Next Article in Journal
Planning and Design Strategies for Green Stormwater Infrastructure from an Urban Design Perspective
Previous Article in Journal
Transformative Trends in Runoff and Sediment Dynamics and Their Influential Drivers in the Wuding River Basin of the Yellow River: A Comprehensive Analysis from 1960 to 2020
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Large Scale Experimental Study on Waves and Submerged Horizontal Cylinders

1
Tianjin Research Institute of Water Transport Engineering, M.O.T., Tianjin 300456, China
2
Tianjin Key Laboratory of Port and Ocean Engineering, Tianjin University, Tianjin 300072, China
*
Authors to whom correspondence should be addressed.
Water 2024, 16(1), 28; https://doi.org/10.3390/w16010028
Submission received: 26 November 2023 / Revised: 13 December 2023 / Accepted: 18 December 2023 / Published: 20 December 2023
(This article belongs to the Section Hydraulics and Hydrodynamics)

Abstract

:
A large-scale physical model experiment was conducted to study the interaction between regular waves and submerged horizontal cylinders, breaking through the bottleneck of a low Reynolds number in a traditional small water tank. The surface pressure distribution and overall force of submerged horizontal cylinders at different submerged depths in waves with different heights and periods under a high Reynolds number were analyzed, and the variation in first- and second-order wave forces with wave height periods was analyzed. Subsequently, the influence of the cylinder on the wave field was analyzed, and the reflection coefficient was obtained through the three-point separation method. According to the analysis of the experimental results, the horizontal and vertical forces acting on the submerged cylinder per linear meter are similar, with the horizontal force slightly greater than the vertical force; at the same time, the submerged cylinder has a certain reflection effect on long-period waves. The summary, analysis, and display of relevant experimental results provide validation support for the current high-precision mathematical model as much as possible.

1. Introduction

Horizontal cylinders are widely used in ocean engineering, and waves are one of the main loads on ocean engineering structures. Under long-term wave action, marine structures are prone to fatigue failure and other problems. Therefore, research on the effect of wave forces on horizontal cylinders has important practical significance and has always been a hot research topic for scholars.
The theoretical research on horizontal cylindrical wave forces is mainly based on potential flow theory. Dean [1] proposed an analytical solution for linearly solving regular waves and submerged horizontal cylinders and found that the reflection coefficient is 0 in deep water. Ursell [2] subsequently confirmed this conclusion using the multipole expansion method. Ogilvie et al. [3] applied Ursell’s method [2] to further calculate the second-order wave force on a submerged cylinder and found that in deep water, the first-order horizontal wave force is equal to the first-order vertical wave force, and the second-order average drift force is zero. However, theoretical analysis usually simplifies the structural form and adopts the assumption of waves with small amplitude, which has certain limitations for strong nonlinear large amplitude sea conditions.
In recent years, with the development of numerical simulation, nonlinear potential flow wave tanks have been used to study the effect of high-order wave forces on submerged cylinders [4,5,6]. Jin et al. [7] compared and analyzed the hydrodynamic characteristics of submerged structures with different shapes using high-order boundary element methods. Yan et al. [8] established a completely nonlinear wave tank and studied the interaction between strongly nonlinear waves and submerged cylinders. They analyzed the changes in high-order free harmonics and locked-in waves behind the cylinder with factors such as incident wave nonlinearity, cylinder size, and submerged water depth. Although numerical calculations based on potential flow theory have been widely applied, they may overestimate the first harmonic wave force under the action of large amplitude waves due to the neglect of fluid viscosity [9]. Numerical calculations based on viscous flow theory can to some extent more realistically reproduce the physical process of the interaction between a horizontal cylinder and waves. The numerical methods used include the time-varying Reynolds equation [10], the Reynolds average equation [11,12,13,14], and the smooth particle method [15], which promote the understanding of the development process of vortices around the cylinder and the related energy dissipation mechanism, however, the simulation results of viscous flow are extremely sensitive to model settings (such as grid quality and turbulence model), and physical model measurement results need to be used in advance for calibration and verification. Gao et al. [16,17,18] applied the Computational Fluid Dynamic (CFD) method to simulated the wave–structure interaction to investigated the subtle changes in the flow field near the structure. Although some model experiments have been conducted, they were mainly based on traditional small wave tanks [14,19,20,21,22,23,24]. Due to experimental conditions, the model scale and the Reynolds number are both small. The accuracy of the simulated high Reynolds number operating conditions using the model validated by small Reynolds number experiments still needs further verification.
In the design of ocean engineering, the Morrison equation is used to calculate the wave force on the cylinder, and it is necessary to determine the drag force coefficient Cd and inertia force coefficient Cm in the equation. Li [25] obtained relevant results for Cd, Cm, and KC numbers under various environmental conditions based on experiments. Chen and Zhu [26] fitted experimental and empirical data to the variation of the inertial force coefficient and resistance coefficient of a cylinder with a Reynolds number and Churchill number, and obtained the relevant coefficients. Hu et al. [27] used an experimental method of forced oscillation of a horizontal cylinder in a towing tank to study the hydrodynamic characteristics of a horizontal cylinder with different immersion depths under flow, wave, and wave current coupling. They obtained the relationships between the drag force coefficient and the added mass coefficient with a Reynolds number, KC number, and Vr at different immersion depths. It should be emphasized that these experiments were still carried out in traditional small wave tanks, and currently there are relatively few large-scale wave tank experiments in the corresponding research.
For this purpose, this paper conducted a series of large-scale physical model tests on waves and submerged cylinders. Since the cylinder itself has a larger dimension, and there will be obvious diffraction phenomena after waves pass through, pressure sensors are placed on the surface of the cylinder to measure the dynamic force on the cylinder, and wave probes are arranged in front of the cylinder to measure the reflection of waves passing through the cylinder. The analysis of the pressure distribution, wave force on the submerged cylinder, as well as the reflection of waves by the large cylinder provided support for the verification of high-precision mathematical models and the calculation of horizontal cylinder wave forces at present.

2. Physical Model Tests

The experiment was conducted in a large-scale wave tank at the Tianjin Research Institute of Water Transport Engineering, Ministry of Transport (TIWTE). The large wave tank is 456 m long, 5 m wide, and 8 m high. The model is located about 230 m away from the wave maker. To display the meanings of parameters in this paper more clearly, Table 1 introduces the meanings of the relevant symbols. The specific layout is shown in Figure 1, and the photo of the model after installation is shown in Figure 2. The diameter of the submerged cylinder is 1.0 m, considering two submerged depths. Three wave probes are arranged in front of the submerged cylinder to measure the reflection coefficient of the waves caused by the cylinder. When waves acted on submerged cylinders, the contribution of viscous shear force to the total wave force was very small [9], therefore, 16 pressure sensors were arranged at equal angles along the circumference of the cylinder surface to measure the pressure time history curve of each point, and then multiplied by their respective force area and azimuth to obtain the horizontal and vertical forces of the cylinder. The distribution of pressure sensors is displayed in Figure 3, and the expressions of the wave force on the cylinder are shown below.
F x = i = 1 16 p i s i cos θ i
F y = i = 1 16 p i s i sin θ i
The experimental groups are shown in Table 2, and all wave elements of the groups are within the range of second-order Stokes waves.

3. Test Results and Analysis

3.1. Dynamic Pressure on the Cylinder Surface

When the wave period is small, the pressure on the surface of the cylinder shows that the dynamic pressure on the upper part of the cylinder is higher, and the dynamic pressure on the lower part is lower. Figure 4 shows the maximum dynamic pressure on the cylindrical surface when the relative submergence depth ds/d is 0.25 and the wave period is 2.25 s. It can be observed that different wave heights exhibit the same conclusion. As the wave period increases and the wave height increases, the reflection effect of the cylinder on the wave field increases, and the maximum pressure position of the cylinder changes to the area of 45° from the vertical axis of the cylinder on the up-wave and back-wave sides. There are two reasons: (1) the trajectory of water quality points in short-period waves is circular, and the radius of motion decays quickly with an increase in water depth. However, the trajectory of water quality points is elliptical when long-period waves act on the cylinder, and the horizontal motion amplitude of water quality points is greater than the vertical motion amplitude. Therefore, the maximum pressure position of the cylinder is not at the top of the cylinder, and the movement of water quality points decreases relatively slowly with water depth; (2) the reason for the large wave front measurement is the reflection effect of the cylinder on the waves, while the reason for the large wave back side is that the waves break through the cylinder, causing bubbles to impact the surface of the cylinder. Figure 5 shows the maximum dynamic pressure on the cylindrical surface when the same relative submergence and the wave period is 3.11 s. It was found that as the wave height increased, the maximum pressure position shifted to both sides. Figure 6 shows the maximum dynamic pressure on the cylindrical surface when the wave period is 3.98 s. Due to the longer wave period and stronger reflection, the above phenomenon is more obvious.
When the depth of cylinder submergence increases to ds/d = 0.50, the pressure distribution trend remains the same. Figure 7 shows the maximum dynamic pressure on the surface of a cylinder when the wave period is 2.77 s. Even if the wave height increases, it still shows the maximum dynamic pressure on the upper part and the minimum dynamic pressure on the lower part of the cylinder. Compared with the case of ds/d = 0.25, when the wave height is large, due to the reflection on the wave field, the pressure at 45° is already higher than that in the middle, as shown in Figure 8. Figure 9 shows the dynamic pressure distribution under the action of long-period waves. It was found that under the action of long-period waves, due to the influence of a cylinder on the wave field being obvious, the position of the dynamic maximum pressure shifted to both sides in large wave heights. Figure 10 shows the experimental photos of interaction between the waves and the submerged cylinder.

3.2. Wave Force Analysis of Submerged Cylinder

Wave forces in horizontal and vertical directions are calculated firstly and the spectrum results are obtained by the FFT method. Figure 11 and Figure 12 show the wave force measurement and analysis results when H = 0.78 m, T = 4.51 s, ds/d = 0.50. The first- and second-order components can be seen in the spectrum analysis results obviously, corresponding to the wave frequency and twice wave frequency, which are defined as first-order and second-order wave force. These two parts have an important role on the ocean engineering. The first-order wave force dominates the force situation of the structure, while the second-order wave force, due to its high frequency, may approach the natural frequency of the ocean engineering and cause the resonance failure of the structure.
Figure 13 shows the variation of the first- and second-order wave force with wave height, taking the relative submergence depth ds/d = 0.25 and the wave periods of 2.51 s and 2.77 s as examples. From the experimental results, the horizontal and vertical forces acting on the submerged cylinder per meter are similar, with the horizontal force being slightly greater than the vertical force. The first-order force on the structure is directly proportional to the wave height, while the second-order force is directly proportional to the square of the wave height. Figure 14 shows the experimental results of all groups at a submerged depth of ds/d = 0.25, indicating that the wave force increases with an increase in wave height. However, as the wave height continues to increase, the impact of the cylinder on the wave field becomes significant, and the waves will break after passing through the submerged cylinder, resulting in a smaller increase in structural force.
Figure 15 and Figure 16 demonstrate that the dimensionless first- and second-order wave force varies with the relative cylinder diameter, respectively. The first-order dimensionless wave force increases with an increase in D/L, indicating that the submerged cylinder has a smaller impact on short waves. However, for long waves, due to the larger size of the structure, it can cause wave breaking, and some energy is dissipated due to wave breaking. Therefore, the structure is subjected to less wave force, and the higher the wave height, the smaller the dimensionless wave force. The second-order wave force first increases and then decreases with an increase in D/L, and the reason for the increase part is the same as the first-order wave force. The main reason for the decrease is that the wave wavelength is small at this time, and the structure is less affected by surface waves, which in turn affects the results of the second-order wave force. Overall, when D/L is approximately 0.1, the structure is most significantly affected by the second-order wave force.
Figure 17 shows the force results of the cylinder with relative submergence depth ds/d = 0.50 and wave period T = 2.25 s and 2.77 s. From the calculation results, the horizontal and vertical forces acting on the submerged cylinder per linear meter are similar, with the horizontal force slightly greater than the vertical force. Due to the deep submergence depth of the cylinder at this time, the structure is subjected to relatively small second-order wave forces, which differ by two orders of magnitude from the first-order wave forces and can be disregarded.
The results of all experimental groups with a relative submergence depth of ds/d = 0.50 for the cylinder are shown in Figure 18. The wave force increases with an increase in wave height. Due to the deep submergence depth of the cylinder, the wave field propagation effect of the submerged cylinder is very small, and even in the case of long-period large waves, there is still no wave breaking phenomenon. Therefore, the force on the cylinder is almost proportional to the wave height.
Figure 19 and Figure 20 show the variation of first- and second-order dimensionless wave forces with ds/d = 0.50 as a function of D/L, respectively. For first-order wave forces, there is little change in the dimensionless wave force of the submerged cylinder with an increase in D/L. At the same time, the influence of wave height on the wave force is not significant, indicating that the influence of the submerged cylinder on surface waves is already very weak. Through the comparison of first- and second-order wave forces, the deeper the submerged depth, the smaller the second-order wave force, which can be disregarded in the calculation.

3.3. Morison Formula Inertia Value

Other scholars have conducted research to obtain the inertial force coefficient values of small-scale submerged cylinders under different conditions, but there are certain differences in the values. For example, Li [25], Chen, and Zhu [26] calculated the inertial force coefficient of a horizontal cylinder between 1.2 and 2.0, while Hu et al. [27] and Mao et al. [28] calculated that in extreme cases, the inertial force coefficient of a horizontal cylinder that may be exposed to the water surface may be greater than 2.0. The submerged cylindrical structure developed in this article has a large scale and is mainly controlled by inertial force. The CM values at different Reynolds numbers are obtained by incorporating the Morison equation. The results are shown in Figure 21. It is found that the inertial force coefficients of the cylinder are between 1.2 and 2.0 at each Reynolds number, indicating that the Morison formula is also applicable when calculating large-scale structures. Relevant calculation coefficients provided by Chinese regulations can be referred to in the calculation.

3.4. Influence on the Wave Field of the Cylinder

Based on the data of three wave probes in front of the cylinder, the three-point method of Mansard and Funke [29] is used to calculate the reflected wave surface curves and spectrum.
Assuming that the distances between the three wave probes are X12 and X13, respectively, the incident wave and reflected wave can be derived as follows:
Z I , k = 1 D k B 1 , k R 1 + i Q 1 + B 2 , k R 2 + i Q 2 + B 3 , k R 3 + i Q 3
Z R , k = 1 D k B 1 , k R 1 i Q 1 + B 2 , k R 2 i Q 2 + B 3 , k R 3 i Q 3
where B1, k is the Fourier transform results of wave elevation, and the expressions of other symbols are as below
D k = 2 · sin 2 β k + sin 2 γ k + sin 2 γ k β k R 1 k = sin 2 β k + sin 2 γ k Q 1 k = sin β k · cos β k + sin γ k · cos γ k R 2 k = sin γ k sin γ k β k Q 2 k = sin γ k · cos γ k β k + 2 · sin β k R 3 k = sin β k sin γ k β k Q 3 k = sin β k · cos γ k β k 2 · sin γ k
βk and γk of the above expressions are shown as follows
β k = 2 π X 12 L γ k = 2 π X 13 L
Figure 22 and Figure 23 show the reflected wave surface curves and amplitude spectra under two different wave heights, with a submergence depth ds/d = 0.25 and a wave period of 2.25 s. The results show that the reflected waves are mainly composed of two types of frequency waves, corresponding to the wave frequency and the lateral oscillation frequency of the water tank. We analyzed the amplitude corresponding to the reflected wave frequency and found that under the action of waves in this period, the reflection coefficient of the submerged cylinder is very small, ranging from approximately 0.070 to 0.080, indicating that the submerged cylinder has little impact on the waves of this period.
Subsequently, the wave field under different wave heights with a larger wave period T = 3.45 s was selected for analysis are shown in Figure 24 and Figure 25. From the figures, the wavelength also increases significantly with an increase in wave period, resulting in a significant increase in the depth of water quality point influence. Under the wave action of this period, the reflected waves are more pronounced compared to the short-period situation, with a reflection coefficient of 0.082~0.098.
The same analysis is applied in the cases of the submergence depth ds/d = 0.50. After statistical analysis, the reflection coefficients under different D/L conditions for two different submergence depths are obtained, as shown in Figure 26. From the results, the deeper the submergence depth of the cylinder, the less obvious the reflection effect. When the submergence depth is constant, the submerged cylinder has a certain reflection effect on long-period waves. The main reason is that the action of long-period waves is deeper, which means that the water quality points still exhibit obvious motion characteristics in deeper situations. Therefore, the submerged cylinder has a certain reflection effect on long-period waves.

4. Conclusions

This paper conducted a series of physical model tests under high Reynolds numbers for waves and underwater submerged cylinders in the large-scale wave tank in TIWTE. By analyzing the pressure distribution, wave force, and reflection of waves by the submerged cylinder, it provides validation support for the current high-precision mathematical model as much as possible. The conclusions are as follows:
(1)
When the wave period is small, the dynamic pressure on the surface of the submerged cylinder shows a higher dynamic pressure in the upper part and a lower dynamic pressure in the lower part. The reflection effect of the cylinder on the wave field is enhanced with increases in wave periods and wave heights. Waves would break when passing through the cylinder, and the maximum stress position of the cylinder changed to an area 45° from the vertical axis of the cylinder on the up-wave and back-wave sides.
(2)
The horizontal and vertical forces acting on the submerged cylinder per linear meter are similar, with the horizontal force slightly greater than the vertical force. The first-order force on the structure is directly proportional to the wave height, while the second-order force is directly proportional to the square of the wave height; the overall force on the structure increases with an increase in wave height. However, as the wave height continues to increase, the cylinder has a significant impact on the wave field. After the wave passes through the submerged cylinder, it will break. Through calculation, the inertia force coefficient of the cylinder is between 1.2 and 2.0, indicating that the Morison formula can also be used for calculating the force on large-scale cylinders. The coefficient can be selected according to these specifications.
(3)
The deeper the submergence depth of the submerged cylinder, the less obvious the reflection effect. When the submergence depth is constant, the submerged cylinder has a certain reflection effect on long-period waves. The main reason is that the long-period wave action is deeper in water depth, which means that water quality points still exhibit obvious motion characteristics in deeper situations. Therefore, the submerged cylinder has a certain reflection effect on long-period waves.

Author Contributions

Conceptualization, R.J.; Data curation, X.Z.; Writing—original draft, R.J.; Writing—review & editing, Y.L. and B.G. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Key research and development program (2022YFB2602800), National Natural Science Foundation of China (U2106223, U21A20123), the Basic Funding of the Central Public Research Institutes (TKS20230106).

Data Availability Statement

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

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Dean, W.R. On the reflexion of surface waves by a submerged circular cylinder. Math. Proc. Camb. Philos. Soc. 1948, 4, 483–491. [Google Scholar] [CrossRef]
  2. Ursell, F. Surface waves on deep water in the presence of a submerged circular cylinder, I and II. Math. Proc. Camb. Philos. Soc. 1950, 1, 141–158. [Google Scholar] [CrossRef]
  3. Ogilvie, T.F. First- and second-order forces on a cylinder submerged under a free surface. J. Fluid Mech. 1963, 3, 451–472. [Google Scholar] [CrossRef]
  4. Guerber, E.; Benoit, M.; Grilli, S.T.; Buvat, C. A fully nonlinear implicit model for wave interactions with submerged structures in forced or free motion. Eng. Anal. Bound. Elem. 2012, 7, 1151–1163. [Google Scholar] [CrossRef]
  5. Hannan, M.A.; Bai, W.; Ang, K.K. Modeling of fully nonlinear wave radiation by submerged moving structures using the higher order boundary element method. J. Mar. Sci. Appl. 2014, 1, 1–10. [Google Scholar] [CrossRef]
  6. Bai, W.; Hannan, M.A.; Ang, K.K. Numerical simulation of fully nonlinear wave interaction with submerged structures: Fixed or subjected to constrained motion. J. Fluids Struct. 2014, 49, 534–553. [Google Scholar] [CrossRef]
  7. Jin, R.J.; Liu, Y.; Geng, B.L.; Zhang, H.Q. Hydrodynamic characteristics of circular tube structure with wave action. J. Waterw. Harb. 2019, 3, 249–258. (In Chinese) [Google Scholar]
  8. Yan, B.; Lin, H.X.; Ning, D.Z. Nonlinear numerical simulation of wave action on submerged horizontal cylinder. Adv. Mar. Sci. 2013, 3, 319–325. (In Chinese) [Google Scholar]
  9. Mao, H.F. Numerical Examination of Wave Forces on Horizontal Circular Cylinder Based on Viscous Fluid Theory. Ph.D. Thesis, Dalian University of Technology, Dalian, China, 2018. (In Chinese). [Google Scholar]
  10. Liu, C.G.; Tao, J.H. Solitary waves interaction with horizontal cylinder submerged in different depth. J. Hydrodyn. 2006, 2, 155–160. (In Chinese) [Google Scholar]
  11. Liu, M.M.; Lu, L.; Teng, B.; Zhao, M.; Tang, G.Q. Numerical modeling of local scour and forces for submarine pipeline under surface waves. Coast. Eng. 2016, 116, 275–288. [Google Scholar] [CrossRef]
  12. Teng, B.; Mao, H.F.; Lu, L. Viscous Effects on Wave Forces on A Submerged Horizontal Circular Cylinder. China Ocean Eng. 2018, 3, 245–255. [Google Scholar] [CrossRef]
  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. 2018, 5, 887–899. [Google Scholar] [CrossRef]
  14. Bai, J.L.; Ma, N.; Gu, X.C. Wave-current loads on the horizontal cylinder with varying submergence depth. J. Shanghai Jiaotong Univ. 2018, 8, 938–945. (In Chinese) [Google Scholar]
  15. Moballa, B.; Chern, M.J.; Borthwick, A.G.L. Incompressible SPH simulation of flow past horizontal cylinder between plane wall and free surface. J. Fluids Struct. 2020, 97, 103091. [Google Scholar] [CrossRef]
  16. Gao, J.L.; Zang, J.; Chen, L.F.; Chen, Q.; Ding, H.Y.; Liu, Y.Y. On hydrodynamic characteristics of gap resonance between two fixed bodies in close proximity. Ocean Eng. 2019, 173, 28–44. [Google Scholar] [CrossRef]
  17. Gao, J.L.; He, Z.W.; Zang, J.; Chen, Q.; Ding, H.Y.; Wang, G. Numerical investigations of wave loads on fixed box in front of vertical wall with a narrow gap under wave actions. Ocean Eng. 2020, 206, 107323. [Google Scholar] [CrossRef]
  18. Gao, J.L.; Jing, L.Y.U.J.; Wang, J.H.; Zhang, J.; Liu, Q.; Zang, J.; Zou, T. Study on transient gap resonance with consideration of the motion of floating body. China Ocean Eng. 2022, 36, 994–1006. [Google Scholar] [CrossRef]
  19. Chaplin, J.R. On the irrotational flow around a horizontal cylinder in waves. J. Appl. Mech. 1981, 4, 689–694. [Google Scholar] [CrossRef]
  20. Chaplin, J.R. Mass transport around a horizontal cylinder beneath waves. J. Fluid Mech. 1984, 140, 175–187. [Google Scholar] [CrossRef]
  21. Chaplin, J.R. Non-linear forces on a horizontal cylinder beneath waves. J. Fluid Mech. 1984, 147, 449–464. [Google Scholar] [CrossRef]
  22. Chaplin, J.R. Non-linear wave interactions with a submerged horizontal cylinder. In Proceedings of the ISOPE International Ocean and Polar Engineering Conference, Stavanger, Norway, 17 June 2001; ISOPE: Stavanger, Norway, 2001. [Google Scholar]
  23. Contento, G.; Codiglia, R. Non-linear free surface induced pressure on a submerged horizontal circular cylinder at low Keulegan-Carpenters numbers. Appl. Ocean Res. 2001, 3, 175–185. [Google Scholar] [CrossRef]
  24. Fu, Y.Y. Analytical Solution for Wave Action on Submerged Horizontal Circular Cylinder. Master’s Thesis, Dalian University of Technology, Dalian, China, 2012. (In Chinese). [Google Scholar]
  25. LI, Y.C. Aspect of the normalization of hydrodynamic coefficients in Morison equation. J. Hydrodyn. 1998, 3, 329–337. (In Chinese) [Google Scholar]
  26. Chen, J.; Zhu, K.Q. Fitting of inertial coefficient and drag coefficient curve due to the changes in Reynolds and Keulegan-Carpecter. Ship Sci. Technol. 2014, 2, 40–43. (In Chinese) [Google Scholar]
  27. Hu, K.; Fu, S.X.; Xu, Y.W.; Ma, L.X. Experimental research about hydrodynamic characteristics of horizontal cylinder under different draft. J. Ship Mech. 2017, 10, 1190–1198. (In Chinese) [Google Scholar]
  28. Mao, H.F.; He, Y.L.; Yuan, J.P.; Pan, X.X.; Jia, B.Z. Hydrodynamic coefficients of a horizontal circular cylinder near free surface under wave action. J. Guangdong Ocean Univ. 2020, 4, 109–115. (In Chinese) [Google Scholar]
  29. Mansard, E.P.D.; Funke, E.R. The measurement of incident and reflected spectra using a least squares method. In Proceedings of the 17th International Conference on Coastal Engineering, Sydney, Australia, 23 March 1980; pp. 154–172. [Google Scholar]
Figure 1. Layout diagram of wave and horizontal submerged cylinder.
Figure 1. Layout diagram of wave and horizontal submerged cylinder.
Water 16 00028 g001
Figure 2. Model installation of ds/d = 0.25.
Figure 2. Model installation of ds/d = 0.25.
Water 16 00028 g002
Figure 3. Pressure sensors distribution in the experiments.
Figure 3. Pressure sensors distribution in the experiments.
Water 16 00028 g003
Figure 4. The maximum dynamic pressure distribution on the cylindrical surface under different wave heights at a wave period of 2.25 s (ds/d = 0.25): (a) H = 0.12 m; (b) H = 0.42 m.
Figure 4. The maximum dynamic pressure distribution on the cylindrical surface under different wave heights at a wave period of 2.25 s (ds/d = 0.25): (a) H = 0.12 m; (b) H = 0.42 m.
Water 16 00028 g004
Figure 5. The maximum dynamic pressure distribution on the cylindrical surface under different wave heights at a wave period of 3.11 s (ds/d = 0.25): (a) H = 0.20 m; (b) H = 0.60 m.
Figure 5. The maximum dynamic pressure distribution on the cylindrical surface under different wave heights at a wave period of 3.11 s (ds/d = 0.25): (a) H = 0.20 m; (b) H = 0.60 m.
Water 16 00028 g005
Figure 6. The maximum dynamic pressure distribution on the cylindrical surface under different wave heights at a wave period of 3.98 s (ds/d = 0.25): (a) H = 0.30 m; (b) H = 0.70 m.
Figure 6. The maximum dynamic pressure distribution on the cylindrical surface under different wave heights at a wave period of 3.98 s (ds/d = 0.25): (a) H = 0.30 m; (b) H = 0.70 m.
Water 16 00028 g006
Figure 7. The maximum dynamic pressure distribution on the cylindrical surface under different wave heights at a wave period of 2.77 s (ds/d = 0.50): (a) H = 0.18 m; (b) H = 0.66 m.
Figure 7. The maximum dynamic pressure distribution on the cylindrical surface under different wave heights at a wave period of 2.77 s (ds/d = 0.50): (a) H = 0.18 m; (b) H = 0.66 m.
Water 16 00028 g007
Figure 8. The maximum dynamic pressure distribution on the cylindrical surface under different wave heights at a wave period of 2.77 s (ds/d = 0.25): (a) H = 0.18 m; (b) H = 0.66 m.
Figure 8. The maximum dynamic pressure distribution on the cylindrical surface under different wave heights at a wave period of 2.77 s (ds/d = 0.25): (a) H = 0.18 m; (b) H = 0.66 m.
Water 16 00028 g008
Figure 9. The maximum dynamic pressure distribution on the cylindrical surface under different wave heights at a wave period of 3.45 s (ds/d = 0.50): (a) H = 0.24 m; (b) H = 0.74 m.
Figure 9. The maximum dynamic pressure distribution on the cylindrical surface under different wave heights at a wave period of 3.45 s (ds/d = 0.50): (a) H = 0.24 m; (b) H = 0.74 m.
Water 16 00028 g009
Figure 10. Experiment photos of wave and submerged cylinder: (a) non-breaking wave; (b) breaking wave.
Figure 10. Experiment photos of wave and submerged cylinder: (a) non-breaking wave; (b) breaking wave.
Water 16 00028 g010
Figure 11. Wave force measurement and analysis results in x direction (H = 0.78 m, T = 4.51 s, ds/d = 0.50): (a) wave force time curves; (b) FFT results.
Figure 11. Wave force measurement and analysis results in x direction (H = 0.78 m, T = 4.51 s, ds/d = 0.50): (a) wave force time curves; (b) FFT results.
Water 16 00028 g011
Figure 12. Wave force measurement and analysis results in z direction (H = 0.78 m, T = 4.51 s, ds/d = 0.50): (a) wave force time curves; (b) FFT results.
Figure 12. Wave force measurement and analysis results in z direction (H = 0.78 m, T = 4.51 s, ds/d = 0.50): (a) wave force time curves; (b) FFT results.
Water 16 00028 g012
Figure 13. First- and second-order wave forces at submerged depth ds/d = 0.25: (a) T = 2.51 s; (b) T = 2.77 s.
Figure 13. First- and second-order wave forces at submerged depth ds/d = 0.25: (a) T = 2.51 s; (b) T = 2.77 s.
Water 16 00028 g013
Figure 14. First- and second-order wave forces in different wave periods at submerged depth ds/d = 0.25: (a) First-order results; (b) second-order results.
Figure 14. First- and second-order wave forces in different wave periods at submerged depth ds/d = 0.25: (a) First-order results; (b) second-order results.
Water 16 00028 g014
Figure 15. First-order nondimensional wave force in different wave periods at submerged depth ds/d = 0.25: (a) x direction; (b) z direction.
Figure 15. First-order nondimensional wave force in different wave periods at submerged depth ds/d = 0.25: (a) x direction; (b) z direction.
Water 16 00028 g015
Figure 16. Second-order nondimensional wave force in different wave periods at submerged depth ds/d = 0.25: (a) x direction; (b) z direction.
Figure 16. Second-order nondimensional wave force in different wave periods at submerged depth ds/d = 0.25: (a) x direction; (b) z direction.
Water 16 00028 g016
Figure 17. First- and second-order wave forces at submerged depth ds/d = 0.50: (a) T = 2.51 s; (b) T = 2.77 s.
Figure 17. First- and second-order wave forces at submerged depth ds/d = 0.50: (a) T = 2.51 s; (b) T = 2.77 s.
Water 16 00028 g017
Figure 18. First-order wave forces in different wave periods at submerged depth ds/d = 0.50.
Figure 18. First-order wave forces in different wave periods at submerged depth ds/d = 0.50.
Water 16 00028 g018
Figure 19. First-order nondimensional wave force in different wave periods at submerged depth ds/d = 0.50: (a) x direction; (b) z direction.
Figure 19. First-order nondimensional wave force in different wave periods at submerged depth ds/d = 0.50: (a) x direction; (b) z direction.
Water 16 00028 g019
Figure 20. Second-order nondimensional wave force in different wave periods at submerged depth ds/d = 0.50: (a) x direction; (b) z direction.
Figure 20. Second-order nondimensional wave force in different wave periods at submerged depth ds/d = 0.50: (a) x direction; (b) z direction.
Water 16 00028 g020
Figure 21. Inertia force coefficient of submerged cylinder at different Reynolds numbers.
Figure 21. Inertia force coefficient of submerged cylinder at different Reynolds numbers.
Water 16 00028 g021
Figure 22. The reflected waves and corresponding frequency domain results (ds/d = 0.25, H = 0.12 m, T = 2.25 s).
Figure 22. The reflected waves and corresponding frequency domain results (ds/d = 0.25, H = 0.12 m, T = 2.25 s).
Water 16 00028 g022
Figure 23. The reflected waves and corresponding frequency domain results (ds/d = 0.25, H = 0.30 m, T = 2.25 s).
Figure 23. The reflected waves and corresponding frequency domain results (ds/d = 0.25, H = 0.30 m, T = 2.25 s).
Water 16 00028 g023aWater 16 00028 g023b
Figure 24. The reflected waves and corresponding frequency domain results (ds/d = 0.25, H = 0.24 m, T = 3.45 s).
Figure 24. The reflected waves and corresponding frequency domain results (ds/d = 0.25, H = 0.24 m, T = 3.45 s).
Water 16 00028 g024
Figure 25. The reflected waves and corresponding frequency domain results (ds/d = 0.25, H = 0.72 m, T = 3.45 s).
Figure 25. The reflected waves and corresponding frequency domain results (ds/d = 0.25, H = 0.72 m, T = 3.45 s).
Water 16 00028 g025aWater 16 00028 g025b
Figure 26. Reflection coefficient of submerged cylinder on the wave field at different submerged depths: (a) ds/d = 0.25; (b) ds/d = 0.50.
Figure 26. Reflection coefficient of submerged cylinder on the wave field at different submerged depths: (a) ds/d = 0.25; (b) ds/d = 0.50.
Water 16 00028 g026
Table 1. The introduction of different parameters.
Table 1. The introduction of different parameters.
ParameterMeaningParameterMeaning
DDiameter of submerged cylinderKCKeulegan–Carpenter number
dsSubmerged depthLWave length
dWater depthηrReflected wave elevation
WPWave probeCRReflection coefficient
HWave heightCMInertia force coefficient
TWave periodF(1)/F(2)First-/Second-order wave force
kWave numberρWater density
ReReynolds numbergGravity acceleration
Table 2. Model test table.
Table 2. Model test table.
No.d/mH/mT/skAH/gT2d/gT2Re
(ds/d = 0.25)
Re
(ds/d = 0.50)
KC
(ds/d = 0.25)
KC
(ds/d = 0.50)
14.20.1232.250.04892.47 × 10−38.43 × 10−21.11 × 1055.07 × 1040.1690.075
24.20.2362.250.09384.74 × 10−38.43 × 10−22.13 × 1059.20 × 1040.3230.144
34.20.3012.250.11966.04 × 10−38.43 × 10−22.72 × 1051.21 × 1050.4120.184
44.20.4172.250.16588.37 × 10−38.43 × 10−23.77 × 1051.65 × 1050.5720.255
54.20.162.510.05162.59 × 10−36.80 × 10−21.42 × 1057.42 × 1040.2610.139
64.20.272.510.08714.37 × 10−36.80 × 10−22.40 × 1051.25 × 1050.4410.235
74.20.372.510.11935.99 × 10−36.80 × 10−23.29 × 1051.72 × 1050.6040.322
84.20.5282.510.17038.55 × 10−36.80 × 10−24.70 × 1052.45 × 1050.8610.459
94.20.1742.770.04662.32 × 10−35.59 × 10−21.52 × 1058.66 × 1040.3250.198
104.20.3132.770.08394.17 × 10−35.59 × 10−22.73 × 1051.69 × 1050.5850.356
114.20.4992.770.13376.64 × 10−35.59 × 10−24.35 × 1052.68 × 1050.9330.568
124.20.652.770.17428.65 × 10−35.59 × 10−25.67 × 1053.30 × 1051.2150.740
134.20.2023.110.04422.13 × 10−034.43 × 10−21.71 × 1051.15 × 1050.4370.301
144.20.4323.110.09464.55 × 10−34.43 × 10−23.66 × 1052.45 × 1050.9350.644
154.20.643.110.14026.75 × 10−34.43 × 10−25.43 × 1053.63 × 1051.3860.953
164.20.2373.450.04372.03 × 10−33.59 × 10−21.97 × 1051.42 × 1050.5790.434
174.20.5033.450.09284.30 × 10−33.59 × 10−24.18 × 1053.19 × 1051.2290.921
184.20.7383.450.13626.31 × 10−33.59 × 10−26.13 × 1054.57 × 1051.8041.352
194.20.2963.980.04411.91 × 10−32.70 × 10−22.41 × 1051.91 × 1050.8500.693
204.20.493.980.07303.15 × 10−32.70 × 10−23.98 × 1053.16 × 1051.4071.148
214.20.6993.980.10424.50 × 10−32.70 × 10−25.68 × 1054.51 × 1052.0081.637
224.20.3514.510.04421.76 × 10−32.11 × 10−22.81 × 1052.30 × 1051.1580.994
234.20.7454.510.09393.74 × 10−32.11 × 10−25.96 × 1054.78 × 1052.4592.111
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

Jin, R.; Zhao, X.; Liu, Y.; Geng, B. Large Scale Experimental Study on Waves and Submerged Horizontal Cylinders. Water 2024, 16, 28. https://doi.org/10.3390/w16010028

AMA Style

Jin R, Zhao X, Liu Y, Geng B. Large Scale Experimental Study on Waves and Submerged Horizontal Cylinders. Water. 2024; 16(1):28. https://doi.org/10.3390/w16010028

Chicago/Turabian Style

Jin, Ruijia, Xu Zhao, Ye Liu, and Baolei Geng. 2024. "Large Scale Experimental Study on Waves and Submerged Horizontal Cylinders" Water 16, no. 1: 28. https://doi.org/10.3390/w16010028

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