiffness, particularly when dealing with high-order self-resonant frequency calculations.

5. Conclusions

This study focuses on developing a precise method for calculating the head static stiffness of semi-rigid piles, without the need to consider the critical length, a parameter commonly used in traditional approaches. To achieve this objective, an extensive analysis comprising nearly twenty thousand cases was conducted using the energy-based variational method with Timoshenko beam theory. These cases encompassed a wide range of scenarios, including steel pipe piles and concrete caissons with diameters ranging from 2 m to 10 m, slenderness ratios between 2 and 10, and various soil types represented by the parameter α (ranging from 0 to 1) with a wide range of elastic moduli (Es) from 2 Mpa to 300 Mpa and Poisson’s ratios (vs) spanning 0.2 to 0.45.
The results demonstrate a limitation of traditional methods, as they fail to cover the full spectrum of foundation conditions within the aforementioned range. In contrast, the polynomial equations proposed in this paper exhibit a strong fit with the outcomes obtained through the energy-based variational method, with a coefficient of determination (R2) exceeding 0.997.
In practical applications, the method introduced in this study offers a distinct advantage. It enables engineers to achieve accurate results for head deflection and rotation of foundations under static loading conditions without the need for FDM or FEM software. This research also contributes to a more precise and accessible method for calculating foundation behavior, with potential applications in the analysis of high-order self-resonant frequencies in Offshore Wind Turbine (OWT) systems.
Notably, while these equations are applicable to a wide range of soil types and foundation scenarios, caution should be exercised when applying them to unique or unconventional cases without the aforementioned range, such as monopiles embedded in rocks.

Author Contributions

Conceptualization, W.L. and X.L.; methodology, X.L.; software, Q.Y.; validation, X.L., W.L. and M.Z.; formal analysis, M.Z.; investigation, X.L.; resources, X.L.; data curation, T.W.; writing—original draft preparation, W.L.; writing—review and editing, X.L.; visualization, T.W.; supervision, Q.Y.; project administration, X.L.; funding acquisition, X.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China [52201324] and the Natural Science Foundation of the Jiangsu Higher Education Institution of China [22KJB560015].

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Acknowledgments

The authors would like to thank the National Natural Science Foundation of China, and the Natural Science Foundation of the Jiangsu Higher Education Institution of China. The authors are grateful for their support.

Conflicts of Interest

The authors declare no conflict of interest.

Notation

The following symbols are used in this paper:
Dthe outer diameter of caisson
dw/dzthe rotation of the beam section
Eeqthe equivalent Young’s modulus of the beam
Epthe Young’s modulus of the beam
Esithe elastic modulus of the ith layer of soil
Es0the initial value of soil elastic modulus
Fathe lateral force at the head of beam
Ipthe second moment of inertia of cross-section
KLlateral stiffness
KRrocking stiffness
KLRcross-coupling stiffness
Lpthe embedment depth of beam
mRNAmass of rotor-nacelle assembly (kg)
Mathe moment at the head of beam
R2a coefficient of determination
rpthe radius of the beam
tthe wall thickness of beam
uzthe vertical displacement
wthe lateral displacement of the beam central line
αthe index of the function
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 soil domain
εpqthe strain in soil domain
λsi, Gsithe Lame’s constants of the ith layer of the multilayered continuum
vsithe Poisson’s ratio of the ith layer of soil
ϕrdimensionless decay functions of the displacement components in the r-directions
ϕθdimensionless decay functions of the displacement components in the θ-directions
ϕzdimensionless decay functions of the displacement components in the z-directions

Appendix A

For the beam at zLp, the governing differential equations of w(z) and ϕ(z) are coupled, and these equations are given by:
κ G p A p d 2 w d z 2 d ϕ d z + 2 t d 2 w d z 2 + k 1 d ϕ d z k w = 0
and
E p I p d 2 ϕ d z 2 + κ G p A p ( d w d z ϕ ) k 3 ϕ k 1 d w d z + 2 t 4 d 2 ϕ d z 2 = 0
where π 2 r p G s ( ϕ r 2 + ϕ θ 2 ) r d r d r = t , 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 1 , 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 , π r p G s ϕ z r ϕ θ + d ϕ z d r ϕ r r p r d r = k 2 , 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 = k 3 , π 2 r p ( λ s + 2 G s ) ϕ z 2 r p 2 r d r = t 4 .

Appendix B

The boundary conditions at the end of the foundation are calculated by:
At the pile head:
(1) Fa = constant, θ0 = 0
κ G p A p d w d z ϕ + k 2 ϕ + 2 t d w d z = F a d w d z = 0
where 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 Φ 1 + κ G p A p + k 2 E p I p + 2 t 4 k κ G p A p + k 3 κ G p A p k 1 + κ G p A p + k 2 κ G p A p k 1 κ G p A p + k 3 + κ G p A p + 2 t Φ 1 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 Φ 2 + κ G p A p + k 2 E p I p + 2 t 4 k κ G p A p + k 3 κ G p A p k 1 + κ G p A p + k 2 κ G p A p k 1 κ G p A p + k 3 + κ G p A p + 2 t Φ 2 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 Φ 3 + κ G p A p + k 2 E p I p + 2 t 4 k κ G p A p + k 3 κ G p A p k 1 + κ G p A p + k 2 κ G p A p k 1 κ G p A p + k 3 + κ G p A p + 2 t Φ 3 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 Φ 4 + κ G p A p + k 2 E p I p + 2 t 4 k κ G p A p + k 3 κ G p A p k 1 + κ G p A p + k 2 κ G p A p k 1 κ G p A p + k 3 + κ G p A p + 2 t Φ 4 z = 0 T C 1 C 2 C 3 C 4 z = 0 ( 1 ) = F a and Φ 1 Φ 2 Φ 3 Φ 4 T z = 0 C 1 C 2 C 3 C 4 z = 0 ( 1 ) = 0 .
(2) w0 = constant, θ0 = 0
w 0 = 1 d w d z = 0
where Φ 1 Φ 2 Φ 3 Φ 4 z = 0 T C 1 C 2 C 3 C 4 z = 0 ( 1 ) = 1 and Φ 1 Φ 1 Φ 3 Φ 4 T z = 0 C 1 C 2 C 3 C 4 z = 0 ( 1 ) = 0 .
(3) Ma = constant, w0 = 0
w 0 = 0 E p I p d ϕ d z + k 1 + k 2 w + 2 t 4 d ϕ d z = M a
where Φ 1 Φ 2 Φ 3 Φ 4 T z = 0 C 1 C 2 C 3 C 4 z = 0 ( 1 ) = 0 and E p I p + 2 t 4 κ G p A p k 1 κ G p A p + 2 t Φ 1 + k E p I p + 2 t 4 κ G p A p k 1 + k 1 + k 2 Φ 1 E p I p + 2 t 4 κ G p A p k 1 κ G p A p + 2 t Φ 2 + k E p I p + 2 t 4 κ G p A p k 1 + k 1 + k 2 Φ 2 E p I p + 2 t 4 κ G p A p k 1 κ G p A p P + 2 t Φ 3 + k E p I p + 2 t 4 κ G p A p k 1 + k 1 + k 2 Φ 3 E p I p + 2 t 4 κ G p A p k 1 κ G p A p P + 2 t Φ 4 + k E p I p + 2 t 4 κ G p A p k 1 + k 1 + k 2 Φ 4 z = 0 T C 1 C 2 C 3 C 4 z = 0 ( 1 ) = M a .
The boundary conditions for the ith layer (i = 1, 2, 3, …, n) of the beam at z = Hi < Lp are calculated by:
w i , z = H i = w i + 1 , z = H i d w d z i , z = H i = d w d z i + 1 , z = H i κ G p A p d w d z ϕ + k 2 ϕ + 2 t d w d z i , z = H i = κ G p A p d w d z ϕ + k 2 ϕ + 2 t d w d z i + 1 , z = H i E p I p d ϕ d z + k 1 + k 2 w + 2 t 4 d ϕ d z i , z = H i = E p I p d ϕ d z + k 1 + k 2 w + 2 t 4 d ϕ d z i + 1 , z = H i
where Φ 1 Φ 2 Φ 3 Φ 4 Φ 1 Φ 2 Φ 3 Φ 4 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 Φ 1 + E p I p + 2 t 4 k κ G p A p + k 3 κ G p A p k 1 + κ G p A p k 1 κ G p A p + k 3 + κ G p A p + 2 t κ G p A p + k 2 Φ 1 . . . E p I p + 2 t 4 κ G p A p k 1 κ G p A p + 2 t Φ 1 + E p I p + 2 t 4 κ G p A p k 1 k + k 1 + k 2 Φ 1 . . . z = z i ( i ) C 1 C 2 C 3 C 4 z = z i ( i ) = Φ 1 Φ 2 Φ 3 Φ 4 Φ 1 Φ 2 Φ 3 Φ 4 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 Φ 1 + E p I p + 2 t 4 k κ G p A p + k 3 κ G p A p k 1 + κ G p A p k 1 κ G p A p + k 3 + κ G p A p + 2 t κ G p A p + k 2 Φ 1 . . . E p I p + 2 t 4 κ G p A p k 1 κ G p A p + 2 t Φ 1 + E p I p + 2 t 4 κ G p A p k 1 k + k 1 + k 2 Φ 1 . . . z = z i ( i + 1 ) C 1 C 2 C 3 C 4 z = z i ( i + 1 ) .
At the pile end:
Fixed end:
w z = L p = 0 d w d z z = L p = 0
where Φ 1 Φ 2 Φ 3 Φ 4 z = 0 T C 1 C 2 C 3 C 4 z = L p ( n ) = 0 , and Φ 1 Φ 2 Φ 3 Φ 4 z = 0 T C 1 C 2 C 3 C 4 z = L p ( n ) = 0 .
Free end:
κ G p A p d w d z ϕ + k 2 ϕ + 2 t d w d z n , z = L p = k 2 e n d ϕ + 2 t e n d d w d z n + 1 , z = L p E p I p d ϕ d z + k 1 + k 2 w + 2 t 4 d ϕ d z n , z = L p = k 1 e n d + k 2 e n d w + 2 t 4 e n d d ϕ d z n + 1 , z = L p
where 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 e n d , 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 G s r p 2 = k 1 e n d , π r p G s ϕ z r ϕ θ + d ϕ z d r ϕ r r p r d r + π G s r p 2 = k 2 e n d , 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 G s r p 2 = k 3 e n d , π 2 r p ( λ s + 2 G s ) ϕ z 2 r p 2 r d r + ( λ s + 2 G s ) π 8 r p 4 = t 4 e n d , and η 3 Φ 1 + η 4 Φ 1 2 t e n d η 1 Φ 1 n , z = L p η 3 Φ 2 + η 4 Φ 2 2 t e n d η 1 Φ 2 n , z = L p η 3 Φ 3 + η 4 Φ 3 2 t e n d η 1 Φ 3 n , z = L p η 3 Φ 4 + η 4 Φ 4 2 t e n d η 1 Φ 4 n , z = L p T C 1 C 2 C 3 C 4 z = L p n = 0 , η 7 Φ 1 + η 8 Φ 4 + η 9 Φ 1 + η 10 Φ 1 η 7 Φ 2 + η 8 Φ 4 + η 9 Φ 2 + η 10 Φ 2 η 7 Φ 3 + η 8 Φ 3 + η 9 Φ 3 + η 10 Φ 3 η 7 Φ 4 + η 8 Φ 4 + η 9 Φ 4 + η 10 Φ 4 T C 1 C 2 C 3 C 4 z = L p n = 0 , η 1 = a + b a 2 t 4 e n d k 3 e n d 2 t e n d k 1 e n d a 3 k 1 e n d k 3 e n d + 2 t 4 e n d k 3 e n d k k 1 e n d a 2 t 4 e n d k 3 e n d 2 t e n d k 1 e n d b 3 a 3 k 1 e n d k 3 e n d + 2 t 4 e n d k 3 e n d k k 1 , e n d b a , η 2 = b a 1 2 t 4 e n d k 3 e n d 2 t e n d k 1 e n d b 3 a 3 k 1 e n d k 3 e n d + 2 t 4 e n d k 3 e n d k k 1 e n d b a , η 3 = κ G p A p + k 2 k 2 e n d 2 t e n d η 2 E p I p + 2 t 4 κ G p A p + 2 t κ G p A p + k 3 κ G p A p k 1 , η 4 = κ G p A p + k 2 k 2 e n d 2 t e n d η 2 E p I p + 2 t 4 k κ G p A p + k 3 κ G p A p k 1 + κ G p A p k 1 κ G p A p + k 3 + κ G p A p + 2 t , η 5 = a 2 a 2 + b 2 2 t 4 e n d k 3 e n d 2 t e n d k 1 e n d a 3 k 1 e n d k 3 e n d + 2 t 4 e n d k 3 e n d k k 1 e n d a 2 t 4 e n d k 3 e n d 2 t e n d k 1 e n d b 3 a 3 k 1 e n d k 3 e n d + 2 t 4 e n d k 3 e n d k k 1 e n d b a , η 6 = a 2 + b 2 2 t 4 e n d k 3 e n d 2 t e n d k 1 e n d b 3 a 3 k 1 e n d k 3 e n d + 2 t 4 e n d k 3 e n d k k 1 e n d b a , η 7 = 2 t e n d 2 t 4 e n d k 1 , e n d η 6 E p I p + 2 t 4 κ G p A p + 2 t κ G p A p + k 3 κ G p A p k 1 , 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 e n d 2 t 4 e n d k 1 , e n d η 6 E p I p + 2 t 4 k κ G p A p + k 3 κ G p A p k 1 + κ G p A p k 1 κ G p A p + k 3 , η 10 = E p I p + 2 t 4 κ G p A p k 1 k + k 1 + k 2 k 1 e n d + k 2 e n d + 2 t e n d 2 t 4 e n d k 1 , e n d η 5 2 k t 4 e n d k 1 , e n d , a = t ˜ + t ˜ 2 k ˜ , b = t ˜ t ˜ 2 k ˜ , t ˜ = k 3 , e n d 2 t e n d + 2 t 4 , e n d k k 1 , e n d 2 8 t 4 , e n d t e n d , k ˜ = k 3 , e n d k 2 t 4 , e n d 2 t e n d .
For the soil area at z > Lp, the governing differential equations are given by:
2 t e n d d 2 w d z 2 + k 1 e n d d ϕ d z k w = 0 k 3 e n d ϕ k 1 e n d d w d z + 2 t 4 e n d d 2 ϕ d z 2 = 0
where w z = L p = w z = L p + , and ϕ z = L p = ϕ z = L p + , therefore, the solution of Equation (11) is given as:
w = C 1 e a z + C 2 e b z
where C 1 = w z = L p ϕ z = L p 2 t 4 , e n d k 3 , e n d 2 t e n d k 1 , e n d a 3 k 1 , e n d k 3 , e n d + 2 t 4 , e n d k 3 , e n d k k 1 , e n d a w z = L p 2 t 4 , e n d k 3 , e n d 2 t e n d k 1 , e n d b 3 a 3 k 1 , e n d k 3 , e n d + 2 t 4 , e n d k 3 , e n d k k 1 , e n d b a e a L p , C 2 = ϕ z = L p 2 t 4 , e n d k 3 , e n d 2 t e n d k 1 , e n d a 3 k 1 , e n d k 3 , e n d + 2 t 4 , e n d k 3 , e n d k k 1 , e n d a w z = L p 2 t 4 , e n d k 3 , e n d 2 t e n d k 1 , e n d b 3 a 3 k 1 , e n d k 3 , e n d + 2 t 4 , e n d k 3 , e n d k k 1 , e n d b a e b L p .

Appendix C

As mentioned previously, the governing differential equations for the soil surrounding the deep foundation can be obtained. The governing differential equations for the functions ϕr are:
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 d ϕ d 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 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 .
The governing differential equations for the functions ϕθ are:
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 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 ϕ r p d w d z λ s d ϕ d z r p w d z 0 G s w w d z .
The governing differential equations for the functions ϕz are:
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 G s ϕ d w d z λ s d ϕ d r w d z 0 G s ϕ r p ϕ d z , γ 9 = 0 ( λ s + 2 G s ) d ϕ d r d ϕ d r d z 0 G s ϕ ϕ d z .
The relationship among ϕr, ϕθ, and ϕz can be described as follows:
ϕ r = K ϕ r CK ϕ z 1 M ( B + CK ϕ z 1 N ) ( K ϕ θ FK ϕ z 1 N ) 1 ( E + FK ϕ z 1 M ) ] 1 A + CK ϕ z 1 G + ( B + CK ϕ z 1 N ) ( K ϕ θ FK ϕ z 1 N ) 1 ( D + FK ϕ z 1 G ) ϕ θ = ( K ϕ θ FK ϕ z 1 N ) 1 D + FK ϕ z 1 G + ( E + FK ϕ z 1 M ) ϕ r ϕ z = K ϕ z 1 G + K ϕ z 1 M ϕ r + K ϕ z 1 N ϕ θ
where the matrixes of Kϕr, Kϕθ, Kϕz, Am, Bm, Cm, P, E, Q, G, M, and N are given as following: Buildings 14 01803 i001Buildings 14 01803 i002
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 0 0 0 . . . . . 0 0 0 0 0 . . . . 0 0 0 γ 8 2 r 0 . . . . . . . γ 8 2 r . . . . . . . . . . . . . . . . . 0 γ 8 2 r 0 0 . . . . . 0 γ 8 2 r 0 0 . . 0 0 0 0 0 0 4 n × 4 n
The above calculation is conducted in MATLAB (Figure A1).
Figure A1. Flow chart of the iterative algorithm in MATLAB [55].
Figure A1. Flow chart of the iterative algorithm in MATLAB [55].
Buildings 14 01803 g0a1

References

  1. 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. Comput. Geotech. 2023, 153, 105085. [Google Scholar] [CrossRef]
  2. 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]
  3. Qin, W.; Gao, J.; Chang, K.; Dai, G.; Wei, H. Set-up effect of large-diameter open-ended thin-walled pipe piles driven in clay. Comput. Geotech. 2023, 159, 105459. [Google Scholar] [CrossRef]
  4. Wu, J.-S.; Hsu, S.-H. A unified approach for the free vibration analysis of an elastically supported immersed uniform beam carrying an eccentric tip mass with rotary inertia. J. Sound Vib. 2006, 291, 1122–1147. [Google Scholar] [CrossRef]
  5. Galvín, P.; Romero, A.; Solís, M.; Domínguez, J. Dynamic characterisation of wind turbine towers account for a monopile foundation and different soil conditions. Struct. Infrastruct. Eng. 2016, 13, 942–954. [Google Scholar] [CrossRef]
  6. Wang, P.; Zhao, M.; Du, X.; Liu, J.; Xu, C. Wind, wave and earthquake responses of offshore wind turbine on monopile foundation in clay. Soil Dyn. Earthq. Eng. 2018, 113, 47–57. [Google Scholar] [CrossRef]
  7. Arany, L.; Bhattacharya, S.; John, M.; Hogan, S.J. Design of monopiles for offshore wind turbines in 10 steps. Soil Dyn. Earthq. Eng. 2017, 92, 126–152. [Google Scholar] [CrossRef]
  8. Subhamoy, B. Design of Foundations for Offshore Wind Turbines; John Wiley & Sons Ltd.: West Sussex, UK, 2017. Available online: https://lccn.loc.gov/2018046931 (accessed on 16 February 2024).
  9. Hu, Q.; Han, F.; Prezzi, M.; Salgado, R.; Zhao, M. Lateral load response of large-diameter monopiles in sand. Géotechnique 2021, 72, 1035–1050. [Google Scholar] [CrossRef]
  10. Hu, Q.; Han, F.; Prezzi, M.; Salgado, R.; Zhao, M. Finite-element analysis of the lateral load response of monopiles in layered sand. J. Geotech. Geoenviron. Eng. 2022, 148, 04022001. [Google Scholar] [CrossRef]
  11. Bergfelt, A. The axial and lateral loading bearing capacity, and failure by buckling of piles in soft clay. In Proceedings of the 4th International Conference on Soil Mechanics and Foundation Engineering, London, UK, 12–14 August 1957; Volume 2, pp. 8–19. [Google Scholar]
  12. Thomas, G. Determination of Spring Constant, Term Project; Rensselaer Polytechnic Institute: Troy, NY, USA, 1978. [Google Scholar]
  13. Randolph, M. The response of flexible piles to lateral loading. Geotechnique 1981, 2, 247–259. [Google Scholar] [CrossRef]
  14. Poulos, H.; Davis, E. Pile Foundation Analysis and Design; Rainbow-Bridge Book Co.: Landon, UK, 1980. [Google Scholar]
  15. Barber, E.S. Discussion to paper by S. M. Gleser. ASTM Spec. Tech. Publ. 1953, 154, 96–99. [Google Scholar]
  16. Budhu, M.; Davies, G. Analysis of Laterally Loaded Piles in Soft Clays. J. Geotech. Eng. 1988, 114, 21. [Google Scholar] [CrossRef]
  17. Pender, M.J. ASeismic Pile Foundation Design Analysis. Bull. N. Z. Natl. Soc. Earthq. Eng. 1993, 26, 49–161. [Google Scholar] [CrossRef]
  18. Gazetas, G. Seismic response of end-bearing single piles. Int. J. Soil Dyn. Earthq. Eng. 1984, 3, 82–93. [Google Scholar] [CrossRef]
  19. Higgins, W.; Vasquez, C.; Basu, D.; Griffiths, D.V. Elastic solutions for laterally loaded piles. J. Geotech. Geoenviron. Eng. 2013, 139, 1096–1103. [Google Scholar] [CrossRef]
  20. Shadlou, M.; Bhattacharya, S. Dynamic stiffness of monopiles supporting offshore wind turbine generators. Soil Dyn. Earthq. Eng. 2016, 8, 15–32. [Google Scholar] [CrossRef]
  21. Aissa, M.; Bouzid, D.A.; Bhattacharya, S. Monopile head stiffness for serviceability limit state calculations in assessing the natural frequency of offshore wind turbines. Int. J. Geotech. Eng. 2017, 6, 267–283. [Google Scholar] [CrossRef]
  22. Anoyatis, G.; Lemnitzer, A. Dynamic pile impedances for laterally-loaded piles using improved Tajimi and Winkler formulations. Soil Dyn. Earthq. Eng. 2017, 92, 279–297. [Google Scholar] [CrossRef]
  23. Rao, S.S. Vibration of Continuous Systems; John Wiley & Sons, Inc.: Hoboken, NJ, USA, 2007. [Google Scholar]
  24. Ko, Y.; Li, Y. Response of a scale-model pile group for a jacket foundation of an offshore wind turbine in liquefiable ground during shaking table tests. Earthq. Eng. Struct. Dyn. 2020, 49, 1682–1701. [Google Scholar] [CrossRef]
  25. European Committee for Standardization. Eurocode 8: Design of Structures for Earthquake Resistance—Part 5: Foundations, Retaining Structures and Geotechnical Aspects; European Committee for Standardization: Brussels, Belgium, 2003. [Google Scholar]
  26. Syngros, C. Seismic Response of Piles and pile-Supported Bridge Piers Evaluated through Case Histories. Ph.D. Thesis, City University of New York, New York, NY, USA, 2004. [Google Scholar]
  27. Carter, J.; Kulhawy, F. Analysis of laterally loaded shafts in rock. J. Geotech. Eng. 1992, 118, 839–855. [Google Scholar] [CrossRef]
  28. Bouzid, D.A.; Bhattacharya, S.; Otsmane, 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]
  29. Abed, Y.; Bouzid, D.A.; Bhattacharya, S. Static impedance functions for monopiles supporting offshore wind turbines in nonhomogeneous soils emphasis on soil/monopile interface characteristics. Earthq. Struct. 2016, 10, 1143–1179. [Google Scholar] [CrossRef]
  30. Salgado, R.; Tehrani, F.S.; Prezzi, M. Analysis of laterally loaded pile groups in multilayered elastic soil. Comput. Geotech. 2014, 62, 136–153. [Google Scholar] [CrossRef]
  31. Han, F. Axial and Lateral Resistance of Non-Displacement Piles. Ph.D. Thesis, Purdue University, West Lafayette, IN, USA, 2017. [Google Scholar]
  32. Han, F.; Prezzi, M.; Salgado, R. Energy-based solutions for non-displacement piles subjected to lateral loads. Int. J. Geomech. 2017, 17, 04017104. [Google Scholar] [CrossRef]
  33. Byrne, B.W.; Burd, H.J.; Houlsby, G.T. New Design Methods for Large Diameter Piles under Lateral Loading for Offshore Wind Applications. In Proceedings of the Third International Symposium on Frontiers in Offshore Geotechnics ISFOG, Oslo, Norway, 10–12 June 2015. [Google Scholar]
  34. Byrne, B.W.; Mcadam, R.A.; Burd, H.J. Field testing of large diameter piles under lateral loading for offshore wind applications. In Proceedings of the XVI European Conference on Soil Mechanics and Geotechnical Engineering (ECSMGE), Edinburg, UK, 13–17 September 2016; pp. 1255–1260. [Google Scholar]
  35. Gupta, B.K.; Basu, D. Applicability of Timoshenko, Euler–Bernoulli and rigid beam theories in the analysis of laterally loaded monopiles and piles. Géotechnique 2018, 9, 772–785. [Google Scholar] [CrossRef]
  36. Gupta, B.K.; Basu, D. Offshore wind turbine monopile foundations: Design perspectives. Ocean Eng. 2020, 213, 107514. [Google Scholar] [CrossRef]
  37. Gupta, B.K. Dynamic pile-head stiffness of laterally loaded end-bearing pile in linear viscoelastic soil—A comparative study. Comput. Geotech. 2022, 145, 104654. [Google Scholar] [CrossRef]
  38. Cao, G.; Hian, S.; Ding, X. Horizontal static for OWT monopiles based on Timoshenko beam theory. Int. J. Geomech. 2023, 23, 04023202. [Google Scholar] [CrossRef]
  39. 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]
  40. Li, X.; Dai, G. Closure to “Energy-Based Analysis of Laterally Loaded Caissons with Large Diameters under Small-Strain Conditions”. Int. J. Geomech. 2023, 23, 07023008. [Google Scholar] [CrossRef]
  41. Li, X.; Dai, G.; Zhu, M.; Zhu, W.; Zhang, F. Investigation of the soil deformation around late-rally loaded deep foundations with large diameters. Acta Geotech. 2023, 19, 2293–2314. [Google Scholar] [CrossRef]
  42. 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]
  43. 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]
  44. Bhattacharya, S.; Carrington, T.; Aldridge, T. Observed increases in offshore piledriving resistance. Proc. Inst. Civ. Eng.—Geotech. Eng. 2009, 162, 71–80. [Google Scholar] [CrossRef]
  45. Skempton, A.W.; Henkel, D.J. Tests on London Clay fromDeepBorings at Paddington, Victoria, and the South Bank. In Proceedings of the 4th International Conference on Soil Mechanics and Foundation Engineering, London, UK, 12–14 August 1957; Volume 1, pp. 100–106. [Google Scholar]
  46. Ward, W.H.; Marsland, A.; Samuels, S.G. Properties of the London Clay at the Ashford Common Shaft: In-Situ and Undrained Strength Tests. Geotechnique 1965, 15, 321–344. [Google Scholar] [CrossRef]
  47. Burland, J.B.; Lord, J.A. The Load-Deformation Behavior of the Middle Chalk at Mundford, Norfolk: A Comparison between Full Scale Performance and In-Situ Laboratory Measurements, Building Research Station CP. 1969. Available online: https://eurekamag.com/research/018/258/018258882.php (accessed on 16 February 2024).
  48. Butler, F.G. Heavily Over-Consolidated Clays. In BGS Conference on Settlement of Structures; Wiley: London, UK, 1974; pp. 531–578. Available online: http://www.geosolve.co.uk/refs/Butler_F_G_1975.pdf (accessed on 16 February 2024).
  49. Hobbs, N.B. Factors Affecting the Prediction of Settlement on Rock with Particular Reference to Chalk and Trias. In BGS Conference on Settlement of Structures; Wiley: London, UK, 1974; p. 579-560. [Google Scholar]
  50. Selvadurai, P. The Analytical Method in Geomechanics. Appl. Mech. Rev. 2007, 60, 87–106. [Google Scholar] [CrossRef]
  51. Malekjafarian, A.; Jalilvand, S.; Doherty, P.; Igoe, D. Foundation damping for monopile supported offshore wind turbines: A review. Mar. Struct. 2021, 77, 102937. Available online: https://www.sciencedirect.com/science/article/pii/S0951833921000058 (accessed on 16 February 2024). [CrossRef]
  52. 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]
  53. ABAQUS. Abaqus Analysis User’s Manual: ABAQUS 6.15; SIMULIA: Johnston, RI, USA, 2018. [Google Scholar]
  54. Laszlo, A.; Bhattacharya, S.; Macdonald, J.H.G.; 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]
  55. The MathWorks Inc. Optimization Toolbox Version: (R2018b); The MathWorks Inc.: Natick, MA, USA, 2018; Available online: https://www.mathworks.com (accessed on 16 February 2024).
Figure 1. The bending behavior of a semi-rigid beam, a flexible beam, and a rigid beam in offshore or bridge projects (free-end boundary condition).
Figure 1. The bending behavior of a semi-rigid beam, a flexible beam, and a rigid beam in offshore or bridge projects (free-end boundary condition).
Buildings 14 01803 g001
Figure 2. The results of KL/(EsD) at (a) 3-D, (b) Lp/D and (c) ln(Eeq/Es) version in homogenous soil conditions compared with different methods (t = 0.01D, vs = 0.30).
Figure 2. The results of KL/(EsD) at (a) 3-D, (b) Lp/D and (c) ln(Eeq/Es) version in homogenous soil conditions compared with different methods (t = 0.01D, vs = 0.30).
Buildings 14 01803 g002
Figure 3. The results of KLR/(EsD2) at (a) 3-D, (b) Lp/D and (c) ln(Eeq/Es) version in homogenous soil conditions compared with different methods (t = 0.01D, vs = 0.30).
Figure 3. The results of KLR/(EsD2) at (a) 3-D, (b) Lp/D and (c) ln(Eeq/Es) version in homogenous soil conditions compared with different methods (t = 0.01D, vs = 0.30).
Buildings 14 01803 g003
Figure 4. The results of KR/(EsD3) at (a) 3-D, (b) Lp/D and (c) ln(Eeq/Es) version in homogenous soil conditions compared with different methods (t = 0.01D, vs = 0.30).
Figure 4. The results of KR/(EsD3) at (a) 3-D, (b) Lp/D and (c) ln(Eeq/Es) version in homogenous soil conditions compared with different methods (t = 0.01D, vs = 0.30).
Buildings 14 01803 g004
Figure 5. The results of KL/(EsD) at (a) 3-D, (b) Lp/D and (c) ln(Eeq/Es) version in linearly inhomogeneous soil conditions compared with different methods (t = 0.01D, vs = 0.30).
Figure 5. The results of KL/(EsD) at (a) 3-D, (b) Lp/D and (c) ln(Eeq/Es) version in linearly inhomogeneous soil conditions compared with different methods (t = 0.01D, vs = 0.30).
Buildings 14 01803 g005aBuildings 14 01803 g005b
Figure 6. The results of KLR/(EsD2) at (a) 3-D, (b) Lp/D and (c) ln(Eeq/Es) version in linearly inhomogeneous soil conditions compared with different methods (t = 0.01D, vs = 0.30).
Figure 6. The results of KLR/(EsD2) at (a) 3-D, (b) Lp/D and (c) ln(Eeq/Es) version in linearly inhomogeneous soil conditions compared with different methods (t = 0.01D, vs = 0.30).
Buildings 14 01803 g006aBuildings 14 01803 g006b
Figure 7. The results of KR/(EsD3) at (a) 3-D, (b) Lp/D and (c) ln(Eeq/Es) version in linearly inhomogeneous soil conditions compared with different methods (t = 0.01D, vs = 0.30).
Figure 7. The results of KR/(EsD3) at (a) 3-D, (b) Lp/D and (c) ln(Eeq/Es) version in linearly inhomogeneous soil conditions compared with different methods (t = 0.01D, vs = 0.30).
Buildings 14 01803 g007aBuildings 14 01803 g007b
Figure 8. The results of KL/(EsD) at (a) 3-D, (b) Lp/D and (c) ln(Eeq/Es) version in homogenous soil conditions compared with different methods (t = 0.01D, vs = 0.25–0.35).
Figure 8. The results of KL/(EsD) at (a) 3-D, (b) Lp/D and (c) ln(Eeq/Es) version in homogenous soil conditions compared with different methods (t = 0.01D, vs = 0.25–0.35).
Buildings 14 01803 g008aBuildings 14 01803 g008b
Figure 9. The results of KLR/(EsD2) at (a) 3-D, (b) Lp/D and (c) ln(Eeq/Es) version in heterogeneous soil conditions compared with different methods (t = 0.01D, vs = 0.25-0.35).
Figure 9. The results of KLR/(EsD2) at (a) 3-D, (b) Lp/D and (c) ln(Eeq/Es) version in heterogeneous soil conditions compared with different methods (t = 0.01D, vs = 0.25-0.35).
Buildings 14 01803 g009
Figure 10. The results of KR/(EsD3) at (a) 3-D, (b) Lp/D and (c) ln(Eeq/Es) version in heterogeneous soil conditions compared with different methods (t = 0.01D, vs = 0.30).
Figure 10. The results of KR/(EsD3) at (a) 3-D, (b) Lp/D and (c) ln(Eeq/Es) version in heterogeneous soil conditions compared with different methods (t = 0.01D, vs = 0.30).
Buildings 14 01803 g010
Figure 11. KL/(Es0D) in heterogeneous soil conditions (α = 0.75, t = 0.1D, Ep = 30 Gpa and 210 Gpa, vs = 0.30).
Figure 11. KL/(Es0D) in heterogeneous soil conditions (α = 0.75, t = 0.1D, Ep = 30 Gpa and 210 Gpa, vs = 0.30).
Buildings 14 01803 g011
Figure 12. KL/(EsD) in heterogeneous soil conditions compared with different methods (α = 0.25, t = 0.05D and 0.15D, concrete material, vs = 0.30).
Figure 12. KL/(EsD) in heterogeneous soil conditions compared with different methods (α = 0.25, t = 0.05D and 0.15D, concrete material, vs = 0.30).
Buildings 14 01803 g012
Figure 13. KL,vs=0.2 to 0.45/KL,vs=0.3 in different conditions (α = 0 to 1, concrete and steel material vs = 0.2 to 0.45).
Figure 13. KL,vs=0.2 to 0.45/KL,vs=0.3 in different conditions (α = 0 to 1, concrete and steel material vs = 0.2 to 0.45).
Buildings 14 01803 g013
Figure 14. A laterally loaded caisson.
Figure 14. A laterally loaded caisson.
Buildings 14 01803 g014
Figure 15. The error of (a) the lateral deflection and (b) the rotation from different methods with Lp/D = 2.4 and Lp/D = 6 (wFEM is the head lateral deflection of caissons from ABAQUS, θFEM is the head rotation of caissons from ABAQUS).
Figure 15. The error of (a) the lateral deflection and (b) the rotation from different methods with Lp/D = 2.4 and Lp/D = 6 (wFEM is the head lateral deflection of caissons from ABAQUS, θFEM is the head rotation of caissons from ABAQUS).
Buildings 14 01803 g015
Figure 16. An OWT model considering the SSI effect.
Figure 16. An OWT model considering the SSI effect.
Buildings 14 01803 g016
Table 1. Different equations of stiffness formula by different researchers for Semi-infinite beams in various soil profile. m* is the equivalent ratio of shear modulus, m* = dGs/dz(1 + 0.75νs)
Table 1. Different equations of stiffness formula by different researchers for Semi-infinite beams in various soil profile. m* is the equivalent ratio of shear modulus, m* = dGs/dz(1 + 0.75νs)
SourceKL, KR and KLR (If Have)Constant ValueBeam and Soil Material
KLKLRKROthers
Randolph [13], Ko [24]KL:
a 1 E s 0 D f ν s E e q E s 0 b 1
KLR:
a 2 E s 0 D 2 f ν s E e q E s 0 b 2
KR:
a 3 E s 0 D 3 f ν s E e q E s 0 b 3
a1 = 3.147
b1 = 1/7
a2 = −0.5311
b2 = 3/7
a3 = 0.2458
b3 = 5/7
E e q = E p I p π D 4 64
f ν s = 1 + 0.75 ν s 1 + ν s
Es0 is the initial value of soil elastic modulus
The Semi-infinite beam in homogeneous and linear inhomogeneous soils (Triangular-linear strain element)
Pender [17] a1 = 1.285
b1 = 0.188
a2 = −0.3075
b2 = 0.47
a3 = 0.18125
b3 = 0.738
f(vs) = 1
m* is the equivalent ratio of shear modulus,
m* = dGs/dz(1 + 0.75νs)
The Semi-infinite beam in homogeneous soil
a1 = 0.85
b1 = 0.29
a2 = −0.24
b2 = 0.53
a3 = 0.15
b3 = 0.77
The Semi-infinite beam in linear inhomogeneous homogeneous soil
a1 = 0.735
b1 = 0.33
a2 = −0.27
b2 = 0.55
a3 = 0.1725
b3 = 0.776
The Semi-infinite beam in parabolic inhomogeneous soil
Gazetas [18] and Eurocode 8 Part 5 [25]a1 = 1.08
b1 = 0.21
a2 = −0.22
b2 = 0.50
a3 = 0.16
b3 = 0.75
The Semi-infinite beam in homogeneous soil
a1 = 0.60
b1 = 0.35
a2 = −0.17
b2 = 0.60
a3 = 0.14
b3 = 0.80
The Semi-infinite beam in linear inhomogeneous soil
a1 = 0.79
b1 = 0.28
a2 = −0.24
b2 = 0.53
a3 = 0.15
b3 = 0.77
The Semi-infinite beam in parabolic inhomogeneous soil
a1 = 0.1967 m*
b1 = 18
a2 = −0.3472 m*
b2 = 0.43
a3 = 0.2083 m*
b3 = 0.72
-
a1=0.1967 m*
b1 = 33
a2=−0.3472 m*
b2 = 0.54
a3=0.2083 m*
b3 = 0.78
-
Syngros [26];
Anoyatis [22]
a1 = 1.24
b1 = 0.18
a2 = −0.21
b2 = 0.50
a3 = 0.15
b3 = 0.75
The Semi-infinite beam in parabolic inhomogeneous soil
Shadlou et al. [20]a1 = 1.45
b1 = 0.186
a2 = −0.30
b2 = 0.50
a3 = 0.18
b3 = 0.73
f ν s = 1 1 + ν s 0.25 The Semi-infinite beam in homogeneous soil
a1 = 0.79
b1 = 0.34
a2 = −0.26
b2 = 0.567
a3 = 0.17
b3 = 0.78
The Semi-infinite beam in linear inhomogeneous homogeneous soil
a1 = 1.02
b1 = 0.27
a2 = −0.29
b2 = 0.52
a3 = 0.17
b3 = 0.76
The Semi-infinite beam in parabolic inhomogeneous soil
Higgins et al. [19] K L = 100 G s r p E p E s 0.18 3.4 E p E s 0.04 K L R = 1 0.3 G s r p 2 E p E s 0.47 E p E s 0.04 3.4 K R = 34 9 G s r p 3 E p E s 0.72 3.4 E p E s 0.04 f(νs) = Gs(1 + 0.75νs)/Es
m* = dGs/dz(1 + 0.75νs)
The Semi-infinite beam in homogeneous soil
K L = 1.23 m r p 2 E p m r p 0.33 0.6765 0.2704 E p m r p 0.03 K L R = 0.52 m r p 3 E p m r p 0.57 0.2704 E p m r p 0.03 0.6765 K R = 0.55 m r p 4 E p m r p 0.78 0.6765 0.2704 E p m r p 0.03 The Semi-infinite beam in linear inhomogeneous soil
a1 = 3\3.66\4.58
(Lp/rp = 40, Ep/Es = 100\300\1000)
a2 = 2.70\4.48\7.81
(Lp/rp = 6, Ep/Es = 100~1000
a3 = 6.02\13.20\31.31
(Lp/rp = 6, Ep/Es = 100~1000)
-
Table 2. Different equations of Stiffness formulae by different researchers for short beams in various soil profile.
Table 2. Different equations of Stiffness formulae by different researchers for short beams in various soil profile.
SourceKL, KR and KLR (If Have)Constant ValueOthersBeam and Soil Material
KLKLRR
Randolph [13], Ko [24]KL:
a 1 E s 0 D f ν s E e q E s 0 b 1
KLR:
a 2 E s 0 D 2 f ν s E e q E s 0 b 2
KR:
a 3 E s 0 D 3 f ν s E e q E s 0 b 3
a1 = 3.150/(1 − 0.3345(Lp/D)0.25
b1 = 1/3
a2 = −2.045/(1 − 0.3345(Lp/D)0.25
b2 = 9/8
a3 = 3.969/(1 − 0.3345(Lp/D)0.25
b3 = 5/3
Finite element methodThe rigid beam in homogeneous and linear inhomogeneous soils (Triangular-linear strain element)
Carter and Kulhawy [27]; Bouzid et al. [28]a1 = 1.884
b1 = 0.627
a2 = −1.048
b2 = 1.483
a3 = 1.91
b3 = 2.049
The rigid beam in homogeneous soil
Higgns et al. [19]; Bouzid et al. [28]a1 = 2.426
b1 = 0.71
a2 = −1.44
b2 = 1.67
a3 = 1.789
b3 = 2.459
The Fourier finite-element methodThe rigid beam in homogeneous soil
a1 = 0.929
b1 = 2.041
a2 = −0.633
b2 = 3.061
a3 = 0.672
b3 = 3.491
The rigid beam in linear inhomogeneous soil
Shadlou and Bhattacharya [20]a1 = 3.2
b1 = 0.62
a2 = −1.7
b2 = 1.56
a3 = 1.65
b3 = 2.5
3D element analysisThe rigid beam in homogeneous soil
a1 = 2.35
b1 = 1.53
a2 = −1.775
b2 = 2.5
a3 = 1.58
b3 = 3.45
The rigid beam in linear inhomogeneous soil
a1 = 2.66
b1 = 1.07
a2 = −1.8
b2 = 2.0
a3 = 1.63
b3 = 3.0
The rigid beam in parabolic inhomogeneous soil
Abed et al. [29]; Bouzid et al. [28]a1 = 1.708
b1 = 1.661
a2 = −1.233
b2 = 2.655
a3 = 0.672
b3 = 3.941
Fourier Series Aided Finite Element (FSAFE) approachThe rigid beam in linear inhomogeneous soil
a1 = 2.841
b1 = 0.977
a2 = −2.933
b2 = 1.767
a3 = 3.894
b3 = 2.562
The rigid beam in soil whose stiffness increases with the square root of the depth
Aissa et al. [21]a1 = 2.756
b1 = 0.668
a2 = −1.595
b2 = 1.636
a3 = 1.731
b3 = 2.495
The semi-analytical finite element analysisThe rigid beam in homogeneous soil
Carter et al. [27] KL: 3.15 G s D 2 3 L p 1 3 1 0.28 2 L p D 1 4 K LR : 2 G s D 7 8 L p 5 3 1 0.28 2 L p D 1 4   K R :   4 G s D 4 3 L p 5 3 1 0.28 2 L p D 1 4 G* is the equivalent shear modulus,
G* = Gs (1 + 0.75 νs)
The rigid beam in bedrock
Table 3. Pile geometries and soil conditions in study cases.
Table 3. Pile geometries and soil conditions in study cases.
Pile Diameter D (m)Slenderness Ratio, Lp/DDiameter-to-Wall Thickness
Ratio
Elastic Modulusvs
Pile (Ep) GPaSoil (Es)
2–102–10t = (0.01 ± 0.005)D210 (Steel pipe pile)Es = Es0(z/D)α (α = 0, 0.25, 05, 0.75, 1)0.20–0.45
2–102–10t = (0.1 ± 0.05)D30 (Concrete caissons)
Table 4. The content of KL.
Table 4. The content of KL.
ValueP00P10P01P20P11P02P30P21P12P03P31P22P13P04
α = 0−0.19461.5850.5968−0.1631−0.43790.0602500.07794−0.01022−0.0016490−0.0051560.003405−0.0006621
α = 0.250.35760.8363−0.1893−0.01239−0.016150.06096−0.010370.02333−0.027330.0032750.003736−0.0061680.004877−0.001146
α = 0.51.3870.10030.13990.1291−0.12170.03236−0.020910.04254−0.035050.010730.004781−0.0089810.007158−0.002031
α = 0.75−0.62450.58820.77910.135−0.40540.06082−0.028380.07991−0.040490.010650.006437−0.014170.01121−0.003105
α = 1.0−0.8280.70340.14630.1117−0.20190.1179−0.030420.07791−0.084250.023350.008512−0.016380.015−0.004651
Table 5. The content of KLR.
Table 5. The content of KLR.
ValueP00P10P01P20P11P02P30P21P12P03P31P22P13P04
α = 0−12.967.6160.1802−1.802−0.57580.19980.14370.2000−0.06428−0.00767−0.034090.0266−0.010540.002288
α = 0.25−9.3915.3910.8606−1.424−0.5084−0.041520.1330.1489−0.01490.003804−0.037470.0351−0.018150.003377
α = 0.5−3.0613.251−0.6617−1.265−0.13930.15740.14430.1388−0.05027−0.01146−0.048320.05271−0.028070.006426
α = 0.75−0.66764.592−3.002−1.9370.29580.41120.21460.2178−0.1840.007226−0.071740.08561−0.044990.01009
α = 1.011.21−0.8236−1.624−1.1920.0036260.15210.19710.216−0.09182−0.005945−0.083040.1031−0.062530.01531
Table 6. The content of KR.
Table 6. The content of KR.
ValueP00P10P01P20P11P02P30P21P12P03P31P22P13P04
α = 0131.2−82.5−15.4817.0514.32−2.878−1.153−3.6070.95850.038460.3045−0.09629−0.0083770.002295
α = 0.2582.54−56.65−13.3513.3810.44−0.9526−1.048−3.0210.8748−0.14540.319−0.1560.03140.001307
α = 0.576.44−58.59−10.9114.6711.83−2.367−1.209−3.71.362−0.15060.4116−0.2490.06592−0.005961
α = 0.75144.1−102.1−21.3123.4821.57−5.23−1.804−6.1662.572−0.28270.6233−0.41380.1193−0.01369
α = 1.063.22−67.89−17.919.3319.84−4.724−1.71−6.2072.814−0.42260.6904−0.52350.1853−0.0248
Table 7. Detail of parameters in studies.
Table 7. Detail of parameters in studies.
Fa (kN)Ma/Fa/Lp (kNm)DLptEsvsIpln(EP/ES0)Lp/D
20000 to 15120.4600.2515.405.52572.4
20.685.52576
Table 8. The values of KL, KLR and KR.
Table 8. The values of KL, KLR and KR.
ParametersPresent StudyShadlou et al. [20] (Rigid)Shadlou [20] (Flexible)
D = 2 mD = 5 mD = 2 mD = 5 mD = 2 m
KL124449716521166486
KLR−7310−1208−9992−6677−1141
KR78,7166606110,42669,8404879
Table 9. The relative parameters of OWT system.
Table 9. The relative parameters of OWT system.
ParametersBelwindWalney [36]Kentish Flats Offshore Wind Farm
Rated power (MW)33.6-
Mass of rotor-nacelle assembly (mRNA) (kg)130,800234,500132,000
Tower height 5367.360.06
Tower bottom diameter (m)4.354.45
Tower top diameter (m)2.332.3
Tower wall bottom thickness (mm)284126
Tower wall top thickness (mm)284115
Tower Young’s modulus (GPa)210210210
Tower density (kg/m3)786078607960
Transfer piece diameter (m)564.3
Platform height above mudline (m)3737.314.94
Monopile diameter (D) (m)564.3
Monopile thickness (mm)608045
Monopile length (m)3523.529.5
Monopile young’s modulus (GPa)210210210
Soil type221
Es0 (Mpa)153052
vs0.30.250.4
α110
From previous studies (Laszlo, [54]; Gupat et al. [36])lateral stiffness of foundation (GN/m3)1.021.530.82
cross stiffness of foundation, KLR (GN)−7.59−13.88−5.42
Rocking stiffness of foundation, KR (GN/rad)91.93205.7258.77
From this paperlateral stiffness of foundation, KL (GN/m3)0.6261.221.04
cross stiffness of foundation, KLR (GN)−5.74−12.94−5.72
Rocking stiffness of foundation, KR (GN/rad)89.24205.2666.69
Table 10. Influence of spring stiffness on the natural frequencies of the elastically supported immersed beam with eccentricity.
Table 10. Influence of spring stiffness on the natural frequencies of the elastically supported immersed beam with eccentricity.
Offshore Wind FarmModeMeasured FrequencyFrequency (ABAQUS)Error (%)
From Previous StudiesFrom This Paper
Belwind10.3720.3640.365−0.27
2-0.3790.383−1.04
3-5.0835.0061.54
4-6.2796.483−3.15
5-15.82715.7200.68
10-88.2088.090.12
18-303.26300.360.97
Walney 10.3500.3370.3370.00
2-0.4270.4045.82
4-10.87610.1956.68
5-10.88710.8760.10
10-50.18749.8230.73
20-163.28162.830.28
Kentish Flats Offshore Wind Farm10.3390.3430.3420.00
3-7.36936.7369.40
5-13.39511.95412.05
7-35.44931.33913.11
10-79.67473.4558.47
18-316.35300.315.34
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.; Wang, T.; Yin, Q.; Zhu, M. The Simplified Method of Head Stiffness Considering Semi-Rigid Behaviors of Deep Foundations in OWT Systems. Buildings 2024, 14, 1803. https://doi.org/10.3390/buildings14061803

AMA Style

Li W, Li X, Wang T, Yin Q, Zhu M. The Simplified Method of Head Stiffness Considering Semi-Rigid Behaviors of Deep Foundations in OWT Systems. Buildings. 2024; 14(6):1803. https://doi.org/10.3390/buildings14061803

Chicago/Turabian Style

Li, Wei, Xiaojuan Li, Tengfei Wang, Qian Yin, and Mingxing Zhu. 2024. "The Simplified Method of Head Stiffness Considering Semi-Rigid Behaviors of Deep Foundations in OWT Systems" Buildings 14, no. 6: 1803. https://doi.org/10.3390/buildings14061803

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