Next Article in Journal
A New Efficient Ship Detection Method Based on Remote Sensing Images by Device–Cloud Collaboration
Previous Article in Journal
Room-Temperature Creep Deformation of a Pressure-Resistant Cylindrical Structure Made of Dissimilar Titanium Alloys
Previous Article in Special Issue
A Decoupled Buckling Failure Analysis of Buried Steel Pipeline Subjected to the Strike-Slip Fault
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Numerical Investigation on the Interaction between a Tsunami-like Solitary Wave and a Monopile on a Sloping Sandy Seabed

State Key Laboratory of Ocean Engineering, Department of Civil Engineering, Shanghai Jiao Tong University, Shanghai 200240, China
*
Author to whom correspondence should be addressed.
J. Mar. Sci. Eng. 2024, 12(8), 1421; https://doi.org/10.3390/jmse12081421
Submission received: 23 July 2024 / Revised: 7 August 2024 / Accepted: 15 August 2024 / Published: 17 August 2024
(This article belongs to the Special Issue Advanced Studies in Marine Geomechanics and Geotechnics)

Abstract

:
An integrated numerical model was developed to investigate the interaction between a tsunami-like solitary wave and a monopile on a sloping sandy seabed in this study. The solitary wave motion is governed by the RANS equations with the k-ε turbulence model. The porous sloping sandy seabed is governed by Biot’s equation (u-p approximation). The solitary wave is validated with previous experimental data. Meanwhile, a further comparison of solitary wave scattering by the monopile is carried out to verify the numerical model. Then, the effects of different monopile locations were examined in investigating the solitary wave–monopile interaction problem. The velocity magnitudes and the free-surface elevation changes in the solitary wave around the monopile are investigated at various monopile locations. In addition, the response of the sloping sandy seabed and monopile under the solitary wave are examined. The numerical results demonstrate the accuracy of the current method in simulating solitary waves and wave height variation around monopiles. Wave run-up is observed in front of the monopile, with a high-velocity forward-moving water jet forming behind it. The maximum fluid velocity, wave run-up height in front of the monopile, excess pore water pressure (EPWP), and bending moment of the monopile increase as the monopile approaches the shoreline. However, at the closest location to the shoreline, due to the strong dynamic interaction between the solitary wave and the monopile, significant wave shoaling and breaking are observed, resulting in a slight decrease in the wave force acting on the monopile.

1. Introduction

The destructive power of tsunamis against coastal areas is enormous, as evidenced by the 2004 Indian Ocean tsunami and the 2011 Tohoku tsunami, which caused significant damage to coastal structures [1,2,3]. Therefore, the stability of coastal structures under the influence of tsunamis is of great importance to coastal engineers and researchers. Offshore wind energy has rapidly developed into an attractive renewable energy source in the world, and a large number of offshore wind turbines have been built [4,5]. Therefore, understanding the behavior of monopiles, the commonly used foundation type for offshore wind turbines, and the responses of the seabed under tsunami loading is essential.
Due to the unpredictability of tsunamis, most previous research was carried out through experimental and numerical methods, with solitary waves commonly used to simulate tsunami waves [6,7,8,9,10,11]. Many researchers have studied the mechanism of solitary wave run-up and breaking on a sloping seabed through potential flow theory and turbulent flow theory [12,13,14,15,16,17,18], but note that they did not consider the existence of offshore structures. Recently, the interactions between solitary waves and offshore structures, such as breakwaters and monopiles on sandy seabeds have attracted much attention [19,20]. Given that breakwaters along coastlines are extremely long, the interactions between solitary waves and breakwaters are commonly considered as two-dimensional (2D) problems in numerical investigations [7].
This is completely different from the interaction between solitary waves and monopiles, where strong three-dimensional (3D) features can be observed. When a solitary wave passes through a monopile, the characteristics of the solitary wave’s run-up in front of the monopile, the acceleration of the solitary wave on the monopile’s sides, and the reconnection wave behind the monopile are obvious [21,22,23,24]. Zhao et al. numerically studied the interaction between a solitary wave and an array of surface-piercing vertical circular cylinders [25]. Mo et al. investigated the interaction between a solitary wave and a vertical cylinder on a constant slope through both experimental and numerical methods [26]. Their results indicated that laboratory waves appear to be a little faster than numerical simulation waves. Recently, some researchers studied the interaction between solitary waves and cylinders by using different CFD models, such as OpenFOAM and REEF3D [27,28,29].
Numerous investigations on wave-induced seabed response have been conducted over the past decades [30,31,32,33,34,35,36]. However, only few researchers have focused on seabed response under solitary waves. Ye et al. numerically studied the stability of a breakwater—that was built on a sloped coast—against a solitary wave [7]. Zhao developed a 2D integrated model to investigate the effects of wave and soil characteristics on pore pressure and liquefaction [13]. Additionally, differences in wave motions on a 1:15 slope and a 1:6 slope were studied in detail. Leng et al. developed a 2D coupled approach to investigate tsunami-like solitary wave-induced soil response, where the effect of various slopes (1:5, 1:10, 1:15, and 1:20) on slope response was studied [37].
In this study, numerical simulations were conducted to study the dynamic interaction between a tsunami-like solitary wave and a monopile on a sloping sandy seabed. Solitary wave motion is governed by the RANS equations and the k-ε model. Under solitary wave loading, the response of the sloping sandy seabed is governed by Biot’s equation (u-p approximation). The numerical results are verified by previous experimental and numerical results. Subsequently, the solitary wave motions around the monopile, as well as the dynamic responses of the monopile and sloping seabed, are examined at various monopile locations.

2. Integrated Numerical Model

2.1. Governing Equations for Waves

Fluid motion is described by the RANS equations, mass conservation, and momentum conservation equations.
u f i x i = 0
ρ f u f i t + ρ f u f i u f j x j = p f x i + x j μ u f i x j + u f j x i + x j ρ f u i u j + ρ f g i
in which u f i is the ensemble mean fluid velocity, xi are the coordinates x, y, and z, ρf is the fluid density, μ is dynamic viscosity, p f is fluid pressure, t is time, and gi is the gravitational acceleration. The term ρ f u i u j in Equation (2) is the Reynolds stress which can be written as
ρ f u i u j = μ t u f i x j + u f j x i 2 3 ρ f δ i j k
where μt is the turbulent viscosity, k is the turbulence kinetic energy, and δij is the Kronecker delta. Substituting Equation (3) into Equation (2), the Equation (2) can be rewritten as
ρ f u f i t + ρ f u f i u f j x j = x i p f + 2 3 ρ f k + x j μ e f f u f i x j + u f j x i + ρ f g i
where μ e f f = μ + μ t .
The k-ε turbulence model is employed to enclose turbulence [38]:
ρ k t + ρ u j k x j = x j μ + μ t σ k k x j + ρ P k ρ ε
ρ ε t + ρ u j ε x j = x j μ + μ t σ ε ε x j + ε k C ε 1 ρ P k C ε 2 ρ ε
μ t = ρ C μ k 2 ε
in which k is TKE, ε is the rate of turbulent kinetic energy dissipation, and μt is the turbulent kinematic viscosity. The values of constants in Equations (5)–(7) are as follows: σk = 1.00, σε = 1.00, Cε1 = 1.00, Cε2 = 1.00, and Cμ = 0.99 [39].

2.2. Governing Equations of Porous Sloping Sandy Seabeds

In this study, Biot’s partly dynamic theory (u-p approximation) [31] is used to control the dynamic response of the sloping sandy seabed. The governing equations can be written as
k z 2 p γ w n β p t + k s ρ f 2 ε s t 2 = γ w ε s t
σ x x + τ x y y + τ x z z = p x + ρ 2 u t 2 τ x y x + σ y y + τ y z z = p y + ρ 2 v t 2 τ x z x + τ y z y + σ z z = p z + ρ 2 w t 2
in which p is the pore pressure; γw is the unit weight water, n is the soil porosity, ρf is the fluid density, ρ is the average density of porous sloping sandy seabed, u, v, and w are the soil displacements, σ x , σ y and σ z are the effective normal stresses, and τxy, τyz and τxz are shear stresses. εs can be written as
ε s = u x + v y + w z
where β is the compressibility of the pore fluid that can be written as
β = 1 K = 1 K w + 1 S r P w 0
where Kw is the bulk modulus of pore water, and Pw0 is the absolute water pressure. The effective normal stress and shear stress can be written as
σ x = 2 G u x + μ 1 2 μ ε s   ,   τ x z = G u z + w x = τ z x σ y = 2 G v y + μ 1 2 μ ε s   ,   τ y z = G v z + w y = τ z y σ z = 2 G w z + μ 1 2 μ ε s   ,   τ x y = G u y + v x = τ y x
where G is the shear modulus.

3. Boundary Conditions and Numerical Method

The integrated numerical model consists of a solitary wave sub-model and sloping bed sub-model, as shown in Figure 1. In the solitary wave sub-model, the symmetry boundary condition is applied to the two lateral boundaries (y-direction) to weaken the wave reflection. Since the sloping seabed is sufficiently long to simulate the solitary wave run-up and run-down, the wave does not reach the outlet area. Therefore, the outflow boundary is applied at the outlet boundary even if it has no effects on the calculation. The solitary wave solution is based on McCowan’s theory, which offers higher order accuracy than Boussinesq’s theory [40]. The governing equations for the water elevation μ, x-velocity uf1, z-velocity uf3, and wave speed c can be expressed as
η h = N M sin M 1 + η h cos M 1 + η h + cosh M x c t h
u f 1 x ,   z ,   t u ¯ g h + H = N 1 + cos M z h cosh M X h cos M z h + cosh M X h 2
u f 3 x ,   z ,   t g h + H = N sin M z h sinh M X h cos M z h + cosh M X h 2
c = u ¯ + g h + H
where H is the wave height, h is the undisturbed water depth, X = x c t , g is the gravitational acceleration, and M and N satisfy
H d = N M tan 1 2 M 1 + H h
N = 2 3 sin 2 M 1 + 2 H 3 h
The initial estimates of M and N are M = 3 H / h and N = 2 H / h . The initial estimate of water elevation η is from Boussinesq’s solution.
η h = H h sech 2 3 H 4 h x c t h
The VOF method is used to capture the free water surface in this research [41]. The volume fraction F is defined as
F = F = 0 air 0 < F < 1 interface F = 1 water
in which F satisfies the following equation:
F t + u i F x i = 0
In the porous sloping sandy seabed sub-model, the four vertical sides of the sloping sandy seabed are all fixed and impermeable.
u = v = w = p n = 0
where n donates the normal vector of the boundaries. The sloping sandy seabed bottom is treated as impermeable and rigid. The wave-induced dynamic pressure is applied on the sloping sandy seabed surface through linear interpolation.
p = p b = p f ρ f g d
where pb is the dynamic wave pressure on the sloping sandy seabed surface.
In the solitary wave sub-model, the commercial software Flow3D v11.2 was adopted to simulate the wave propagation and determine the wave pressure acting on the seabed and monopile. The scale of the model is 250 m long, 30 m wide, and 20 m high. The bed slope is 1/10. The computation domain is subdivided into the mesh of fixed hexahedron cells. The fluid motion is solved through the finite-difference method (FDM). The computational mesh of the solitary wave sub-model is shown in Figure 2. The total element number is 12,000,000. The minimum mesh size around the monopile is 0.2 m (0.1D) in x-, y-, and z-directions to achieve a satisfactory accuracy in simulating the wave motion around the monopile. The computation time for each numerical model is approximately 1 week. The presented numerical model as well as the commercial software Flow3D v11.2 were carried out with up to 16 cores on a desktop computer (CPU 3.8 GHz and 16.0 GB RAM).
In the sloping sandy seabed sub-model, FEM codes are constructed within COMSOL Multiphysics to solve Biot’s u-p equations. The sloping sandy seabed sub-model is discretized into structured tetrahedral elements, as shown in Figure 3. The maximum element size of the monopile is 0.4 m. The total element number is 103,223. In the integrated model, the solitary wave sub-model is adopted to simulate the wave motion and determines the dynamic pressures acting on the seabed and monopile. The porous sloping bed sub-model simulates the responses of the sloping sandy seabed and monopile under solitary wave loading. More details about the numerical method can be found in Zhang’s study [32].

4. Results and Discussion

4.1. Comparison of the Solitary Wave Surface Elevation

To examine the accuracy of the numerical model, a comparison of the free-surface elevation in a solitary wave was conducted firstly [26]. In the experiment, the normalized solitary wave height is H/h = 0.33, and the water depth is h = 0.205 m. As shown in Figure 4, the numerical simulation of the solitary wave agrees well with Mo’s experimental result, which confirms that the numerical approach is accurate enough to generate solitary waves.
The numerical model is further validated against a classic example of solitary wave reflection from a vertical wall. Figure 5 illustrates the comparison of the maximum solitary wave run-up height on the vertical wall with the previous experimental results [42,43], analytical results [44,45], and numerical results [46,47], in which Byatt-Smith’s analytical result is based on the Bousinessq equation that can be written as
R max h = 2 H h + H h 2
where Rmax is the maximum solitary wave run-up height on the vertical wall. Su and Mirie gave the following expression [45]:
R max h = 2 H h + 1 2 H h 2 + 3 4 H h 3
Figure 5 shows that the differences among these results are relatively small. The present numerical result is closer to Su and Mirie’s analytical result than to Byatt-Smith’s analytical result, especially for larger H/h, which is probably due to the fact that a larger wave height has strong nonlinearity so that the third-order component may have a greater impact. Additionally, the present result is close to Chen’s experimental result and Chambarel’s numerical result.

4.2. Comparison of the Maximum Run-Up Height on a Vertical Cylinder

A further comparison of solitary wave scattering by a vertical cylinder is conducted to verify the present model. Previous experimental and numerical results [27] are used for the validation. The experiment was conducted in a wave tank with dimensions of 7.6 m in length, 0.76 m in width, and 0.6 m in depth. The water depth is h = 0.04 m, the radius of the cylinder is R = 0.0635 m, and the normalized wave height is H/h = 0.4. Eight wave probes were arranged around the cylinder, as shown in Figure 6. The comparisons of the free-surface elevation along the upstream center line ( θ = 0 ° , where θ is the polar angle) are illustrated in Figure 7, in which r/R is the radial distance from the cylinder center to the wave probe. The numerical results by Chen et al. are also included in the comparison [28].
It can be observed that the calculated free-surface elevations at different wave probes upstream of the cylinder generally agree well with previous results. As the wave probes move closer to the cylinder, the maximum wave surface elevation increases. Additionally, a secondary peak is separately observed for r/R = 4.5 and 2.61, attributed to reflected waves as illustrated in Figure 7a,b. A similar result was also observed in Zhao’s numerical results [25].
Comparing with two numerical results, the present result is closer to the Yates’s experimental data for r/R = 4.5, 2.61, and 1.66. However, both the present and previous numerical results slightly overestimate the maximum wave surface elevation at r/R = 1.03. Figure 8 shows the time histories of free-surface elevations at four wave probes along the downstream centerline of the cylinder ( θ = 18 0 ° ). The present results and Chen’s numerical results show good agreement with previous findings. Both models offer improved accuracy in calculating the free-surface elevation compared to Yates and Wang’s numerical results. In general, these comparisons confirm that the numerical model used in this research is accurate for simulating the interaction between solitary waves and cylinders.

4.3. Solitary Wave Interaction with the Monopile on the Sloping Sandy Seabed

The sloping seabed is modeled with dimensions of 300 m in length, 20 m in width, and a slope ratio of 1:10. As shown in Figure 1, five different monopile locations (C1, C2, C3, C4, and C5) are simulated to investigate the solitary wave–monopile interaction problem. The other solitary wave, monopile, and sloping sandy seabed characteristics are listed in Table 1.
Figure 9 illustrates the velocity magnitudes and the variations in free-surface elevation in the solitary wave interacting with the monopile positioned at C1 at different times. Given that C1 is situated at the tip of a sloping sandy seabed, distant from the shoreline, the solitary wave’s interaction with the monopile mirrors that of a flat bed. As the solitary wave approaches the monopile, the wave run-up is observed in front of the monopile. The solitary wave separates into two symmetric waves with respect to the center line in y-direction, which then accelerate downstream, as shown in Figure 9c. The free-surface elevation decreases along the monopile surface from upstream to downstream. The two symmetric waves merge again behind the monopile. Meanwhile, a forward-moving water jet with high velocity is formed behind the monopile. As the wave progresses downstream, the wave run-down in front of the monopile is observed, with the height and velocity of the water jet decreasing gradually (Figure 9d).
Figure 10, Figure 11 and Figure 12 illustrate the velocity magnitudes and changes in free-surface elevation when the monopile is located at C2, C3, and C4. It is evident that the wave processes through the monopiles at C2 and C3 are similar to those at C1. For these three conditions, the solitary wave is distant from the shoreline as it passes through the monopile, with no significant wave shoaling or breaking observed on the sloping seabed (Figure 9, Figure 10 and Figure 11). However, the maximum fluid velocity and wave run-up height in front of the monopile increase as the monopile moves closer to the shoreline. As shown in Figure 12, the wave approaches the shoreline at the downstream of the monopile, where wave shoaling, run-up, and breaking are prominently observed. The wave front sharpens gradually as it reaches to the shoreline.
Figure 13 shows the entire process of wave shoaling, breaking, run-up, and drawdown on the sloping sandy seabed when the monopile is located at C5. The solitary wave breaks before its peak reaches the monopile. Simultaneously, the velocity magnitude of the water increases gradually as the wave climbs the sloping sandy seabed leading to a strong dynamic interaction between the solitary wave and the monopile, as shown in Figure 13c. The water rises up higher in front of the monopile, and the fluid velocity intensifies, with some sea water even splashing into the air. At this moment, the water jet is formed gradually behind the monopile. After the solitary wave passes the monopile, the water jet and the sea water, along with the splashed waters falling from the air, rapidly climb the sloping sandy seabed (Figure 13d,e). The height of the water jet and the maximum fluid velocity gradually decrease. At t = 30 s, the solitary wave reaches its farthest distance on the sloping sandy seabed and begins to draw down under the effect of gravity (Figure 13f). As the fluid flows down from the sloping sandy seabed, the water rises behind the monopile, forming a water jet in front of the monopile (Figure 13g,h).
Figure 14 presents the time histories of free-surface elevation at three wave probes for the monopile positioned at C1, C2, C3, C4, and C5, respectively. The maximum wave elevation at wave probe A is significantly higher than at wave probes B and C when the monopile is located at C1 (Figure 14a). Additionally, a secondary peak is observed in front of the monopile at t = 14.5 s, due to the effects of the reflection wave. As the monopile shifts from C1 to C4, the maximum free-surface elevation in wave probe A increases, and the secondary peak gradually disappears (Figure 14a–c). When the monopile is located at C5, the maximum free-surface elevation in front of the monopile rapidly increases to approximately 15.2 m, then decreases to a constant value of about 14.8 m. The maximum free-surface elevation at wave probe B surpasses that at wave probe A, due to the strong dynamic interaction between the solitary wave and the monopile. These results are also reflected in Figure 15a. It is evident that as the monopile location moves from C1 to C5, the maximum free-surface elevations at wave probes A and B increase significantly. However, the maximum free-surface elevations at wave probe C remain almost unchanged across different monopile locations. Figure 15b displays the maximum fluid velocity when the solitary passes the monopile at different locations. Comparing these results, it is evident that the fluid velocity gradually increases as the solitary climbs the sloping sandy seabed. The maximum fluid velocity increases from approximately 4.1 m/s to 15.5 m/s as the monopile moves from C1 to C5.
Figure 16 illustrates the cloud figures of the solitary wave-induced excess pore water pressure (EPWP) in the sloping sandy seabed, capturing the moment when the solitary wave passes the monopile. At this point, the EPWP around the monopile achieves its peak value. It is observed that the EPWP decreases in vertically from the surface of the sloping sandy seabed. However, the effect of the solitary wave on the seabed is confined to a narrow region.
Figure 17 illustrates the distributions of EPWP along the vertical direction at three gauge locations around the monopile, with the monopile bottom located at z = −25 m. As depicted in Figure 17a, the EPWP in front of the monopile (G1) along the vertical direction decreases rapidly. The maximum EPWP at the seabed surface increases as the monopile moves from C1 to C4. However, when the monopile is located at C5, the maximum EPWP at G1 is significantly lower than that at C3 and C4. This is mainly because C5 is at the shoreline, where the solitary wave is already breaking upon arrival. Additionally, the effects of the monopile bottom on the EPWP distributions around z = −25 m are significant. Figure 17b illustrates the distributions of the EPWP at the lateral side of the monopile (G2). Similar trends in EPWP distributions are observed as observed in Figure 17a. However, the maximum EPWP at G2 for both monopile locations is smaller than at G1. Figure 17c displays the vertical distributions of EPWP at G3 for different monopile locations. The maximum EPWP at the seabed surface increases as the monopile moves from C1 to C2. However, as the wave breaks while climbing the sloping seabed, the maximum EPWP at the seabed surface becomes irregular as the monopile moves from C3 to C5. Overall, the effects of the solitary wave on the sloping sandy seabed are significant.
The response of the monopile under the solitary wave is also examined. Figure 18 shows the distribution of the bending moment M of the monopile along the vertical direction for various monopile locations. It is evident that the maximum M occurs near the sloping sandy seabed surface for each monopile locations with different height. In addition, the maximum M slightly decreases from about 450 kN·m to 400 kN·m as the monopile moves from C1 to C5, indicating that the solitary wave forces acting on the monopile decrease as the monopile moves from C1 to C5. The bending moments at both the monopile’s top and bottom are zero, which is reasonable since the monopile head is unconstrained in the air and the insertion depth of the monopile below the sloping sandy seabed is sufficiently deep.

5. Conclusions

In this study, the dynamic interaction between the tsunami-like solitary wave and monopile on a sloping sandy seabed was investigated numerically. An integrated numerical model was developed for the investigation, in which the solitary wave sub-model is governed by the RANS equations and the porous sloping sandy seabed sub-model is governed by Biot’s partly dynamic theory (u-p approximation). The main conclusions are as follows:
(1) The present numerical model is accurate for simulating solitary waves. Due to the larger wave height’s strong nonlinearity, the numerical result of the maximum run-up height on a vertical wall is closer to Su and Mirie’s analytical result.
(2) The comparison of the free-surface elevations at different wave probes around the cylinder is set up between the present model and previous results. Good agreement is observed. The maximum wave surface elevation increases as the wave probes move towards the cylinder upstream. Due to the reflected waves, a secondary peak is separately observed for r/R = 4.5 and 2.61 upstream the cylinder.
(3) When the monopile is located at C1, C2, and C3, the wave run-up is observed in front of the monopile, and a forward-moving water jet with a high velocity is formed behind it. When the monopile is located at C4 and C5, wave shoaling and breaking on the sloping sandy seabed are observed. Strong dynamic interaction between the solitary wave and the monopile is observed when the monopile is located at C5, with some sea water even splashing into the air.
(4) The EPWP decreases quickly in a vertical direction from the seabed surface. The maximum EPWP at the seabed surface at G1 and G2 increases as the monopile moves from C1 to C4. Due to the strong dynamic interaction between the solitary wave and monopile at C4 and C5, the maximum EPWP at G3 becomes irregular as the monopile moves from C3 to C5. The maximum M occurs near the sloping sandy seabed surface. The maximum M slightly decreases as the monopile moves from C1 to C5.
(5) This study investigates the dynamic response of a monopile and sloping sandy seabed under the excitation of a solitary wave, focusing on variations in the interaction between the monopile and solitary wave at different slope locations, differences in wave flow around the monopile, and variations in the bending moment response of the monopile. This study provides a valuable reference for understanding the response of offshore monopile foundations subjected to solitary wave excitation on sloping sandy seabeds.

Author Contributions

Conceptualization, W.X. and Q.Z.; formal analysis, W.X.; investigation, W.X. and H.C.; resources, Q.Z.; data curation, M.F.; writing—original draft preparation, W.X.; writing—review and editing, Q.Z.; visualization, H.C.; supervision, Q.Z. and M.F. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by National Natural Science Foundation of China (Grant No. 42207186), the Natural Science Foundation of Shanghai (Grant No. 22ZR1435000).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The original contributions presented in this study are included in this article; further inquiries can be directed to the corresponding author.

Conflicts of Interest

The authors declare no conflicts of interest.

Glossary

u f i ensemble mean fluid velocity
xicoordinates x, y, and z
ρffluid density
ρaverage density of porous sloping sandy seabed
μdynamic viscosity
p f fluid pressure
ttime
gigravitational acceleration
μtturbulent viscosity
kturbulence kinetic energy
δijKronecker delta
εrate of turbulent kinetic energy dissipation
μtturbulent kinematic viscosity
ppore pressure
γwunit weight water
nsoil porosity
u, v, and wsoil displacements
σ x , σ y and σ z effective normal stresses
τxy, τyz, and τxzshear stresses
βcompressibility of pore fluid
Kwbulk modulus of pore water
Pw0absolute water pressure
Gshear modulus
Hwave height
hundisturbed water depth
ηinitial estimate of water elevation
Fvolume fraction
nnormal vector of boundaries
pbdynamic wave pressure on sloping sandy seabed surface
Rmaxmaximum solitary wave run-up height on vertical wall
ρsdensity of sloping sandy seabed
vsPoisson ratio of sloping sandy seabed
EsYoung’s modulus of sloping sandy seabed
kzpermeability of sloping sandy seabed
Srdegree of saturation of sloping sandy seabed
EmYoung’s modulus of monopole
vmPoisson ratio of monopole
ρmdensity of monopole
Dmonopile diameter

References

  1. Chen, J.; Huang, Z.; Jiang, C.; Deng, B.; Long, Y. Tsunami-Induced Scour at Coastal Roadways: A Laboratory Study. Nat. Hazards 2013, 69, 655–674. [Google Scholar] [CrossRef]
  2. Bricker, J.D.; Francis, M.; Nakayama, A. Scour Depths near Coastal Structures due to the 2011 Tohoku Tsunami. J. Hydraul. Res. 2012, 50, 637–641. [Google Scholar] [CrossRef]
  3. Tanaka, N.; Sato, M. Scoured Depth and Length of Pools and Ditches Generated by Overtopping Flow from Embankments during the 2011 Great East Japan Tsunami. Ocean Eng. 2015, 109, 72–82. [Google Scholar] [CrossRef]
  4. Lombardi, D.; Bhattacharya, S.; Muir Wood, D. Dynamic Soil-Structure Interaction of Monopile Supported Wind Turbines in Cohesive Soil. Soil Dyn. Earthq. Eng. 2013, 49, 165–180. [Google Scholar] [CrossRef]
  5. Jung, S.; Kim, S.R.; Patil, A.; Hung, L.C. Effect of Monopile Foundation Modeling on the Structural Response of a 5-MW Offshore Wind Turbine Tower. Ocean Eng. 2015, 109, 479–488. [Google Scholar] [CrossRef]
  6. Sassa, S.; Takahashi, H.; Morikawa, Y.; Takano, D. Effect of Overflow and Seepage Coupling on Tsunami-Induced Instability of Caisson Breakwaters. Coast. Eng. 2016, 117, 157–165. [Google Scholar] [CrossRef]
  7. Ye, J.; Jeng, D.; Ren, W.; Zhu, C. Numerical Study of the Stability of Breakwater Built on a Sloped Porous Seabed under Tsunami Loading. Appl. Math. Model. 2013, 37, 9575–9590. [Google Scholar]
  8. Clamond, D.; Dutykh, D. Fast accurate computation of the fully nonlinear solitary surface gravity waves. Comput. Fluids 2013, 84, 35–38. [Google Scholar] [CrossRef]
  9. Dutykh, D.; Clamond, D. Efficient computation of steady solitary gravity waves. Wave Motion 2014, 51, 86–99. [Google Scholar] [CrossRef]
  10. Abdalazeez, A.; Didenkulova, I.; Dutykh, D. Nonlinear deformation and run-up of single tsunami waves of positive polarity: Numerical simulations and analytical predictions. Nat. Hazards Earth Syst. Sci. 2019, 19, 2905–2913. [Google Scholar] [CrossRef]
  11. Watanabe, M.; Arikawa, T.; Kihara, N.; Tsurudome, C.; Hosaka, K.; Kimura, T.; Hashimoto, T.; Ishihara, F.; Shikata, T.; Morikawa, D.; et al. Validation of tsunami numerical simulation models for an idealized coastal industrial site. Coast. Eng. J. 2022, 64, 302–343. [Google Scholar] [CrossRef]
  12. Grilli, S. Shoaling of Solitary Waves on Plane Beaches. J. Waterw. Port Coast. Ocean. Eng. 2008, 120, 609–628. [Google Scholar] [CrossRef]
  13. Zhao, H.; Jeng, D.; Zhang, H.; Zhang, J.; Zhang, H. 2-D Integrated Numerical Modeling for the Potential of Solitary Wave-Induced Residual Liquefaction over a Sloping Porous Seabed. J. Ocean. Eng. Mar. Energy 2016, 2, 1–18. [Google Scholar] [CrossRef]
  14. Hsieh, C.; Hwang, R.R.; Peng, Y.; Yang, W. Numerical Simulations of Solitary Wave Running and Breaking on a Slopping bed. Isope 2007, 2321–2326. [Google Scholar]
  15. Liang, D.; Gotoh, H.; Khayyer, A. Boussinesq Modelling of Solitary Wave and N-Wave Runup on Coast. Appl. Ocean Res. 2013, 42, 144–154. [Google Scholar] [CrossRef]
  16. Zelt, J.A. The Run-up of Nonbreaking and Breaking Solitary Waves. Coast. Eng. 1991, 15, 205–246. [Google Scholar] [CrossRef]
  17. Abdalazeez, A.; Didenkulova, I.; Dutykh, D.; Denissenko, P. Comparison of dispersive and nondispersive models for wave run-up on a beach. Izv. Atmos. Ocean. Phys. 2020, 56, 494–501. [Google Scholar] [CrossRef]
  18. Abdalazeez, A.; Didenkulova, I.; Dutykh, D. Dispersive effects during long wave run-up on a plane beach. In Advances in Natural Hazards and Hydrological Risks: Meeting the Challenge: Proceedings of the 2nd International Workshop on Natural Hazards (NATHAZ’19), Pico Island—Azores; Springer: Cham, Switzerland, 2019; pp. 143–146. [Google Scholar]
  19. Yao, Y.; He, T.; Deng, Z.; Chen, L.; Guo, H. Large eddy simulation modeling of tsunami-like solitary wave processes over fringing reefs. Nat. Hazards Earth Syst. Sci. 2019, 19, 1281–1295. [Google Scholar] [CrossRef]
  20. Tripepi, G.; Aristodemo, F.; Veltri, P. On-bottom stability analysis of cylinders under tsunami-like solitary waves. Water 2018, 10, 487. [Google Scholar] [CrossRef]
  21. Isaacson, M. Solitary Wave Diffraction around Large Cylinder. J. Waterw. Port Coast. Ocean Eng. 1977, 109, 121–127. [Google Scholar] [CrossRef]
  22. Yates, G.T.; Wang, K.H. Solitary Wave Scattering by a Vertical Cylinder. In Proceedings of the The Fourth International Offshore and Polar Engineering Conference, Osaka, Japan, 10–15 April 1994. [Google Scholar]
  23. Ohyama, T. A Numerical Method of Solitary Wave Forces Acting on a Large Vertical Cylinder. Coast. Eng. 1991, 1990, 840–852. [Google Scholar]
  24. Isaacson, M. Nonlinear-Wave Effects on Fixed and Floating Bodies. J. Fluid Mech. 2006, 120, 267–281. [Google Scholar] [CrossRef]
  25. Zhao, M.; Cheng, L.; Teng, B. Numerical Simulation of Solitary Wave Scattering by a Circular Cylinder Array. Ocean Eng. 2007, 34, 489–499. [Google Scholar] [CrossRef]
  26. Mo, W.; Jensen, A.; Liu, P. Plunging Solitary Wave and Its Interaction with a Slender Cylinder on a Sloping Beach. Ocean Eng. 2013, 74, 48–60. [Google Scholar] [CrossRef]
  27. Cao, H.; Wan, D. RANS-VOF Solver for Solitary Wave Run-up on a Circular Cylinder. China Ocean Eng. 2015, 29, 183–196. [Google Scholar] [CrossRef]
  28. Chen, Q.; Zang, J.; Kelly, D.M.; Dimakopoulos, A.S. A 3-D Numerical Study of Solitary Wave Interaction with Vertical Cylinders Using a Parallelised Particle-in-Cell Solver. J. Hydrodyn. 2017, 29, 790–799. [Google Scholar] [CrossRef]
  29. Alagan, C.M.; Bihs, H.; Myrhaug, D.; Muskulus, M. Breaking Solitary Waves and Breaking Wave Forces on a Vertically Mounted Slender Cylinder over an Impermeable Sloping Seabed. J. Ocean Eng. Mar. Energy 2017, 3, 1–19. [Google Scholar] [CrossRef]
  30. Biot, M.A. Mechanics of Deformation and Acoustic Propagation in Porous Media. J. Appl. Phys. 1962, 33, 1482–1498. [Google Scholar] [CrossRef]
  31. Zienkiewicz, O.C.; Chang, C.T.; Bettess, T. Drained, Undrained, Consolidating and Dynamic Behaviour Assumptions in Soils. Geotechnique 1980, 30, 385–395. [Google Scholar] [CrossRef]
  32. Zhang, Q.; Zhou, X.; Wang, J.; Guo, J. Wave-Induced Seabed Response around an Offshore Pile Foundation Platform. Ocean Eng. 2017, 130, 567–582. [Google Scholar] [CrossRef]
  33. Jeng, D.S.; Rahman, M.S. Effective Stresses in a Porous Seabed of Finite Thickness: Inertia Effects. Can. Geotech. J. 2011, 37, 1383–1392. [Google Scholar] [CrossRef]
  34. Zhou, X.; Zhang, J.; Wang, J.; Xu, Y.; Jeng, D. Stability and Liquefaction Analysis of Porous Seabed Subjected to Cnoidal Wave. Appl. Ocean Res. 2014, 48, 250–265. [Google Scholar] [CrossRef]
  35. Liao, C.; Zhao, H.; Jeng, S. Poro-Elasto-Plastic Model for the Wave-Induced Liquefaction. J. Offshore Mech. Arct. Eng. 2015, 137, 042001. [Google Scholar] [CrossRef]
  36. Tong, D.; . Liao, C.; Jeng, D.; Zhang, L.; Wang, J.; Chen, L. Three-Dimensional Modeling of Wave-Structure-Seabed Interaction around Twin-Pile Group. Ocean Eng. 2017, 145, 416–429. [Google Scholar] [CrossRef]
  37. Leng, J.; Ye, G.; Liao, C.; Jeng, D. On the Soil Response of a Coastal Sandy Slope Subjected to Tsunami-like Solitary Wave. Bull. Eng. Geol. Environ. 2018, 77, 999–1014. [Google Scholar] [CrossRef]
  38. Launder, B.E.; Spalding, D.B. The Numerical Computation of Turbulent Flows. Comput. Methods Appl. Mech. Eng. 1974, 3, 269–289. [Google Scholar] [CrossRef]
  39. Rodi, W. Turbulence Models and Their Application in Hydraulics: A State of the Art Review; International Association for Hydraulic Research, 1984; Available online: https://www.taylorfrancis.com/books/mono/10.1201/9780203734896/turbulence-models-application-hydraulics-wolfgang-rodi (accessed on 14 August 2024).
  40. Munk, W.H. The Solitary Wave Theory and Its Application to Surf Problems. Ann. N. Y. Acad. Sci. 1949, 51, 376–424. [Google Scholar] [CrossRef]
  41. Hirt, C.W.; Nichols, B.D. Volume of Fluid (VOF) Method for the Dynamics of Free Boundaries. J. Comput. Phys. 1981, 39, 201–225. [Google Scholar] [CrossRef]
  42. Maxworthy, T. Experiments on Collisions between Solitary Waves. J. Fluid Mech. 1976, 76, 177–186. [Google Scholar] [CrossRef]
  43. Chen, Y.Y.; Kharif, C.; Yang, J.H.; Hsu, H.C.; Touboul, J.; Chambarel, J. An Experimental Study of Steep Solitary Wave Reflection at a Vertical Wall. Eur. J. Mech.—B/Fluids 2015, 49, 20–28. [Google Scholar] [CrossRef]
  44. Byatt-Smith, J.G.B. An Integral Equation for Unsteady Surface Waves and a Comment on the Boussinesq Equation. J. Fluid Mech. 1971, 49, 625–633. [Google Scholar] [CrossRef]
  45. Su, C.H.; Mirie, R.M. On Head-on Collisions between Two Solitary Waves. J. Fluid Mech. 1980, 39, 781–791. [Google Scholar] [CrossRef]
  46. Chambarel, J.; Kharif, C.; Touboul, J. Head-on Collision of Two Solitary Waves and Residual Falling Jet Formation. Nonlinear Process. Geophys. 2009, 16, 111–122. [Google Scholar] [CrossRef]
  47. Sun, J.L.; Wang, C.Z.; Wu, G.X.; Khoo, B.C. Fully Nonlinear Simulations of Interactions between Solitary Waves and Structures Based on the Finite Element Method. Ocean Eng. 2015, 108, 202–215. [Google Scholar] [CrossRef]
Figure 1. Sketch of the numerical model.
Figure 1. Sketch of the numerical model.
Jmse 12 01421 g001
Figure 2. Mesh of the solitary wave sub-model.
Figure 2. Mesh of the solitary wave sub-model.
Jmse 12 01421 g002
Figure 3. Mesh of the sloping bed sub-model.
Figure 3. Mesh of the sloping bed sub-model.
Jmse 12 01421 g003
Figure 4. Comparison of the free-surface elevation in the solitary wave [26].
Figure 4. Comparison of the free-surface elevation in the solitary wave [26].
Jmse 12 01421 g004
Figure 5. Comparison of the maximum run-up height versus wave amplitude [25,42,43,44,45,46,47].
Figure 5. Comparison of the maximum run-up height versus wave amplitude [25,42,43,44,45,46,47].
Jmse 12 01421 g005
Figure 6. Sketch of the arrangement of the wave probes.
Figure 6. Sketch of the arrangement of the wave probes.
Jmse 12 01421 g006
Figure 7. Comparisons of the free-surface elevation at different wave probes at the upstream of the cylinder [22,43]. (a) WP1, (b) WP2, (c) WP3, (d) WP4.
Figure 7. Comparisons of the free-surface elevation at different wave probes at the upstream of the cylinder [22,43]. (a) WP1, (b) WP2, (c) WP3, (d) WP4.
Jmse 12 01421 g007
Figure 8. Comparisons of the free-surface elevation at different wave probes at the downstream of the cylinder [22,44]. (a) WP5, (b) WP6, (c) WP7, (d) WP8.
Figure 8. Comparisons of the free-surface elevation at different wave probes at the downstream of the cylinder [22,44]. (a) WP5, (b) WP6, (c) WP7, (d) WP8.
Jmse 12 01421 g008aJmse 12 01421 g008b
Figure 9. Velocity magnitudes and the free-surface elevation changes in the solitary wave interacting with the monopile that is located at C1. (a) t = 9.5 s, (b) t = 10.5 s, (c) t = 11.5 s, (d) t = 12.5 s.
Figure 9. Velocity magnitudes and the free-surface elevation changes in the solitary wave interacting with the monopile that is located at C1. (a) t = 9.5 s, (b) t = 10.5 s, (c) t = 11.5 s, (d) t = 12.5 s.
Jmse 12 01421 g009aJmse 12 01421 g009b
Figure 10. Velocity magnitudes and the free-surface elevation changes in the solitary wave interacting with the monopile that is located at C2. (a) t = 12 s, (b) t = 13 s, (c) t = 14 s, (d) t = 15 s.
Figure 10. Velocity magnitudes and the free-surface elevation changes in the solitary wave interacting with the monopile that is located at C2. (a) t = 12 s, (b) t = 13 s, (c) t = 14 s, (d) t = 15 s.
Jmse 12 01421 g010
Figure 11. Velocity magnitudes and the free-surface elevation changes in the solitary wave interacting with the monopile that is located at C3. (a) t = 14 s, (b) t = 15 s, (c) t = 16 s, (d) t = 17 s.
Figure 11. Velocity magnitudes and the free-surface elevation changes in the solitary wave interacting with the monopile that is located at C3. (a) t = 14 s, (b) t = 15 s, (c) t = 16 s, (d) t = 17 s.
Jmse 12 01421 g011
Figure 12. Velocity magnitudes and the free-surface elevation changes in the solitary wave interacting with the monopile that is located at C4. (a) t = 16.5 s, (b) t = 17.5 s, (c) t = 18.5 s, (d) t = 19.5 s.
Figure 12. Velocity magnitudes and the free-surface elevation changes in the solitary wave interacting with the monopile that is located at C4. (a) t = 16.5 s, (b) t = 17.5 s, (c) t = 18.5 s, (d) t = 19.5 s.
Jmse 12 01421 g012
Figure 13. Velocity magnitudes and the free-surface elevation changes in the solitary wave interacting with the monopile that is located at C5. (a) t = 16.5 s, (b) t = 17.5 s, (c) t = 20.5 s, (d) t = 22 s, (e) t = 25 s, (f) t = 30 s, (g) t = 35 s, (h) t = 40 s.
Figure 13. Velocity magnitudes and the free-surface elevation changes in the solitary wave interacting with the monopile that is located at C5. (a) t = 16.5 s, (b) t = 17.5 s, (c) t = 20.5 s, (d) t = 22 s, (e) t = 25 s, (f) t = 30 s, (g) t = 35 s, (h) t = 40 s.
Jmse 12 01421 g013
Figure 14. Time histories of free-surface elevation at three wave probes of the monopile at (a) C1, (b) C2, (c) C3, (d) C4, and (e) C5, respectively.
Figure 14. Time histories of free-surface elevation at three wave probes of the monopile at (a) C1, (b) C2, (c) C3, (d) C4, and (e) C5, respectively.
Jmse 12 01421 g014
Figure 15. Numerical results of the (a) maximum free-surface elevations around the monopile, and (b) maximum fluid velocity when the solitary passes through the monopile at different locations.
Figure 15. Numerical results of the (a) maximum free-surface elevations around the monopile, and (b) maximum fluid velocity when the solitary passes through the monopile at different locations.
Jmse 12 01421 g015
Figure 16. Solitary wave-induced EPWP in the sloping seabed around the monopile at different locations. (a) C1, (b) C2, (c) C3, (d) C4, (e) C5.
Figure 16. Solitary wave-induced EPWP in the sloping seabed around the monopile at different locations. (a) C1, (b) C2, (c) C3, (d) C4, (e) C5.
Jmse 12 01421 g016aJmse 12 01421 g016b
Figure 17. Distributions of the EPWP along the vertical direction around the monopile for various monopile locations. (a) G1, (b) G2, (c) G3.
Figure 17. Distributions of the EPWP along the vertical direction around the monopile for various monopile locations. (a) G1, (b) G2, (c) G3.
Jmse 12 01421 g017
Figure 18. Distributions of the bending moment of the monopile along the vertical direction for various monopile locations.
Figure 18. Distributions of the bending moment of the monopile along the vertical direction for various monopile locations.
Jmse 12 01421 g018
Table 1. Calculation parameters of the numerical model.
Table 1. Calculation parameters of the numerical model.
Water Characteristics
Water Depth (h)10 (m)Wave Height (H)3 (m)
Slopping seabed characteristics
Density (ρs)1850 (kg/m3)Young’s modulus (Es)4.05 × 107 (N/m2)
Poisson ratio (vs)3.448Permeability (kz)1 × 10−3 (m/s)
Porosity (n)0.4Degree of saturation (Sr)0.98
Monopile characteristics
Young’s modulus (Em)0.5Density (ρm)2300 (kg/m3)
Poisson ratio (vm)0.25Monopile diameter (D)2 (m)
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

Xie, W.; Zhang, Q.; Cai, H.; Fu, M. Numerical Investigation on the Interaction between a Tsunami-like Solitary Wave and a Monopile on a Sloping Sandy Seabed. J. Mar. Sci. Eng. 2024, 12, 1421. https://doi.org/10.3390/jmse12081421

AMA Style

Xie W, Zhang Q, Cai H, Fu M. Numerical Investigation on the Interaction between a Tsunami-like Solitary Wave and a Monopile on a Sloping Sandy Seabed. Journal of Marine Science and Engineering. 2024; 12(8):1421. https://doi.org/10.3390/jmse12081421

Chicago/Turabian Style

Xie, Wenbo, Qi Zhang, Hao Cai, and Miao Fu. 2024. "Numerical Investigation on the Interaction between a Tsunami-like Solitary Wave and a Monopile on a Sloping Sandy Seabed" Journal of Marine Science and Engineering 12, no. 8: 1421. https://doi.org/10.3390/jmse12081421

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