Next Article in Journal
Assessing the Performance of Shipboard Instruments Used to Monitor Total Residual Oxidants
Previous Article in Journal
Self-Organizing of Waves and Sandy Bottom Relief—Laboratory Experiments
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Numerical Investigation on the Hydrodynamic Characteristics of Submarine Power Cables for Offshore Wind Turbines Under Combined Wave–Current Loading

1
Haikou Sub-Bureau of Guangzhou Bureau, Extra High Voltage Transmission Company of China Southern Power Grid Co., Ltd., Haikou 570216, China
2
State Key Laboratory of Coastal and Offshore Engineering, Dalian University of Technology, Dalian 116024, China
*
Author to whom correspondence should be addressed.
J. Mar. Sci. Eng. 2025, 13(11), 2067; https://doi.org/10.3390/jmse13112067
Submission received: 23 September 2025 / Revised: 24 October 2025 / Accepted: 28 October 2025 / Published: 29 October 2025
(This article belongs to the Topic Wind, Wave and Tidal Energy Technologies in China)

Abstract

A 2D numerical model for viscous flow is established in OpenFOAM version 10 to analyze the hydrodynamic response of submarine power cables for offshore wind turbines under combined wave–current conditions. It focuses on analyzing the effect of the cable suspension ratio e/D and the current-to-wave velocity ratio Uc/Um on the Morison coefficient of the suspended cable. The results indicate that for the cable suspension ratio e/D of less than 0.5, the strength of the dependence of both the drag coefficient Cd and inertia coefficient CM on the cable suspension ratio e/D is significantly influenced by the current-wave-ratio Uc/Um, while this dependence becomes less pronounced for e/D greater than 0.5. And the inertia force coefficient CM decreases monotonically with the current-to-wave velocity ratio Uc/Um, while the drag force coefficient Cd demonstrates a more complex, non-monotonic relationship with it. Based on the simulation results in this paper, a quantitative relationship between Cd, CM, and the key governing parameters is established using a two-layer feedforward neural network model, providing a method for predicting wave–current forces on subsea suspended cables.

1. Introduction

As an integral component of marine renewable energy systems, submarine cables connect onshore and offshore facilities, playing a pivotal role in the transmission of electric power harnessed from wind, wave, and tidal energy. This role is particularly critical in China, given its national strategic emphasis on energy transition and its ambitious “Dual Carbon” goals, which have catalyzed the rapid expansion of offshore wind turbines along the coastline. However, complex seabed topography or seabed scouring due to ocean currents can result in suspended spans of submarine cables. And the on-bottom stability and structural safety of subsea cables are governed by the hydrodynamic loads on the suspended spans [1,2]. Therefore, conducting research on the hydrodynamic characteristics of suspended submarine cables for offshore wind turbines holds significant engineering importance.
The suspended span of a submarine cable may be formed due to local scour that extends axially along the cable. Although the scouring process exhibits typical three-dimensional characteristics, as it gradually expands along the cable to a reasonable stage, the scouring and force characteristics at the cable cross-section can be considered to be a two-dimensional process. Liang et al. [3] compared the results of two-dimensional numerical simulations of local scouring with experimental data from three-dimensional physical models, indicating that the two-dimensional model can effectively simulate the development process of local scouring, and the suspended height of the cylinder remains basically at the order of 1.0D, where D represents the cylinder diameter. Cheng et al. [4] conducted a three-dimensional scour experiment, measuring the local scour depth beneath the cylinder using a specially developed conductivity scour probe. A prediction formula for scour depth was derived from the results. Bearman et al. [5] studied the flow around cylinders at different heights above a plane boundary through wind tunnel experiments and found that the vortex-shedding frequency is inhibited by the plane boundary, and the vibration response of the cylinder is, consequently, affected. Lei et al. [6] conducted a numerical study on the drag coefficient of a circular cylinder in unidirectional flow, investigating Reynolds numbers of 80–1000 and clearance ratios of 0.1–3.0. Here, the Reynolds number Re under unidirectional flow conditions is defined as Re = UcD/υ, where Uc is the unidirectional flow velocity, υ is the kinematic viscosity of the water, and the clearance ratio is defined as e/D, where e is the clearance between the cylinder bottom and the wall. Similarly, Ong et al. [7] studied the vortex-shedding mechanism of a near-wall circular cylinder at a Reynolds number Re = 1.31 × 104 and clearance ratios of 0.1 ≤ e/D ≤ 1.0 based on a two-dimensional k-ε turbulence model. The study showed that vortex shedding is suppressed at e/D ≤ 0.3, resulting in a decrease in the root mean square coefficient of the cylinder. According to Lei et al. [8], the critical clearance ratio for vortex suppression ranges from 0.2 to 0.3 and varies inversely with the boundary layer thickness δ. The drag coefficient’s dependence on the clearance ratio is dominated by the cylinder’s position relative to the boundary layer. It remains nearly constant outside the layer, increases with e/δ within the layer, and becomes negligible for e/D > 2.0. Buresti and Lanciotti [9] reported that the mean lift coefficient decreases with the clearance ratio, and the magnitude of this reduction is directly proportional to the boundary layer thickness, making the drop more dramatic in thicker layers. However, the average drag coefficient varies non-monotonically with the clearance ratio, a trend that is highly sensitive to both the Reynolds number and the boundary layer thickness. For oscillatory flow conditions, An et al. [10] conducted 2D and 3D simulations within the parameter range of Re = 2000 and KC = 1~26.2 (under oscillatory flow conditions, Re = UmD/υ and KC = UmT/D, where Um and T are defined as the velocity amplitude and period of the oscillatory flow). The computational results indicated that the vortex structures obtained from the 2D and 3D models under the same conditions were essentially consistent, and the predicted differences in the Morison force coefficient were within 18%. Munir et al. [11] investigated the influence of the plane boundary on the vortex-induced vibration (VIV) using a two-degree-of-freedom (2DOF) VIV model. The results showed that the reduced velocity Vr (Vr = Um/(fnD), where fn is the natural frequency of the cylinder) affects the vortex shedding significantly, and the vortices are shed only from the bottom side of the cylinder due to the effect of the shear layer on the plane boundary.
To date, there has been limited research on the hydrodynamic problems of cylinders subjected to combined unidirectional and oscillatory flow. Tong et al. [12] used a viscous flow model to study the flow field structure around a forced oscillating cylinder in uniform flow at Re = 100 and explored the influence of different oscillation directions on the flow field structure. Teng et al. [13] utilized a numerical model to calculate the forces acting on a horizontally oscillating cylinder under forced excitation and analyzed the amplitude differences in various harmonic forces through flow field visualization. Zhao et al. [14] employed a 3D direct numerical simulation to study current-to-wave ratio effects on the force acting on the cylinder, finding that the frequency of lift force increases with the current-to-wave ratio. Cao and Li [15] found that the vortex shedding under the combined steady and oscillatory flow is not only dependent on the Reynolds number but also on the KC number and AR number (AR = Um/Uc). Deng et al. [16] numerically investigated the flow around a transversely oscillating circular cylinder subjected to combined uniform and oscillatory flow. Their study found that the vortex formation length near the cylinder decreases as the amplitude of the oscillatory flow increases. Furthermore, the drag and inertia force coefficients increase when the frequency ratio f/fs (where f is the oscillation frequency of the cylinder and fs is the vortex-shedding frequency) approaches the lock-in region. Zhao et al. [17] investigated the vortex-induced vibration of a circular cylinder subjected to combined steady and oscillatory flow and found that the lock-in regime under this condition is wider than that in either pure steady flow or pure oscillatory flow. Chang et al. [18] investigated the flow structure around a vertical cylinder. They found that under a constant KC number, both the strength and lifespan of the main horseshoe vortex formed during each oscillatory cycle increase with the velocity ratio Uc/Um. Additionally, the phase lag between the force and the oscillatory velocity component decreases with increasing KC number and shows little dependence on Uc/Um. Teng et al. [19] investigated the hydrodynamic forces on small-diameter cylinders, finding that Fourier coefficients can effectively reconstruct the time-history of hydrodynamic forces, and provided empirical formulas for peak coefficients and Fourier coefficients as functions of KC and Uc/Um.
However, under combined wave–current loading, the forces on a suspended cable are influenced by multiple control parameters, including KC, e/D, and Uc/Um. Scientifically quantifying the relationship between the hydrodynamic coefficients of the cylinder and these control parameters is key to predicting the forces on subsea suspended cables. Therefore, this paper investigates the hydrodynamics of subsea suspended cables under combined wave–current conditions, utilizing a viscous flow numerical model. Based on the simulation results, the present study further proposes a method for predicting the forces on cables under combined wave–current loading, providing a reference for related engineering design.

2. Numerical Method

2.1. Governing Equations

The flow is governed by the two-dimensional, incompressible Reynolds-averaged Navier–Stokes and continuity equations:
u i x i = 0
u i t + u j u i x j = 1 ρ p x i + x j 2 υ S i j u i u j ¯
where xi represents the coordinate component, and where i = 1 and 2 represent the x-component and y-component, respectively. ui represents the velocity component corresponding to the coordinate xi, where u1 = u and u2 = v. t is time, ρ is fluid density, p is pressure, and υ is fluid kinematic viscosity. Sij is the strain rate tensor, defined as Sij = (∂ui/∂xj + ∂uj/∂xi)/2. The Reynolds stress u i u j ¯ is defined as follows:
u i u j ¯ = υ t u i / x j + u j / x i + 2 3 k δ i j
where k represents turbulent kinetic energy, and δij represents the Kronecker operator.
Due to the unknown Reynolds stress term u i u j ¯ in the Navier–Stokes equation, a reasonable turbulence model needs to be further introduced for the closed-form solution of the equation. The SST k-ω two-equation turbulence model [20,21] is adopted in the present study: the standard k-ω model is used near the wall, and it is transformed into the standard k-ε model in the free stream and core region, which can accurately predict boundary layer flows with adverse pressure gradients. The governing equations of the SST k-ω two-equation turbulence model are as follows:
k t + u j k x j = x j υ + σ k υ t k x j + P k β ω k
ω t + u j ω x j = x j υ + σ ω υ t ω x j + α p k υ t β ω 2 + 2 1 F 1 σ ω 2 1 ω k x j ω x j
where k represents the turbulent kinetic energy, ω represents the turbulent energy dissipation rate, Pk is the product of turbulent kinetic energy, υt is the turbulent eddy viscosity, and F1 is the mixing function, which are expressed as follows:
P k = min υ t u i x j u i x j + u j x i , 10 β k ω
υ t = α 1 k max α 1 ω , S F 2
F 1 = tanh min min max k β ω y + , 500 υ y + 2 ω , 4 ρ σ ω 2 k D k ω y + 2 , 10 4
where S = (2SijSij)1/2 represents the invariant measure of strain rate, y+ denotes the non-dimensional distance of the grid closest to the wall, and the variables F2 and D are expressed as follows:
F 2 = tanh min max 2 k β ω y + , 500 υ y + 2 ω , 100 2
D k ω = max 2 ρ σ ω 2 1 ω k x j ω x j , 10 10
This is based on the mixed function F1, σk = F1σk1 + (1 − F1)σk2, σω = F1σω1 + (1 − F1)σω2, α = F1α1 + (1 − F1)α2, and β = F1β1 + (1 − F1)β2. In addition, the basic theoretical parameters of the SST k-ω turbulence model used in Formulas (4)–(10) are shown in Table 1.

2.2. Boundary Conditions

Figure 1 shows the schematic of the 2D numerical model for simulating the hydrodynamics of a suspended cable under combined wave–current loading. The coordinate origin O is located on the seabed, directly below the cable center. To minimize boundary effects on the near-cable flow, the inlet (left) and outlet (right) of the fluid are both set at a distance of 25D from the center of the cable. The center of the cable is located at a height of e + D/2 above the seabed. The specific boundary conditions are set as follows: (1) the left boundary is the inflow boundary, with boundary conditions of u = Uc + Umsin(2πt/T), v = 0, ∂p/∂x = (2πUm/T)cos(2πt/T), ∂k/∂n = 0, and ∂ω/∂n = 0; (2) the right boundary is the outflow boundary, with boundary conditions of ∂u/∂x = 0, ∂v/∂y = 0, p = 0, ∂k/∂n = 0, and ∂ω/∂n = 0; (3) the upper boundary is the symmetric boundary condition, with ∂u/∂n = 0, v = 0, ∂p/∂y = 0, ∂k/∂n = 0, and ∂ω/∂n = 0; and (4) both the lower boundary and the cable surface are non-slip boundary conditions, with boundary conditions of u = 0, v = 0, ∂p/∂n = 0, and ω = 6υ/0.075Δ2. The boundary conditions for turbulent kinetic energy k at the lower boundary and cable surface are ∂k/∂n = 0 and k = 0, respectively. Here, n represents the unit normal vector, and Δ represents the first-layer height of the grid.

3. Validation of Numerical Model

3.1. Grid Convergence Verification of Numerical Model

Prior to performing numerical simulations, grid convergence verification was first conducted. The grids used for numerical calculations in this paper are all structured grids, as shown in Figure 2. Local densification of the grid around the cable was achieved mainly through the following two methods: adjusting the circumferential nodes count and the first-layer height of the grid on the cable surface (An et al. [22]). Next, convergence studies were performed with respect to the circumferential node density and the first-layer grid height. The specific information of the four sets of grids is shown in Table 2. The calculation conditions were set as follows: Reynolds number Re = 0.8 × 105, KC = 10, current wave ratio Uc/Um = 0, and suspension ratio e/D = 0.5.
Based on the numerical results using the four different sets of grids, the horizontal forces (the inertia coefficient CM and drag coefficient Cd) can be approximated using the Morison equation (Equation (11)). By fitting using the least squares method, the optimal solutions for Cd and CM can be obtained, as shown in Table 2:
F D t = 1 2 ρ D C d u ( t ) u ( t ) + ρ π D 2 4 C M d u ( t ) d t
Table 2 and Table 3 present the results for oscillatory flow and steady current, respectively. Comparing the results corresponding to the grids G4 and G3 in Table 2, the force coefficients show a difference below 1%, indicating that the grid G3 meets the convergence requirements. As shown in Table 3, the results for mesh G7 (with 140 nodes) and G8 (with 160 nodes) show an insignificant difference. Therefore, the mesh that features a first-layer height of 0.001 D and 160 nodes on the cable surface was selected to ensure accuracy while maintaining computational efficiency. Figure 3 shows the variation in the force coefficients with dimensionless time. The results for Δt* = 0.005 (Δt* = u0Δt/D) are in close agreement with those for a finer time step of Δt* = 0.002, with a negligible difference. Consequently, a time step of Δt* = 0.005 was adopted for all subsequent simulations.

3.2. Accuracy Verification of Numerical Model

This section conducted the accuracy verification on the hydrodynamic characteristics of suspended cables. The computational condition is adopted from Oner et al. [23], i.e., Reynolds number Re = 9500 and suspension ratio e/D = 0.2 and 0.3. Figure 4 compares the numerically simulated and experimental [23] distributions of the average horizontal velocity around the cable, showing good agreement between the present results (red line) and the data from Oner et al. [23] (black circles). The model shows its capability to faithfully capture the velocity distribution around the suspended cable.
The model is subsequently validated against the hydrodynamics of a cylinder in oscillatory flow using the experimental data of Sumer et al. [24], i.e., (a) KC = 4, Re = 0.4 × 105, e/D = 0.1, (b) KC = 10, Re = 0.8 × 105, e/D = 0.1, and (c) KC = 10, Re = 0.8 × 105, e/D = 1.0. As shown in Figure 5, the calculated Cd and CM values are in good agreement with the results of Sumer et al. [24], demonstrating the model’s accuracy for simulating the hydrodynamic forces subjected to cylinders under oscillating flow. Although the flow around a circular cylinder exhibits inherent three-dimensional characteristics, the two-dimensional assumption remains a widely adopted and practical approach for estimating hydrodynamic forces acting on the subsea structures. The reliability of the 2D model is supported by the present study and prior studies [10,25], which demonstrate that it is capable of accurately capturing the dominant characteristics of drag and inertia forces. These force coefficients could be used for on-bottom stability design and buckling analysis. It is noteworthy that the present results are typically workable for a single, endless cross-section, where the effects of span flexibility, the 3D vortex shedding, and the contact at span ends are not considered.

4. Numerical Results and Analysis

This paper conducts research on the hydrodynamic characteristics of underwater suspended cables under the combined wave–current conditions. To simulate the combined effect of wave and current, the oscillatory flow was superimposed onto the steady unidirectional flow, where the oscillatory flow was prescribed by a sine function. The definition of the flow velocity u(t) is as follows:
u t = U c + U m sin 2 π t / T
Under the combined wave–current conditions, the current-to-wave ratio Uc/Um and KC number are the two key parameters. The operating conditions in this article are set as follows: Reynolds number Re = 0.8 × 105, KC number KC = 4, 7, 10, suspension ratio e/D = 0.1–2.0, and the ratio of unidirectional flow velocity to oscillating flow amplitude Uc/Um = 0.2, 0.5, 0.8, 1.0, and 1.5, totaling 120 calculation conditions. The drag coefficient Cd and inertia coefficient CM are still solved by fitting the Morrison equation (Equation (11)) using the least squares method.

4.1. Flow Field Analysis

Figure 6, Figure 7 and Figure 8 show the evolution of vorticity contours over one oscillatory cycle for different current-to-oscillatory ratios and suspension ratios. When the oscillatory flow dominates, the vortex oscillates around the cable and eventually decays. In contrast, as the current intensifies, the vortex tends to propagate downstream. Under the effect of oscillatory flow, vortex shedding persists even at small clearance ratios and is not suppressed by the wall. A comparison of Figure 7 and Figure 8 reveals that the bottom wall significantly influences the shed vortex path. At a small gap ratio, the vortex is deflected upward, whereas at a larger ratio, it propagates horizontally downstream.

4.2. The Influence of Suspension Ratio e/D on Force Coefficients

Figure 9 shows the relationship between the drag coefficient Cd and the suspension ratio e/D for a suspended cable under combined wave–current conditions, where (a)–(c) correspond to KC numbers of 4, 7, and 10, respectively. The results show that under the conditions of three KC numbers, when the current-to-wave ratio Uc/Um ≤ 0.5, the drag coefficient Cd decreases with the suspension ratio, with its influence diminishing markedly for Uc/Um > 0.5. Across all tested conditions, the effect of suspension ratio on the drag coefficient Cd exhibits a marked reduction for e/D values exceeding 0.5. Taking the results at KC = 4 shown in Figure 9a as an example, it is found that when the component of current-induced velocity is relatively weak, say Uc/Um = 0.2, the drag coefficient Cd decreases noticeably with the increasing suspension ratio e/D. This behavior is similar to that found under pure wave conditions, namely Uc/Um = 0, as reported by Teng et al. [26], indicating that the force amplification is due to the effect of wall proximity. However, when the current-induced velocity is strong, say Uc/Um = 1, Cd is less sensitive to the increase in e/D. This phenomenon agrees well with that found under pure current conditions, as reported by Teng et al. [27], indicating that the force is dominated by the velocity distribution in the boundary layer rather than the effect of wall proximity.
Figure 10 presents the variation in the inertia force coefficient CM with the suspension ratio e/D for KC numbers of 4, 7, and 10. The results indicate that the inertia force coefficient CM tends to decrease with e/D. However, when KC = 10 and Uc/Um ≥ 1, the dependence of CM on e/D is weak. Across all tested conditions, the influence of suspension ratio on the inertia force coefficient CM exhibits a marked reduction for e/D values exceeding 0.5.

4.3. The Influence of Current-to-Wave Ratio Uc/Um on Force Coefficients

The drag and inertia coefficients under different current-to-wave ratios Uc/Um are shown in Figure 11 and Figure 12, respectively. The results indicate that when Uc/Um < 1, the variation trend of Cd with Uc/Um is different under the same KC number but different e/D conditions, and it is also different under the same suspension ratio but different KC number conditions. However, under all KC numbers and e/D conditions, CM shows a monotonic decreasing trend with the increase in Uc/Um, as the flow becomes increasingly dominated by the steady current component, which contributes less to the inertial force. For Uc/Um greater than 1.0, both Cd and CM remain nearly constant with a further increase in Uc/Um. The above results indicate that predicting the horizontal force on suspended cables under combined wave–current conditions, especially when Uc/Um < 1, requires establishing quantitative relationships for the force coefficients Cd, CM with KC number, current-to-wave ratio Uc/Um, and the suspension ratio e/D. Taking the results at KC = 7, shown in Figure 11b, as an example, opposite variation trends with the increasing Uc/Um are found in the regions of Uc/Um < 0.5 and Uc/Um > 0.5 when e/D ≤ 0.3. These opposite trends reflect the complicated effect of the co-existence of wave and current. However, the physics responsible for the local minimum drag force at Uc/Um = 0.5 is not well understood.

4.4. Prediction Method of Horizontal Force on Suspended Cables Under Combined Wave–Current Conditions

The force acting on a suspended cable under combined wave–current conditions cannot be considered as merely a superposition of the forces induced by unidirectional flow and oscillatory flow alone. As mentioned earlier, the force coefficients of the cable are governed by three key parameters (i.e., KC number, flow wave ratio Uc/Um, and suspension ratio e/D), and the non-trivial relationships between these parameters make it impractical to define the quantitative influence on the Morrison hydrodynamic coefficient using simple empirical expressions. Therefore, based on a two-layer feedforward neural network model, as shown in Figure 13, the quantitative relationships between the force coefficients Cd, CM, and {KC, Uc/Um, e/D} are established, where the hidden layer contains 10 sigmoid neurons and the output layer contains one linear neuron, where ‘w’ and ‘b’ represent weight and bias, respectively. The Levenberg–Marquardt algorithm is used to train the neural network. The hyperparameters were determined by trials, and those that led to the minimum rms deviation were chosen. Out of the 120 simulation results in this article, 84 were used for training, 18 for validation, and 18 for testing. The root mean square deviations of Cd and CM for model training are less than 1% and 2%, respectively. Note that the applicable range of this model is as follows: 4 ≤ KC ≤ 10, 0 ≤ Uc/Um ≤ 1.5, 0.1 ≤ e/D ≤ 2.
Figure 14 shows the drag and inertia coefficients under the conditions of KC = 7, Uc/Um = 0.5, and e/D = 0.1~2. For ease of comparison, the figure also includes results from simulations with KC = 7 and Uc/Um = 0.5. It should be noted that these simulations have been excluded from any phase of model training or validation. As shown in the figure, the prediction results based on the model proposed in this paper are generally in good agreement with the simulations. The deviation of Cd is relatively larger compared to that of CM, because the variation trend of the training data Cd is more complicated under the effect of the co-existing wave and current.

5. Conclusions

This study investigates the hydrodynamic characteristics of underwater suspended cables under combined wave–current conditions. A two-dimensional viscous flow calculation and analysis model is established by solving the RANS equations with the SST k-ω turbulence model. The influences of the suspension ratio e/D and the current-to-wave ratio Uc/Um on the Morrison coefficient of the cable are analyzed within the parameter ranges of 4 ≤ KC ≤ 10, 0 ≤ Uc/Um ≤ 1.5, and 0.1 ≤ e/D ≤ 2. The low KC values working together with a wide range of Uc/Um in the present study reflect steady currents in co-existence with mild waves. The primary findings are summarized below:
(1)
For the cable suspension ratio e/D of less than 0.5, the strength of the dependence of both the drag and inertia coefficients on the cable suspension ratio e/D is significantly influenced by the current–wave ratio Uc/Um, while this dependence becomes less pronounced for e/D that is greater than 0.5.
(2)
When Uc/Um < 1, the variation trend of Cd with Uc/Um is different under the same KC number but different e/D conditions, and the variation trend of Cd with Uc/Um is also different under the same suspension ratio but different KC number conditions. However, under all KC numbers and e/D conditions, CM shows a monotonically decreasing trend with the increase in Uc/Um. When Uc/Um is greater than 1.0, both Cd and CM almost no longer change with Uc/Um variation.
(3)
Based on the simulation results in this article, a double-layer feedforward neural network model was used to establish quantitative relationships between Cd, CM, and {KC, Uc/Um, e/D}, with root mean square deviations of less than 1% and 2%, respectively.

Author Contributions

Conceptualization, D.Z. and X.H.; methodology, D.Z. and Z.C.; software, J.R. and G.T.; validation, X.H., J.R. and G.T.; formal analysis, D.Z.; investigation, D.Z. and X.H.; resources, D.Z. and Z.C.; data curation, B.Z. and G.T.; writing—original draft preparation, X.H. and J.R.; writing—review and editing, B.Z. and G.T.; visualization, X.H. and B.Z.; supervision, J.R.; project administration, J.R.; funding acquisition, Z.C. 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, grant number 52371262.

Data Availability Statement

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

Conflicts of Interest

Authors Deping Zhao, Xiaowei Huang, Zhenjin Cen were employed by the Extra High Voltage Transmission Company of China Southern Power Grid Co., Ltd. The remaining authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

References

  1. Zhu, J.; Ren, B.; Dong, P.; Cheng, W. Vortex-induced vibrations of a free spanning submarine power cable. Ocean Eng. 2023, 272, 113792. [Google Scholar] [CrossRef]
  2. Zhang, J.; Chen, Y.; Jia, F.; Zhou, Y.; Li, H.; Li, C.; Huang, X.; Guo, Q. Damage Analysis of Suspended Submarine Cable under Current Force Based on Finite Element Simulation. In Proceedings of the 2021 International Conference on Sensing, Measurement & Data Analytics in the Ara of Artificial Intelligence (ICSMD), Nanjing, China, 21–23 October 2021. [Google Scholar]
  3. Liang, D.F.; Cheng, L.; Li, F.J. Numerical modeling of flow and scour below a pipeline in currents: Part Ⅱ. Scour simulation. Coast. Eng. 2005, 52, 43–62. [Google Scholar] [CrossRef]
  4. Cheng, L.; Yeow, K.; Zhang, Z.; Teng, B. Three-dimensional scour below offshore pipelines in steady currents. Coast. Eng. 2009, 56, 577–590. [Google Scholar] [CrossRef]
  5. Bearman, P.W.; Zdravkovich, M.M. Flow around a circular cylinder near a plane boundary. J. Fluid Mech. 1978, 89, 33–47. [Google Scholar] [CrossRef]
  6. Lei, C.W.; Cheng, L.; Armfield, S.W.; Kavanagh, K. Vortex shedding suppression for flow over a circular cylinder near a plane boundary. Ocean Eng. 2000, 27, 1109–1127. [Google Scholar] [CrossRef]
  7. Ong, M.C.; Utnes, T.; Holmedal, L.E.; Myrhaug, D.; Pettersen, B. Numerical simulation of flow around a circular cylinder close to a flat seabed at high Reynolds numbers using a k-ε model. Coast. Eng. 2010, 57, 931–947. [Google Scholar] [CrossRef]
  8. Lei, C.; Cheng, L.; Kavanagh, K. Re-examination of the effect of a plane boundary on force and vortex shedding of a circular cylinder. J. Wind Eng. Ind. Aerodyn. 1999, 80, 263–286. [Google Scholar] [CrossRef]
  9. Buresti, G.; Lanciotti, A. Mean and fluctuating forces on a circular cylinder in cross-flow near a plane surface. J. Wind Eng. Ind. Aerodyn. 1992, 41, 639–650. [Google Scholar] [CrossRef]
  10. An, H.W.; Cheng, L.; Zhao, M. Two-dimensional and three-dimensional simulations of oscillatory flow around a circular cylinder. Ocean Eng. 2015, 109, 270–286. [Google Scholar] [CrossRef]
  11. Munir, A.; Zhao, M.; Wu, H.; Ning, D.; Lu, L. Numerical investigation of the effect of plane boundary on two-degree-of-freedom of vortex-induced vibration of a circular cylinder in oscillatory flow. Ocean Eng. 2018, 148, 17–32. [Google Scholar] [CrossRef]
  12. Tong, F.F.; Cheng, L.; An, H.W.; Jiang, H.F. Modes of synchronisation for oscillatory cylinder in steady flow at various angles of attack. In Proceedings of the 21st Australasian Fluid Mechanics Conference, Adelaide, Australia, 10–13 December 2018. [Google Scholar]
  13. Teng, B.; Mao, H.F.; Ning, D.Z.; Zhang, C.W. Viscous numerical examination of hydrodynamic forces on a submerged horizontal circular cylinder undergoing forced oscillation. J. Hydrodyn. 2019, 31, 887–899. [Google Scholar] [CrossRef]
  14. Zhao, M.; Cheng, L.; An, H. Three-dimensional numerical simulation of flow around a circular cylinder under combined steady and oscillatory flow. J. Hydrodyn. 2010, 22, 144–149. [Google Scholar] [CrossRef]
  15. Cao, S.; Li, M. Numerical study of flow over a circular cylinder in oscillatory flows with zero-mean and non-zero-mean velocities. J. Wind Eng. Ind. Aerodyn. 2015, 144, 42–52. [Google Scholar] [CrossRef]
  16. Deng, Y.; Huang, W.; Zhao, J. Combined action of uniform flow and oscillating flow around marine riser at low keulegan-carpenter number. J. Ocean Univ. China 2014, 13, 390–396. [Google Scholar] [CrossRef]
  17. Zhao, M.; Kaja, K.; Xiang, Y.; Yan, G. Vortex-induced vibration (VIV) of a circular cylinder in combined steady and oscillatory flow. Ocean Eng. 2013, 73, 83–95. [Google Scholar] [CrossRef]
  18. Chang, W.Y.; Constantinescu, G. Flow structure around a vertical cylinder placed in an open channel under combined wave-current flows. Phys. Rev. Fluids 2025, 10, 024804. [Google Scholar] [CrossRef]
  19. Teng, Y.; Griffiths, T.; An, H.; Draper, S.; Cheng, L.; Mohr, H.; Zhang, C.; Tang, G.; Lu, L. Hydrodynamic forces on small-diameter pipelines in combined wave-and-current flows. Coast. Eng. 2025, 202, 104852. [Google Scholar] [CrossRef]
  20. Menter, F.R. Two-equation eddy-viscosity turbulence models for engineering applications. AIAA J. 1994, 32, 1598–1605. [Google Scholar] [CrossRef]
  21. Menter, F.R.; Kuntz, M.; Langtry, R. Ten years of industrial experience with the SST turbulence model. Turbul. Heat Mass Transf. 2003, 4, 625–632. [Google Scholar]
  22. An, H.; Cheng, L.; Zhao, M. Steady streaming around a circular cylinder near a plane boundary due to oscillatory flow. J. Hydraul. Eng.-ASCE 2011, 137, 23–33. [Google Scholar] [CrossRef]
  23. Oner, A.A.; Kirkgoz, M.S.; Akoz, M.S. Interaction of a current with a circular cylinder near a rigid bed. Ocean Eng. 2008, 35, 1492–1504. [Google Scholar] [CrossRef]
  24. Sumer, B.M.; Jensen, B.L.; Fredsøe, J. Effect of a plane boundary on oscillatory flow around a circular-cylinder. J. Fluid Mech. 1991, 225, 271–300. [Google Scholar] [CrossRef]
  25. Liang, D.F.; Cheng, L. Numerical modeling of flow and scour below a pipeline in currents Part I. Flow simulation. Coast. Eng. 2005, 52, 25–42. [Google Scholar] [CrossRef]
  26. Teng, Y.; Griffiths, T.; Tang, G.; An, H.; Draper, S.; Cheng, L.; Mohr, H. Peak force coefficients on small-diameter spanning pipelines under waves. Coast. Eng. 2022, 177, 104189. [Google Scholar] [CrossRef]
  27. Teng, Y.; Griffiths, T.; Cheng, L.; Qu, Y.; An, H.; Draper, S.; Gao, Z.; Tang, G.; Lu, L. Effect of steady boundary-layer flow on the hydrodynamic forces on a near-wall cylinder in currents. Ocean Eng. 2024, 313, 119338. [Google Scholar] [CrossRef]
Figure 1. Schematic diagram of the computational domain.
Figure 1. Schematic diagram of the computational domain.
Jmse 13 02067 g001
Figure 2. Near-field mesh around the cable: (a) General view of the cable; (b) detailed view of the cable.
Figure 2. Near-field mesh around the cable: (a) General view of the cable; (b) detailed view of the cable.
Jmse 13 02067 g002
Figure 3. Time step convergence verification results: (a) Drag coefficient (CD); (b) lift coefficient (CL).
Figure 3. Time step convergence verification results: (a) Drag coefficient (CD); (b) lift coefficient (CL).
Jmse 13 02067 g003
Figure 4. Average horizontal velocity distribution around the cable: (a) e/D = 0.2; (b) e/D = 0.3.
Figure 4. Average horizontal velocity distribution around the cable: (a) e/D = 0.2; (b) e/D = 0.3.
Jmse 13 02067 g004
Figure 5. Comparisons of drag and inertia coefficients with experimental results (Sumer et al. [24]).
Figure 5. Comparisons of drag and inertia coefficients with experimental results (Sumer et al. [24]).
Jmse 13 02067 g005
Figure 6. Flow field around the cable at different phases (e/D = 0.1, Uc/Um = 0.2, KC = 7).
Figure 6. Flow field around the cable at different phases (e/D = 0.1, Uc/Um = 0.2, KC = 7).
Jmse 13 02067 g006
Figure 7. Flow field around the cable at different phases (e/D = 0.1, Uc/Um = 1, KC = 7).
Figure 7. Flow field around the cable at different phases (e/D = 0.1, Uc/Um = 1, KC = 7).
Jmse 13 02067 g007
Figure 8. Flow field around the cable at different phases (e/D = 0.7, Uc/Um = 1, KC = 7).
Figure 8. Flow field around the cable at different phases (e/D = 0.7, Uc/Um = 1, KC = 7).
Jmse 13 02067 g008
Figure 9. Drag coefficient Cd against gap ratio e/D at various current-to-wave ratios.
Figure 9. Drag coefficient Cd against gap ratio e/D at various current-to-wave ratios.
Jmse 13 02067 g009aJmse 13 02067 g009b
Figure 10. Inertia coefficient CM against gap ratio e/D at various current-to-wave ratios.
Figure 10. Inertia coefficient CM against gap ratio e/D at various current-to-wave ratios.
Jmse 13 02067 g010aJmse 13 02067 g010b
Figure 11. Drag coefficient Cd against Uc/Um at different gap ratios.
Figure 11. Drag coefficient Cd against Uc/Um at different gap ratios.
Jmse 13 02067 g011aJmse 13 02067 g011b
Figure 12. Inertia coefficient CM against Uc/Um at different gap ratios.
Figure 12. Inertia coefficient CM against Uc/Um at different gap ratios.
Jmse 13 02067 g012
Figure 13. Sketch of the neural network model.
Figure 13. Sketch of the neural network model.
Jmse 13 02067 g013
Figure 14. Predictions of Cd and CM based on the present neural network model: (a) Cd; (b) CM.
Figure 14. Predictions of Cd and CM based on the present neural network model: (a) Cd; (b) CM.
Jmse 13 02067 g014aJmse 13 02067 g014b
Table 1. Parameters in the SST k-ω turbulence model.
Table 1. Parameters in the SST k-ω turbulence model.
β*α1β1σk1σω1α2β2σk2σω2
0.095/93/400.850.50.440.08281.00.856
Table 2. Grid convergence verification for the height of the first grid layer on the cable.
Table 2. Grid convergence verification for the height of the first grid layer on the cable.
Grid NumberThe Height of the First Grid Layer on the CableCdCM
G10.003 D1.251.60
G20.002 D1.512.05
G30.001 D1.422.07
G40.0005 D1.412.08
Table 3. Grid convergence verification for the number of nodes on the cable surface.
Table 3. Grid convergence verification for the number of nodes on the cable surface.
Grid NumberNumber of Nodes on the Cable SurfaceCDCL
G51201.180.151
G61401.270.124
G71601.280.115
G81801.290.118
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

Zhao, D.; Huang, X.; Cen, Z.; Ren, J.; Zhan, B.; Tang, G. Numerical Investigation on the Hydrodynamic Characteristics of Submarine Power Cables for Offshore Wind Turbines Under Combined Wave–Current Loading. J. Mar. Sci. Eng. 2025, 13, 2067. https://doi.org/10.3390/jmse13112067

AMA Style

Zhao D, Huang X, Cen Z, Ren J, Zhan B, Tang G. Numerical Investigation on the Hydrodynamic Characteristics of Submarine Power Cables for Offshore Wind Turbines Under Combined Wave–Current Loading. Journal of Marine Science and Engineering. 2025; 13(11):2067. https://doi.org/10.3390/jmse13112067

Chicago/Turabian Style

Zhao, Deping, Xiaowei Huang, Zhenjin Cen, Jianfeng Ren, Bolin Zhan, and Guoqiang Tang. 2025. "Numerical Investigation on the Hydrodynamic Characteristics of Submarine Power Cables for Offshore Wind Turbines Under Combined Wave–Current Loading" Journal of Marine Science and Engineering 13, no. 11: 2067. https://doi.org/10.3390/jmse13112067

APA Style

Zhao, D., Huang, X., Cen, Z., Ren, J., Zhan, B., & Tang, G. (2025). Numerical Investigation on the Hydrodynamic Characteristics of Submarine Power Cables for Offshore Wind Turbines Under Combined Wave–Current Loading. Journal of Marine Science and Engineering, 13(11), 2067. https://doi.org/10.3390/jmse13112067

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