Skip to main content
Log in

Dynamic responses of TLP considering coupled transverse and axial effect of tether under second-order wave and freak wave

  • Review article
  • Published:
Journal of Marine Science and Technology Aims and scope Submit manuscript

Abstract

The objective of this paper is to analyze the dynamic responses of Tension Leg Platform (TLP). Under the combined action of waves and currents, a tether model considering the coupled effects of transverse and axial responses is established. The finite difference method is used in the numerical algorithm of the tether vibration model. Three simulations are performed, including a first-order irregular wave, a second-order wave, and a freak wave impact against the platform. The dynamic response characteristics of the tension leg platform under the action of freak waves are calculated by using the superposition model to generate the time history of freak waves. The numerical simulation results show that the platform motions are significantly affected by the freak waves, especially in the focusing location. And the effects of second-order wave force and freak wave force on the platform responses, horizontal offset, and vertical subsidence are analyzed.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Fig. 9
Fig. 10
Fig. 11
Fig. 12
Fig. 13

Similar content being viewed by others

References

  1. Malaeke H, Moeenfard H (2016) Analytical modeling of large amplitude free vibration of non-uniform beams carrying a both transversely and axially eccentric tip mass. J Sound Vib 366:211–229

    Article  Google Scholar 

  2. Prescott J (1961) Applied elasticity. Dover Publications Inc, New York

    MATH  Google Scholar 

  3. Fox EN (1959) Applied mathematics for engineers and physicists. Phys Today 12(4):40–42

    Article  Google Scholar 

  4. Goel RP (1973) Vibrations of a beam carrying a concentrated mass. J Appl Mech 40(3):821–821

    Article  MathSciNet  Google Scholar 

  5. Jefferys ER, Patel MH (1981) Dynamic analysis models of the tension leg platform. In: Offshore technology conference

  6. Chucheepsakul S, Buncharoen S, Wang CM (1994) Large deflection of beams under moment gradient. J Eng Mech 120(9):1848–1860

    Article  Google Scholar 

  7. Patel MH, Park HI (1995) Tensioned buoyant platform tether response to short duration tension loss. Mar Struct 8(5):543–553

    Article  Google Scholar 

  8. Han SM, Benaroya H (2000) Non-linear coupled transverse and axial vibration of a compliant structure, part 1: formulation and free vibration. J Sound Vib 237(5):837–873

    Article  Google Scholar 

  9. Han SM, Benaroya H (2000) Non-linear coupled transverse and axial vibration of a compliant structure, part 2: forced vibration. J Sound Vib 237(5):875–900

    Article  Google Scholar 

  10. Han SM, Benaroya H (2002) Comparison of linear and nonlinear responses of a compliant tower to random wave forces. Chaos Solitons Fract 14(2):269–291

    Article  Google Scholar 

  11. Yigit AS, Christoforou AP (1996) Coupled axial and transverse vibrations of oilwell drillstrings. J Sound Vib 195(4):617–627

    Article  Google Scholar 

  12. Kim N, Kim CH (2003) Investigation of a dynamic property of draupner freak wave. Int J Offshore Polar Engineers 13(1):38–42

    Google Scholar 

  13. Deng Y et al (2016) Freak wave forces on a vertical cylinder—ScienceDirect. Coast Eng 114:9–18

    Article  Google Scholar 

  14. Alagan-Chella M, Bihs H, Myrhaug D (2019) Wave impact pressure and kinematics due to breaking wave impingement on a monopile. J Fluids Struct 86:94–123

    Article  Google Scholar 

  15. Fedele F (2008) Rogue waves in oceanic turbulence. Phys D 237(14–17):2127–2131

    Article  MathSciNet  Google Scholar 

  16. Kharif C, Pelinovsky E (2003) Physical mechanisms of the rogue wave phenomenon. Eur J Mech B Fluids 22(6):603–634

    Article  MathSciNet  Google Scholar 

  17. Fochesato C, Grilli S, Dias F (2007) Numerical modeling of extreme rogue waves generated by directional energy focusing. Wave Motion 44(5):395–416

    Article  MathSciNet  Google Scholar 

  18. Zhao X, Hu C (2012) Numerical and experimental study on a 2-D floating body under extreme wave conditions. Appl Ocean Res 35:1–13

    Article  Google Scholar 

  19. Cheng L, Lin P (2018) The numerical modeling of coupled motions of a moored floating body in waves. Water 10:12

    Google Scholar 

  20. Wei Y et al (2016) Wave interaction with an oscillating wave surge converter. Part II: slamming. Ocean Eng 113(1):319–334

    Article  Google Scholar 

  21. Rudman M, Cleary PW (2013) Rogue wave impact on a tension leg platform: The effect of wave incidence angle and mooring line tension. Ocean Eng 61(6):123–138

    Article  Google Scholar 

  22. Rudman M, Cleary PW (2016) The influence of mooring system in rogue wave impact on an offshore platform. Ocean Eng 115:168–181

    Article  Google Scholar 

  23. Chuang WL, Chang KA, Mercier R (2017) Impact pressure and void fraction due to plunging breaking wave impact on a 2D TLP structure. Exp Fluids 58:6

    Article  Google Scholar 

  24. Abdussamie N et al (2017) Experimental investigation of extreme wave impacts on a rigid TLP model in cyclonic conditions. Ships Offshore Struct 12(1–2):153–170

    Article  Google Scholar 

  25. Clauss G et al (1995) Offshore structures. Springer, Berlin, p 298

    Google Scholar 

  26. Chakrabarti S (2005) Handbook of offshore engineering. Springer, Berlin, pp 79–131

    Book  Google Scholar 

  27. Morison JR, Johnson JW, Schaaf SA (1950) The force exerted by surface waves on piles. Petrol Trans A 189(5):149–154

    Google Scholar 

  28. Akhmediev N, Ankiewicz A, Sotocrespo JM (2009) Rogue waves and rational solutions of the nonlinear Schrödinger equation. Phys Rev E 80(2):026601

    Article  Google Scholar 

  29. Peregrine DH (1983) Water waves, nonlinear Schrdinger equations and their solutions. ANZIAM J 25(1):16–43

    MATH  Google Scholar 

  30. Slunyaev A et al (2013) Super-rogue waves in simulations based on weakly nonlinear and fully nonlinear hydrodynamic equations. Phys Rev E 88(1):12909

    Article  Google Scholar 

  31. Hu Z et al (2015) Numerical study of Rogue waves as nonlinear Schrdinger breather solutions under finite water depth. Wave Motion 52:81–90

    Article  MathSciNet  Google Scholar 

  32. Mei CC (1992) The applied dynamics of ocean surface waves. In: The applied dynamics of ocean surface waves

  33. Pei YG, Zhang NC, Zhang YQ (2007) Efficient generation of freak waves in laboratory. China Ocean Eng 21(3):515–523

    Google Scholar 

  34. Tan SG, Boom WD (1981) the wave induced motions of a tension leg platform in deep water. In: Offshore Technology Conference

  35. Wang J et al (2021) Numerical study of uncoupled and coupled TLP models. Polish Maritime Res 28(3):44–60

    Article  Google Scholar 

  36. Masciola MD (2011) Dynamic analysis of a coupled and an uncoupled tension leg platform. McGill University, Canada

    Google Scholar 

  37. Wang J et al (2020) A method based on LS-SVM to estimate time-domain Green function. J Mar Sci Technol 5:1–13

    Google Scholar 

  38. Clément AH (1998) An ordinary differential equation for the Green function of time-domain free-surface hydrodynamics. J Eng Math 33(2):201–217

    Article  MathSciNet  Google Scholar 

  39. Angelus J, Bureau (2011) Bureau of ocean energy management, regulation and enforcement

  40. Kara F (2000) Time domain hydrodynamics and hydroelastic analysis of floating bodies with forward speed. In: University of Strathclyde: Glasgow Scotland

Download references

Acknowledgements

This work was financially supported by National Natural Science Foundation of China (No. 52075469) and Shaanxi Key Laboratory of Hydraulic Technology Fund (Grant No.: YYJS2022KF15).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Rui Guo.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Appendices

Appendix A: the platform hydrodynamic coefficients

Under the action of waves, the platform hydrodynamic coefficients are determined by BEM. Since the platform is a slender structure, the effect of wave viscous on the slender structure cannot be ignored. Then, the drag force on floating structure is determined by Morison equation. Therefore, hydrodynamic coefficient of the platform is calculated by the methods of BEM and Morison equation.

1.1 Added mass and memory function

The added mass and memory function can be determined by calculating the radiation potential of the platform. The integral equation of the radiation potential is calculated by using the boundary element method. The expression [25, 26] is shown as follows:

$$\begin{gathered} \hfill 2\pi \Phi_{k} (p,t) + \iint\limits_{{s_{b} }} {\left[ \begin{gathered} \Phi_{k} (q,t)\frac{\partial }{{\partial n_{q} }}\left( {\frac{1}{r} - \frac{1}{{r^{\prime}}}} \right) \hfill \\ - \left( {\frac{1}{r} - \frac{1}{{r^{\prime}}}} \right)\frac{{\partial \Phi_{k} (q,t)}}{{\partial n_{q} }} \hfill \\ \end{gathered} \right]{\text{d}}S_{q} } \\ \hfill = \int_{0}^{t} {{\text{d}}\tau } \iint\limits_{{s_{b} }} {\left( \begin{gathered} \tilde{G}\left( {p,t;q,\tau } \right)\frac{{\partial \Phi_{k} (q,t)}}{{\partial n_{q} }} \hfill \\ - \Phi_{k} (q,t)\frac{{\partial \tilde{G}\left( {p,t;q,\tau } \right)}}{{\partial n_{q} }} \hfill \\ \end{gathered} \right){\text{d}}S_{q} }, \\ \end{gathered}$$
(47)

where \(\Phi_{k}\) is the velocity potential which needs to be evaluated; the coordinates of field point \(p\) and source point \(q\) are \((x,y,z)\) and \((\xi ,\eta ,\zeta )\), respectively, with both lying in the water \((z \le 0)\); \(R = \sqrt {(x - \xi )^{2} + (y - \eta )^{2} }\); \(r = \sqrt {R^{2} + (z - \zeta )^{2} }\); \(r^{\prime} = \sqrt {R^{2} + (z + \zeta )^{2} }\) represent the horizontal or the three-dimensional distance between field point \(p\) and source point \(q\) or the \(q\)’s image about the still water surface; \(s_{b}\) is the wet surface of the floating structure; \({\mathbf{n}}_{q}\) is the unit normal vector of any source point on the wet surface; \(\tilde{G}\left( {p,t;q,\tau } \right)\) is the memory term of the corresponding time domain Green function. Its expression [37] is

$$\tilde{G}\left( {p,q;t} \right) = 2\int_{0}^{\infty } {{\text{d}}k\sqrt {gk} \sin \left( {\sqrt {gk} t} \right)e^{{k\left( {z + \zeta } \right)}} J_{0} \left( {kR} \right)} ,$$
(48)

where \(J_{0}\) is the Bessel function of the first kind with order 0.

The calculation method of Green function is based on Clement [38]. The memory of Green function is transformed into a fourth-order ordinary differential equation, and then the Runge–Kutta method is adopted to calculate the differential equation.

According to the method of Fual [39], Eq. (47) is decomposed into instantaneous part and memory part. The expressions of them are

$$\begin{gathered} \psi_{k} \left( p \right) + \frac{1}{2\pi }\iint\limits_{{s_{b} }} {\psi_{k} }\left( q \right)\frac{\partial }{{\partial n_{q} }}\left( {\frac{1}{r} - \frac{1}{{r^{\prime}}}} \right){\text{d}}S_{q} \hfill \\ = \frac{1}{2\pi }\iint\limits_{{s_{b} }} {\left( {\frac{1}{r} - \frac{1}{{r^{\prime}}}} \right)n_{k} {\text{d}}S_{q} } \hfill \\ \end{gathered}$$
(49)
$$\begin{aligned} \chi_{k} (p,t) & + \frac{1}{2\pi }\iint\limits_{{s_{b} }} {\chi_{k} (q,t)\frac{\partial }{{\partial n_{q} }}\left( {\frac{1}{r} - \frac{1}{{r^{\prime}}}} \right){\text{d}}S_{q} } \\ & \; = \frac{1}{2\pi }\iint\limits_{{s_{b} }} {n_{k} (q)\tilde{G}\left( {p,t;q,0} \right){\text{d}}S_{q} } \\ & \;\;\; - \frac{1}{2\pi }\iint\limits_{{s_{b} }} {\psi_{k} \left( q \right)\frac{\partial }{{\partial n_{q} }}\tilde{G}\left( {p,t;q,0} \right){\text{d}}S_{q} } \\ & \;\;\; - \frac{1}{2\pi }\int_{0}^{t} {{\text{d}}\tau } \iint\limits_{{s_{b} }} {\chi_{k} (q,\tau )\frac{{\partial \tilde{G}\left( {p,t;q,\tau } \right)}}{{\partial n_{q} }}{\text{d}}S_{q} }. \\ \end{aligned}$$
(50)

Equations (49) and (50) are, respectively, numerical discrete. Wet surface of the 3-d object is used by a series of quadrilateral grid to approximate. In each quadrilateral grid, the velocity potential is taken as a constant, and then the algebraic equation sets of instantaneous part and memory part are

$$\sum\limits_{j = 1}^{M} {A_{ij} \left( {\psi_{k} } \right)_{j} = B_{i} } \, i = 1,2, \ldots ,M$$
(51)
$$\sum\limits_{j = 1}^{M} {A_{ij} \left[ {\chi_{k} \left( {t_{N} } \right)} \right]_{j} = B\left( {t_{N} } \right)_{i} } \, i = 1,2, \ldots ,M,$$
(52)

where \(M\) is the number of quadrilateral elements; \(N\) is the current time steps;\(\left( {\psi_{k} } \right)_{j}\) is the instantaneous part of the j element; \(\left[ {\chi_{k} \left( {t_{N} } \right)} \right]_{j}\) is the memory part of the j element in the time of \(t_{N}\); \(A_{ij}\), \(B_{i}\), and \(B\left( {t_{N} } \right)_{i}\) are the coefficient matrixes, and the expressions are

$$A_{ij} = \left\{ \begin{gathered} \iint\limits_{{s_{j} }} {\frac{\partial }{{\partial n_{j} }}\left( {\frac{1}{{r_{ij} }} - \frac{1}{{r_{ij}^{\prime } }}} \right)}{\text{d}}S, \, i \ne j \hfill \\ 2\pi - \iint\limits_{{s_{j} }} {\frac{\partial }{{\partial n_{j} }}\left( {\frac{1}{{r_{ij}^{\prime } }}} \right)}{\text{d}}S, \, i = j \hfill \\ \end{gathered} \right.$$
(53)
$$B_{i} = \sum\limits_{j = 1}^{M} {\iint\limits_{{s_{j} }} {\left( {\frac{1}{{r_{ij} }} - \frac{1}{{r_{ij}^{\prime } }}} \right)\left( {n_{k} } \right)_{j} {\text{d}}S}}$$
(54)
$$\begin{aligned} B\left( {t_{N} } \right)_{i} & = \sum\limits_{j = 1}^{M} {\iint\limits_{{s_{j} }} {\left[ {n_{k} } \right]_{j} \tilde{G}\left( {p_{i} ,q_{j} ,t_{N} } \right){\text{d}}S}} \\ & \;\;\; - \sum\limits_{j = 1}^{M} {\iint\limits_{{s_{j} }} {\left[ {\psi_{k} } \right]_{j} \frac{\partial }{{\partial n_{j} }}\tilde{G}\left( {p_{i} ,q_{j} ,t_{N} } \right){\text{d}}S}} \\ & \;\;\; - \Delta t\sum\limits_{n = 1}^{N} {\sum\limits_{j = 1}^{M} {\iint\limits_{{s_{j} }} {\left[ {\chi_{k} \left( {q_{j} ,t_{n} } \right)} \right]_{j} \frac{\partial }{{\partial n_{j} }}\tilde{G}\left( {p_{i} ,q_{j} ,t_{N} - t_{n} } \right){\text{d}}S}} } , \\ \end{aligned}$$
(55)

where \(\Delta t\) is the time step;\(t_{N} = N\Delta t\); \(t_{n} = n\Delta t\).

By the calculation of algebraic equation sets, the integral equations of added mass and memory function are

$$\mu_{ik} = \rho \iint\limits_{{s_{b} }} {\psi_{k} \left( q \right)n_{j} {\text{d}}S_{q} }$$
(56)
$$\overline{K}_{jk} \left( t \right) = \rho \iint\limits_{{s_{b} }} {\frac{\partial }{\partial t}\chi_{k} \left( {q,t} \right){\text{d}}S_{q} }$$
(57)
$$A_{jk} \left( \omega \right) = \mu_{jk} - \frac{1}{\omega }\int_{0}^{t} {K_{jk} \left( \tau \right)\sin \left( {\omega \tau } \right){\text{d}}\tau }$$
(58)
$$B_{jk} \left( \omega \right) = \int_{0}^{t} {K_{jk} \left( \tau \right)\cos \left( {\omega \tau } \right){\text{d}}\tau } ,$$
(59)

where\(\overline{K}_{jk} \left( t \right)\) is the memory function;\(A_{jk} \left( \omega \right)\) is the added mass;\(B_{jk} \left( \omega \right)\) is the damping coefficient.

1.2 The damping matrix

The damping matrix assumption is related to the mass matrix and the restoring force matrix, which can be expressed as follows:

$$\Phi^{T} \left[ K \right]\Phi = \left[ {2\zeta_{i} \omega_{i} m_{i} } \right],$$
(60)

where\(\Phi\) is the shape matrix, \(\omega_{i}\) is the structure natural frequency, K is the damping matrix, \(m_{i}\) is the modal mass, \(m_{i} = \Phi^{T} \left[ M \right]\Phi\), and \(\zeta_{i}\) is the dimensionless damping ratio of each mode, generally 0.05.

The solution of \(\Phi\) and \(\omega_{i}\):

For the undamped vibration:

$$M\ddot{Y} + KY = 0.$$
(61)

The characteristic determinant is

$$\left| {T - \lambda M} \right| = 0.$$
(62)

The characteristic vector \(\left[ {\Phi_{1} ,\Phi_{2} \ldots \Phi_{n} } \right]\) and characteristic value \(\left[ {\begin{array}{*{20}c} {\lambda_{1} } & {} & {} & {} \\ {} & {\lambda_{2} } & {} & {} \\ {} & {} & \ddots & {} \\ {} & {} & {} & {\lambda_{n} } \\ \end{array} } \right]\) are obtained:

$$\lambda_{i} = \omega_{i}^{2} \;\;\;\omega_{i} = \sqrt {\lambda_{i} } .$$
(63)

The modal mass is

$$\left[ {\begin{array}{*{20}c} {m_{1} } & {} & {} & {} \\ {} & {m_{2} } & {} & {} \\ {} & {} & \ddots & {} \\ {} & {} & {} & {m_{n} } \\ \end{array} } \right] = \left[ \Phi \right]^{T} \left[ M \right]\left[ \Phi \right].$$
(64)

The K matrix can be obtained by solving the following matrix equation:

$$\Phi^{T} \left[ c \right]\Phi = 2\xi_{i} \left[ {\begin{array}{*{20}c} {\sqrt {\lambda_{1} } } & {} & {} & {} \\ {} & {\sqrt {\lambda_{2} } } & {} & {} \\ {} & {} & \ddots & {} \\ {} & {} & {} & {\sqrt {\lambda_{n} } } \\ \end{array} } \right]\left[ {\begin{array}{*{20}c} {m_{1} } & {} & {} & {} \\ {} & {m_{2} } & {} & {} \\ {} & {} & \ddots & {} \\ {} & {} & {} & {m_{n} } \\ \end{array} } \right].$$
(65)

Wave Force

2.1 Froude–Krylov force

Froude–Krylov force is caused by the incident potential, and its expression is

$$F_{jI} \left( t \right) = \iint\limits_{{s_{b} }} {p\left( {q,t} \right)n_{j} {\text{d}}S_{q} },$$
(66)

where \(F_{jI}\) is the Froude–Krylov force in the \(j\left( {j = 1,2, \ldots 6} \right)\) mode of motions. \(n_{j}\) is the unit surface normal vector.

The expression [40] of the dynamic pressure acting on the floating structure by the incident waves is

$$p\left( {p,t} \right) = \int_{ - \infty }^{ + \infty } {\hat{p}\left( {p,t - \tau } \right)\zeta_{0} \left( \tau \right){\text{d}}\tau } ,$$
(67)

where \(\hat{p}\left( {p,t} \right)\) is the impulse response function of pressure. The expression is

$$\hat{p}\left( {p,t} \right) = \frac{\rho g}{\pi }{\text{Re}} \left\{ {\int_{0}^{\infty } {e^{{k\left( {z - i\alpha } \right)}} e^{i\omega t} {\text{d}}\omega } } \right\}.$$
(68)

When the incident wave is a regular wave per unit wavelength, \(\zeta_{0} \left( t \right) = e^{i\omega t}\).

2.2 Diffraction force

The diffraction force is mainly caused by the diffraction potential, and its expression is

$$\begin{aligned} 2\pi \hat{\Phi }_{7} (p,t) & + \iint\limits_{{s_{b} }} {\hat{\Phi }_{7} (q,t)\frac{\partial }{{\partial n_{q} }}\left( {\frac{1}{r} - \frac{1}{{r^{\prime}}}} \right){\text{d}}S_{q} } \\ & = - \iint\limits_{{s_{b} }} {\left( {\frac{1}{r} - \frac{1}{{r^{\prime}}}} \right)\left[ {{\vec{\mathbf{n}}} \cdot \hat{K}\left( {q,t} \right)} \right]{\text{d}}S_{q} } \\ & \;\;\; - \int_{0}^{t} {{\text{d}}\tau } \iint\limits_{{s_{b} }} {\left\{ \begin{gathered} \tilde{G}\left( {p,t;q,\tau } \right)\left[ {{\vec{\mathbf{n}}} \cdot \hat{K}\left( {q,t} \right)} \right] \hfill \\ + \hat{\Phi }_{7} \left( {q,\tau } \right)\frac{{\partial \tilde{G}\left( {p,t;q,\tau } \right)}}{{\partial n_{q} }} \hfill \\ \end{gathered} \right\}{\text{d}}S_{q} }, \\ \end{aligned}$$
(69)

where the impulse response function \(\hat{K}\left( {p,t} \right)\) is

$$\hat{K}\left( {p,t} \right) = \frac{1}{\pi }{\text{Re}} \left\{ {\left[ \begin{gathered} {\vec{\mathbf{i}}}\cos \beta \hfill \\ {\vec{\mathbf{j}}}\sin \beta \hfill \\ {\vec{\mathbf{k}}}i \hfill \\ \end{gathered} \right]\int_{0}^{\infty } {e^{{k\left( {z - i\alpha } \right)}} e^{i\omega t} {\text{d}}\omega } } \right\}.$$
(70)

The calculation method of diffraction potential is the same as that of radiation potential. By solving the integral equation of the diffraction potential, the expression of diffraction force is

$$F_{j7} = \int_{ - \infty }^{\infty } {\zeta_{0} \left( \tau \right){\text{d}}\tau \left[ { - \rho \iint\limits_{{s_{b} }} {\frac{\partial }{\partial t}\hat{\Phi }_{7} \left( {q,t - \tau } \right)n_{j} {\text{d}}S_{q} }} \right]} .$$
(71)

2.3 Drag force

For slender columns, the influence of the drag force on the platform can be calculated by the Morison equation, and the drag force per unit length of column is

$$f_{{\text{D}}} = \frac{1}{2}\rho C_{D} D\left( {w_{x} - \dot{x}} \right)\left| {\left( {w_{x} - \dot{x}} \right)} \right|,$$
(72)

where \(D\) is the diameter of column;\(C_{{\text{D}}}\) is the coefficient of drag force;\(w_{x}\) is horizontal component of the velocity of the fluid particle;\(\dot{x}\) is the velocity of floating structure.

The determination method of drag force coefficient [36] is

$$C_{{\text{D}}} = \left\{ \begin{gathered} 60.566 - 5.93{\text{Re}} {\text{ Re}} \le 10 \hfill \\ 1.25{\text{ 10 < Re}} \le 5 \times 10^{5} \hfill \\ 0.7 \, {\text{Re}} > 5 \times 10^{5} \hfill \\ \end{gathered} \right.,$$
(73)

where Re is the Reynolds number that is \({\text{Re}} = UD/\nu\). \(U\) is the relative velocity of fluid. D is characteristic length. In this paper, \(D\) is the diameter of column. \(\nu\) is the coefficient of kinematic viscosity, which is 1.346e−6 m2/s when the temperature of seawater is 10 °C.

The velocity of fluid particle is calculated by Airy linear wave theory. The velocity of water particle (x direction) is

$$w_{{x^{\prime}}} = \frac{{\pi W_{{\text{H}}} }}{{W_{{\text{T}}} }}\frac{{\sinh \left( {kx} \right)}}{{\sinh \left( {kd} \right)}}\sin \left( {kz - \omega t} \right).$$
(74)

The velocity of water particle (y direction) is

$$w_{{y^{\prime}}} = 0.$$
(75)

The velocity of water particle (z direction) is

$$w_{{z^{\prime}}} = \frac{{\pi W_{{\text{H}}} }}{{W_{{\text{T}}} }}\frac{{\cosh \left( {kx} \right)}}{{\sinh \left( {kd} \right)}}\cos \left( {kz - \omega t} \right),$$
(76)

where \(W_{{\text{H}}}\) is the wave height;\(W_{{\text{T}}}\) is the wave period; k is the wave number;\(\omega\) is the wave frequency; and d is the depth of water.

According to method of the coordinate transformation, transforms the particle velocity in the wave coordinate system \(O^{\prime}X^{\prime}Y^{\prime}Z^{\prime}\) to the static coordinate system OXYZ:

$$\left\{ \begin{gathered} w_{x} = w_{{x^{\prime}}} \cos \alpha \hfill \\ w_{y} = w_{{y^{\prime}}} \cos \alpha \hfill \\ w_{z} = w_{{z^{\prime}}} \hfill \\ \end{gathered} \right.$$
(77)

\(\alpha\) is the angle between the coordinate system \(O^{\prime}X^{\prime}Y^{\prime}Z^{\prime}\) and the static coordinate system OXYZ.

The relative velocity between water particle and floating structure in the static coordinate system OXYZ is

$$\vec{V} = \left( {\dot{x}_{1} \vec{i} + \dot{x}_{2} \vec{j} + \dot{x}_{3} \vec{k}} \right) + \vec{\omega } \times \vec{r}$$
(78)
$$\vec{r} = x_{1} \vec{i} + x_{2} \vec{j} + x_{3} \vec{k},$$
(79)

where \(\vec{\omega } \times \vec{r} = \left| {\begin{array}{*{20}c} {\vec{i}} & {\vec{j}} & {\vec{k}} \\ {\dot{x}_{4} } & {\dot{x}_{5} } & {\dot{x}_{6} } \\ {x_{1} } & {x_{2} } & {x_{3} } \\ \end{array} } \right|\), and substitute it into Eq. 78. Then the relative velocity is

$$\begin{gathered} \vec{w}_{r} = w_{rx} \vec{i} + w_{ry} \vec{j} + w_{rz} \vec{k} \\ = \left[ {w_{x} - \left( {\dot{x}_{5} x_{3} - \dot{x}_{6} x_{2} + \dot{x}_{1} } \right)} \right]\vec{i} \\ + \left[ {w_{y} - \left( { - \dot{x}_{4} x_{3} + \dot{x}_{6} x_{1} + \dot{x}_{2} } \right)} \right]\vec{j} \\ + \left[ {w_{z} - \left( {\dot{x}_{4} x_{2} - \dot{x}_{5} x_{1} + \dot{x}_{3} } \right)} \right]\vec{k}. \\ \end{gathered}$$
(80)

The wet surface of the platform is separated into a series of differential rings, and the centers of the circles are taken as the discrete points. Based on the Morison equation, the drag force calculating formula in the case of the i discrete element is

$$\left[ \begin{gathered} F_{Dxi} \hfill \\ F_{Dyi} \hfill \\ F_{Dzi} \hfill \\ \end{gathered} \right] = \frac{1}{2}\rho C_{{\text{D}}} D\left( \begin{gathered} w_{rxi} \hfill \\ w_{ryi} \hfill \\ w_{rzi} \hfill \\ \end{gathered} \right)\left| \begin{gathered} w_{rxi} \hfill \\ w_{ryi} \hfill \\ w_{rzi} \hfill \\ \end{gathered} \right|$$
(81)

The calculation formula of the drag forces of the platform in waves is

$$\left\{ \begin{gathered} F_{Dx} \hfill \\ F_{Dy} \hfill \\ F_{Dz} \hfill \\ \end{gathered} \right. = \left\{ \begin{gathered} \sum\limits_{i = 1}^{N} {F_{Dxi} } \hfill \\ \sum\limits_{i = 1}^{N} {F_{Dyi} } \hfill \\ \sum\limits_{i = 1}^{N} {F_{Dzi} } \hfill \\ \end{gathered} \right.,$$
(82)

where N is the number of discrete elements on the platform.

The calculation formula of the drag moments of the platform in waves is

$$\left\{ \begin{gathered} M_{Dx} = F_{Dz} y - F_{Dy} z \hfill \\ M_{Dy} = F_{Dx} z - F_{Dz} x \hfill \\ M_{Dz} = F_{Dy} x - F_{Dx} y \hfill \\ \end{gathered} \right..$$
(83)

When the relative velocities are calculated, the velocities of the platform are required. Therefore, the formulas of wave drag forces are linked with the differential motion equations of the floating structure and the time iteration of the differential equations is carried out. Then the drag forces of waves on the platform can be calculated.

2.4 Second-order wave forces

The wave force on the platform includes not only the first-order wave force calculated above, but also the second-order wave force. The second-order wave forces consist mainly of the mean drift force, the sum frequency wave force, and the differential frequency wave force [20].

There are two methods of calculating the mean drift force, including the near-field method and the far-field method, respectively. The near-field method is based on the direct solution of the waterline surface and object area components to obtain the average drift force for the six degrees of freedom of the platform. The far-field method is achieved mainly through the law of conservation of energy, while the far-field method can only result in the calculation of the platform's component in the horizontal plane. The near-field and far-field formulations of the mean drift force are given separately [26].

Near-field equation for mean drift force.

$$F_{{{\text{d}}x}} = - \frac{\rho }{2}\left\{ {\iint_{{S_{{\text{H}}} }} {\left[ {\phi_{1} \frac{\partial }{\partial n}\left( {\frac{{\partial \phi_{1} }}{\partial x}} \right) - \frac{{\partial \phi_{1} }}{\partial x}.\frac{{\partial \phi_{1} }}{\partial n}} \right]}{\text{d}}s} \right\}_{m}$$
(84)
$$F_{{{\text{d}}y}} = - \frac{\rho }{2}\left\{ {\iint_{{S_{{\text{H}}} }} {\left[ {\phi_{1} \frac{\partial }{\partial n}\left( {\frac{{\partial \phi_{1} }}{\partial y}} \right) - \frac{{\partial \phi_{1} }}{\partial y}.\frac{{\partial \phi_{1} }}{\partial n}} \right]}{\text{d}}s} \right\}_{m} .$$
(85)

Far-field equation for mean drift force.

$$F_{{{\text{d}}x}} = - \frac{\rho }{2g}\left\{ {\int_{{l_{0} }} {\left( {\frac{{\partial \phi_{1} }}{\partial t}} \right)^{2} n_{x} {\text{d}}l} } \right\}_{m} - \rho \left\{ {\iint_{{C_{0} }} {\left[ {\frac{{\partial \phi_{1} }}{\partial x}.\frac{{\partial \phi_{1} }}{\partial n} - \frac{1}{2}\left( {\nabla \phi_{1} } \right)^{2} n_{x} } \right]}{\text{d}}s} \right\}_{m}$$
(86)
$$F_{{{\text{d}}y}} = - \frac{\rho }{2g}\left\{ {\int_{{l_{0} }} {\left( {\frac{{\partial \phi_{1} }}{\partial t}} \right)^{2} n_{y} {\text{d}}l} } \right\}_{m} - \rho \left\{ {\iint_{{C_{0} }} {\left[ {\frac{{\partial \phi_{1} }}{\partial y}.\frac{{\partial \phi_{1} }}{\partial n} - \frac{1}{2}\left( {\nabla \phi_{1} } \right)^{2} n_{y} } \right]}{\text{d}}s} \right\}_{m} .$$
(87)

As the near-field formulation of the mean drift force can calculate the motion of the platform in six degrees of freedom, while the far-field formulation can only calculate the surge, sway and yaw motions, this paper focuses on the effect of the mean drift force calculated by the near-field method on the motion response of the TLP.

The expression for the second-order force [26] can be expressed as

$$F^{\left( 2 \right)} = \frac{1}{2}\rho g\int_{\Gamma 0} {\eta^{{\left( 1 \right)^{2} }} n{\text{d}}\Gamma - \rho \iint\limits_{S} {\left( {\phi_{t}^{\left( 2 \right)} + \frac{1}{2}\left( {\nabla \phi^{\left( 1 \right)} } \right)^{2} } \right)}n{\text{d}}S} .$$
(88)

The above equation is bounded using second-order bypassing-radiation theory, and the standard form of the second-order force calculation equation

$$\begin{gathered} F^{\left( 2 \right)} \left( t \right) = A_{1}^{2} f_{{\text{d}}} \left( {\omega_{1} } \right) + h\left\{ {A_{1}^{2} + f_{ + }^{\left( 2 \right)} \left( {\omega_{1} ,\omega_{1} } \right)e^{{ - 2i\omega_{1} t}} + } \right.A_{2}^{2} f_{ + }^{\left( 2 \right)} \left( {\omega_{2} ,\omega_{2} } \right)e^{{ - 2i\omega_{2} t}} \\ \left. {{ + }2A_{1} A_{2} f_{ - }^{\left( 2 \right)} \left( {\omega_{1} ,\omega_{2} } \right)e^{{ - 2i\left( {\omega_{1} - \omega_{2} } \right)t}} + 2A_{1} A_{2} f_{ + }^{\left( 2 \right)} \left( {\omega_{1} ,\omega_{2} } \right)e^{{ - 2i\left( {\omega_{1} + \omega_{2} } \right)t}} } \right\} + A_{2}^{2} f_{{\text{d}}} \left( {\omega_{2} } \right), \\ \end{gathered}$$
(89)

where\(f_{{\text{d}}}\) is the normalized mean drift force; \(f_{ + }^{\left( 2 \right)}\) is the second-order transfer function (QTF) of the normalized sum frequency force; \(f_{ - }^{\left( 2 \right)}\) is the second-order transfer function (QTF) of the normalized difference frequency force; \(\omega_{1}\) and are the \(\omega_{2}\) wave frequencies of the two incident waves; \(A_{1}\) and \(A_{2}\) are the wave amplitudes of the two incident waves, respectively.

In this paper, the frequency domain solution of the second-order force is calculated based on the HydroD module, including the average drift force \(f_{{\text{d}}}\) and the second-order transfer function (QTF) of the sum frequency force and differential frequency force, respectively, and embedded in the coupled TLP model compiled in this paper to obtain the TLP motion responses under the influence of the second-order wave forces.

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Wang, J., Wang, Y., Guo, R. et al. Dynamic responses of TLP considering coupled transverse and axial effect of tether under second-order wave and freak wave. J Mar Sci Technol 27, 1084–1103 (2022). https://doi.org/10.1007/s00773-021-00871-5

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00773-021-00871-5

Keywords

Navigation