Skip to main content
Log in

Study on the behavior of weakly nonlinear water waves in the presence of random wind forcing

  • Original paper
  • Published:
Nonlinear Dynamics Aims and scope Submit manuscript

Abstract

Specific solutions of the nonlinear Schrödinger equation, such as the Peregrine breather, are considered to be prototypes of extreme or freak waves in the oceans. An important question is whether these solutions also exist in the presence of gusty wind. Using the method of multiple scales, a nonlinear Schrödinger equation is obtained for the case of wind-forced weakly nonlinear deep water waves. Thereby, the wind forcing is modeled as a stochastic process. This leads to a stochastic nonlinear Schrödinger equation, which is calculated for different wind regimes. For the case of wind forcing which is either random in time or random in space, it is shown that breather-type solutions such as the Peregrine breather occur even in strong gusty wind conditions.

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
Fig. 14
Fig. 15
Fig. 16
Fig. 17

Similar content being viewed by others

References

  1. Kharif, C., Pelinovsky, E., Slunyaev, A.: Rogue Waves in the Ocean. Springer, New York (2009)

    Google Scholar 

  2. Brockwell, P.J., Stramer, O.: On the approximation of continuous time threshold ARMA processes. Ann. Inst. Stat. Math. 47, 1–20 (1995)

    Google Scholar 

  3. Dostal, L., Kreuzer, E.: Probabilistic approach to large amplitude ship rolling in random seas. Proc. Inst. Mech. Eng. Part C J. Mech. Eng. Sci. 225, 2464–2476 (2011)

    Google Scholar 

  4. Rosenthal, W., Lehner, S.: Rogue waves: results of the MaxWave project. J. Offshore Mech. Arct. Eng. 130, 021006 (2008)

    Google Scholar 

  5. Onorato, M., Osborne, A.R., Serio, M., Cavaleri, L., Brandini, C., Stansberg, C.T.: Observation of strongly non-Gaussian statistics for random sea surface gravity waves in wave flume experiments. Phys. Rev. E 70, 067302 (2004)

    Google Scholar 

  6. Dysthe, K., Krogstad, H.E., Müller, P.: Oceanic rogue waves. Ann. Rev. Fluid Mech. 40, 287–310 (2008)

    Google Scholar 

  7. Sharma, J., Dean, R., et al.: Second-order directional seas and associated wave forces. Soc. Pet. Eng. J. 21, 129–140 (1981)

    Google Scholar 

  8. Forristall, G.Z.: Wave crest distributions: observations and second-order theory. J. Phys. Oceanogr. 30, 1931–1943 (2000)

    Google Scholar 

  9. Agarwal, P., Manuel, L.: Incorporating irregular nonlinear waves in coupled simulation and reliability studies of offshore wind turbines. Appl. Ocean Res. 33, 215–227 (2011)

    Google Scholar 

  10. Gibson, R., Swan, C.: The evolution of large ocean waves: the role of local and rapid spectral changes. Proc. R. Soc. Lond. A Math. Phys. Eng. Sci. 463, 21–48 (2007)

    Google Scholar 

  11. Janssen, P.A.: On some consequences of the canonical transformation in the Hamiltonian theory of water waves. J. Fluid Mech. 637, 1–44 (2009)

    Google Scholar 

  12. Annenkov, S.Y., Shrira, V.I.: Large-time evolution of statistical moments of wind-wave fields. J. Fluid Mech. 726, 517–546 (2013)

    Google Scholar 

  13. Tayfun, M.A.: Narrow-band nonlinear sea waves. J. Geophys. Res. Oceans 85, 1548–1552 (1980)

    Google Scholar 

  14. Gramstad, O., Trulsen, K.: Influence of crest and group length on the occurrence of freak waves. J. Fluid Mech. 582, 463–472 (2007)

    Google Scholar 

  15. Marino, E., Borri, C., Peil, U.: A fully nonlinear wave model to account for breaking wave impact loads on offshore wind turbines. J. Wind Eng. Ind. Aerodyn. 99, 483–490 (2011)

    Google Scholar 

  16. Dold, J.W., Peregrine, D.H.: An efficient boundary-integral method for steep unsteady water waves. Numer. Methods Fluid Dyn. II, 671–679 (1986)

    Google Scholar 

  17. Longuet-Higgins, M.S., Cokelet, E.D.: The deformation of steep surface waves on water-I. A numerical method of computation. Proc. R. Soc. Lond. A Math. Phys. Sci. 350, 1–26 (1976)

    Google Scholar 

  18. Dommermuth, D.G., Yue, D.K., Lin, W.M., Rapp, R.J., Chan, E.S., Melville, W.K.: Deep-water plunging breakers: a comparison between potential theory and experiments. J. Fluid Mech. 189, 423–442 (1988)

    Google Scholar 

  19. Grilli, S.T., Skourup, J., Svendsen, I.A.: An efficient boundary element method for nonlinear water waves. Eng. Anal. Bound. Elem. 6, 97–107 (1989)

    Google Scholar 

  20. Marino, E., Lugni, C., Borri, C.: A novel numerical strategy for the simulation of irregular nonlinear waves and their effects on the dynamic response of offshore wind turbines. Comput. Methods Appl. Mech. Eng. 255, 275–288 (2013)

    Google Scholar 

  21. Zakharov, V.E.: Stability of periodic waves of finite amplitude on the surface of a deep fluid. J. Appl. Mech. Tech. Phys. 9, 190–194 (1968)

    Google Scholar 

  22. Mei, C.C.: The Applied Dynamics of Ocean Surface Waves. Wiley, New York (1983)

    Google Scholar 

  23. Kuznetsov, E.: Solitons in a parametrically unstable plasma. Akademiia Nauk SSSR Doklady 236, 575–577 (1977)

    Google Scholar 

  24. Akhmediev, N., Eleonskii, V., Kulagin, N.: Generation of periodic trains of picosecond pulses in an optical fiber: exact solutions. Sov. Phys. JETP 62, 894–899 (1985)

    Google Scholar 

  25. Peregrine, D.: Water waves, nonlinear Schrödinger equations and their solutions. ANZIAM J. 25, 16–43 (1983)

    Google Scholar 

  26. Akhmediev, N., Ankiewicz, A., Taki, M.: Waves that appear from nowhere and disappear without a trace. Phys. Lett. A 373, 675–678 (2009)

    Google Scholar 

  27. Chabchoub, A., Hoffmann, N., Akhmediev, N.: Rogue wave observation in a water wave tank. Phys. Rev. Lett. 106, 204502 (2011)

    Google Scholar 

  28. Chabchoub, A.: Tracking breather dynamics in irregular sea state conditions. Phys. Rev. Lett. 117, 144103 (2016)

    Google Scholar 

  29. Dysthe, K.B.: Note on a modification to the nonlinear Schrödinger equation for application to deep water waves. Proc. R. Soc. A 369, 105–114 (1979)

    Google Scholar 

  30. Onorato, M., Proment, D., Clauss, G., Klein, M.: Rogue waves: from nonlinear Schrödinger breather solutions to sea-keeping test. PloS ONE 8, e54629 (2013)

    Google Scholar 

  31. Klein, M., Clauss, G.F., Rajendran, S., Soares, C.G., Onorato, M.: Peregrine breathers as design waves for wave-structure interaction. Ocean Eng. 128, 199–212 (2016)

    Google Scholar 

  32. Jeffreys, H.: On the formation of water waves by wind. Proc. R. Soc. Lond. Ser. A Contain. Papers Math. Phys. Character 107, 189–206 (1925)

    Google Scholar 

  33. Phillips, O.M.: On the generation of waves by turbulent wind. J. Fluid Mech. 2, 417–445 (1957)

    Google Scholar 

  34. Miles, J.W.: On the generation of surface waves by shear flows. J. Fluid Mech. 3, 185–204 (1957)

    Google Scholar 

  35. Miles, J.W.: On the generation of surface waves by shear flows. Part 2. J. Fluid Mech. 6, 568–582 (1959)

    Google Scholar 

  36. Montalvo, P., Kraenkel, R., Manna, M., Kharif, C.: Wind-wave amplification mechanisms: possible models for steep wave events in finite depth. Nat. Hazards Earth Syst. Sci. 13, 2805–2813 (2013)

    Google Scholar 

  37. Chambarel, J., Kharif, C., Kimmoun, O.: Generation of two-dimensional steep water waves on finite depth with and without wind. Eur. J. Mech.-B/Fluids 29, 132–142 (2010)

    Google Scholar 

  38. Janssen, P.: The Interaction of Ocean Waves and Wind. Cambridge University Press, Cambridge (2004)

    Google Scholar 

  39. Alexakis, A., Calder, A.C., Dursi, L.J., Rosner, R., Truran, J.W., Fryxell, B., Zingale, M., Timmes, F.X., Olson, K., Ricker, P.: On the nonlinear evolution of wind-driven gravity waves. Phys. Fluids 16, 3256–3268 (2004)

    Google Scholar 

  40. Hristov, T., Miller, S., Friehe, C.: Dynamical coupling of wind and ocean waves through wave-induced air flow. Nature 422, 55–58 (2003)

    Google Scholar 

  41. Leblanc, S.: Amplification of nonlinear surface waves by wind. Phys. Fluids 19, 101705 (2007)

    Google Scholar 

  42. Kharif, C., Kraenkel, R., Manna, M., Thomas, R.: The modulational instability in deep water under the action of wind and dissipation. J. Fluid Mech. 664, 138–149 (2010)

    Google Scholar 

  43. Chabchoub, A., Hoffmann, N., Branger, H., Kharif, C., Akhmediev, N.: Experiments on wind-perturbed rogue wave hydrodynamics using the Peregrine breather model. Phys. Fluids 25, 101704 (2013)

    Google Scholar 

  44. Brunetti, M., Marchiando, N., Berti, N., Kasparian, J.: Nonlinear fast growth of water waves under wind forcing. Phys. Lett. A 378, 1025–1030 (2014)

    Google Scholar 

  45. Brunetti, M., Kasparian, J.: Modulational instability in wind-forced waves. Phys. Lett. A 378, 3626–3630 (2014)

    Google Scholar 

  46. Eeltink, D., Lemoine, A., Branger, H., Kimmoun, O., Kharif, C., Carter, J., Chabchoub, A., Brunetti, M., Kasparian, J.: Spectral up-and downshifting of Akhmediev breathers under wind forcing. Phys. Fluids 29, 107103 (2017)

    Google Scholar 

  47. Carter, J.D., Curtis, C.W., Kalisch, H.: Particle Trajectories in Nonlinear Schrödinger Models. Water Waves (2019). https://doi.org/10.1007/s42286-019-00008-7

  48. Morison, J.R., Johnson, J.W., Schaaf, S.A.: The force exerted by surface waves on piles. J. Pet. Technol. 2(05), 149–154 (1950)

    Google Scholar 

  49. Antoine, X., Bao, W., Besse, C.: Computational methods for the dynamics of the nonlinear Schrödinger/Gross-Pitaevskii equations. Comput. Phys. Commun. 184, 2621–2633 (2013)

    Google Scholar 

  50. Rapti, Z., Kevrekidis, P.G., Frantzeskakis, D.J., Ba, M.: On the modulational instability of the nonlinear Schrödinger equation with dissipation. Phys. Scr. T113, 74–77 (2004)

    Google Scholar 

  51. Theocharis, G., Rapti, Z., Kevrekidis, P.G., Frantzeskakis, D.J., Konotop, V.V.: Modulational instability of Gross-Pitaevskii-type equations in \(1+1\) dimensions. Phys. Rev. A 67, 063610 (2003)

    Google Scholar 

  52. Onu, K., Namachchivaya, N.S.: Stochastically forced water waves in a circular basin. Proc. R. Soc. A 466, 2363–2381 (2010)

    Google Scholar 

  53. Rapti, Z., Kevrekidis, P.G., Smerzi, A., Bishop, A.R.: Parametric and modulational instabilities of the discrete nonlinear Schrödinger equation. J. Phys. B At. Mol. Opt. Phys. 37, 257–264 (2004)

    Google Scholar 

  54. de Bouard, A., Debussche, A.: A Stochastic nonlinear Schrödinger equation with multiplicative noise. Commun. Math. Phys. 205, 161–181 (1999)

    Google Scholar 

  55. de Bouard, A., Debussche, A., Di Menza, L.: Theoretical and numerical aspects of stochastic nonlinear Schrödinger equations. Monte Carlo Methods Appl. 7, 55–63 (2001)

    Google Scholar 

  56. Debussche, A., Di Menza, L.: Numerical simulation of focusing stochastic nonlinear Schrödinger equations. Phys. D Nonlinear Phenom. 162, 131–154 (2002)

    Google Scholar 

  57. de Bouard, A., Debussche, A.: Weak and strong order of convergence of a semidiscrete scheme for the stochastic nonlinear Schrödinger equation. Appl. Math. Optim. 54, 369–399 (2006)

    Google Scholar 

  58. de Bouard, A., Debussche, A., Di Menza, L.: Theoretical and numerical aspects of stochastic nonlinear Schrödinger equations. Journées équations aux dérivées partielles. Article no. 3 (2001)

  59. Leblanc, S.: Wind-forced modulations of finite-depth gravity waves. Phys. Fluids 20, 28–31 (2008)

    Google Scholar 

  60. Miles, J.W.: Surface-wave generation: a viscoelastic model. J. Fluid Mech. 322, 131–145 (1996)

    Google Scholar 

  61. Conte, S., Miles, J.: On the numerical integration of the Orr-Sommerfeld equation. J. Soc. Ind. Appl. Math. 7, 361–366 (1959)

    Google Scholar 

  62. Onorato, M., Proment, D.: Approximate rogue wave solutions of the forced and damped nonlinear Schrödinger equation for water waves. Phys. Lett. A 376, 3057–3059 (2012)

    Google Scholar 

  63. Besse, C.: A relaxation scheme for the nonlinear Schrödinger equation. SIAM J. Numer. Anal. 42, 934–952 (2004)

    Google Scholar 

  64. Van der Hoven, I.: Power spectrum of horizontal wind speed in the frequency range from 0.0007 to 900 cycles per hour. J. Meteorol. 14, 160–164 (1957)

    Google Scholar 

  65. Leithead, W., De la Salle, S., Reardon, D.: Role and objectives of control for wind turbines. IEE Proc. C-Gener. Trans. Distrib. 138, 135–148 (1991). IET

    Google Scholar 

  66. Nichita, C., Luca, D., Dakyo, B., Ceanga, E.: Large band simulation of the wind speed for real time wind turbine simulators. IEEE Trans. Energy Convers. 17, 523–529 (2002)

    Google Scholar 

  67. Sarpkaya, T.: Vortex shedding and resistance in harmonic flow about smooth and rough circular cylinders at high Reynolds numbers (No. NPS-59SL76021). Naval Postgraduate School Monterey CA (1976)

  68. Antoine, X., Duboscq, R.: Gpelab, a Matlab toolbox to solve Gross-Pitaevskii equations II: dynamics and stochastic simulations. Comput. Phys. Commun. 193, 95–117 (2015)

    Google Scholar 

  69. Antoine, X., Duboscq, R.: Modeling and computation of Bose-Einstein condensates: stationary states, nucleation, dynamics, stochasticity. In: Besse, C., Garreau, J.C. (eds.) Nonlinear optical and atomic systems. Lecture notes in mathematics, vol. 2146, pp. 49–145. Springer, Cham (2015)

  70. Osborne, A.R.: Nonlinear Ocean Waves and the Inverse Scattering Transform. Academic Press, Amsterdam (2010)

    Google Scholar 

  71. Dias, F., Dyachenko, A., Zakharov, V.: Theory of weakly damped free-surface flows: a new formulation based on potential flow solutions. Phys. Lett. A 372, 1297–1302 (2008)

    Google Scholar 

  72. Davey, A., Stewardson, K.: On three-dimensional packets of surface waves. Proc. R. Soc. Lond. A 338, 101–110 (1974)

    Google Scholar 

  73. Hasimoto, H., Ono, H.: Nonlinear modulation of gravity waves. J. Phys. Soc. Jpn 33, 805–811 (1972)

    Google Scholar 

  74. Djordjevic, V.D., Redekopp, L.G.: On two-dimensional packets of capillary-gravity waves. J. Fluid Mech. 79(4), 703–714 (1977)

    Google Scholar 

  75. Carter, J.D., Govan, A.: Frequency downshift in a viscous fluid. Eur. J. Mech.-B/Fluids 59, 177–185 (2016)

    Google Scholar 

  76. Kloeden, P.E., Platen, E.: Numerical Solution of Stochastic Differential Equations. Springer, Berlin (1992)

    Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Leo Dostal.

Ethics declarations

Conflict of interest

The authors declare that they have no conflict of interest.

Additional information

Publisher's Note

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

Appendices

Appendix 1: Derivation of the forced nonlinear Schrödinger equation for time- and space-variant wind-induced pressure

It can be shown that weakly nonlinear solutions of the Euler equations can be reduced to solutions described by a complex envelope, which satisfies the NLS. Such reduction can be achieved by the method of multiple scales, cf. [22]. In the following the multiple scales analysis is presented for the case of forcing of water waves by a time- and space-variant wind field. For the forced Euler equations, an incompressible fluid with density \(\rho _{\mathrm{w}}\), a finite depth h, and a free surface \(\eta (x,t)\) is assumed, where x is the spatial variable and t is time, and the corresponding velocity field is irrotational. This leads to four equations, where the nonlinear gravity wave problem is governed by two linear equations, which are the Laplace equation and the kinematic condition at the sea bed, and by kinematic and dynamic free surface conditions, which are nonlinear [70]. If in addition a time- and space-variant wind-induced pressure \(P_a(x,t)\) at the free surface is assumed, and if dissipative effects are included by means of the water kinematic viscosity \(\nu \), as suggested by Dias et al. [71], then the forced Euler equations become

$$\begin{aligned}&\phi _{xx}+\phi _{zz}=0, \quad \text {for }-h\le z\le \eta (x,t), \end{aligned}$$
(21a)
$$\begin{aligned}&\eta _t+\phi _x\eta _x-\phi _z-2\nu \eta _{xx}=0,\quad \text {for } z=\eta (x,t), \end{aligned}$$
(21b)
$$\begin{aligned}&\phi _t+\frac{1}{2}\left( \phi _x^2+\phi _z^2\right) +g\,\eta =-\frac{1}{\rho _{\mathrm{w}}}P_a-2\nu \phi _{zz},\nonumber \\&\quad \text {for } z=\eta (x,t), \end{aligned}$$
(21c)
$$\begin{aligned}&\phi _{z}=0, \quad \text {for } z=-h, \end{aligned}$$
(21d)

where \(\phi (x,z,t)\) is the velocity potential and g is acceleration due to gravity.

The nonlinear boundary conditions at the free surface \(\eta (x,t)\) make the solution of Eq. (21) very difficult, since the nonlinear boundary \(\eta (x,t)\) is unknown. Thus, one seeks to use simplifications.

Here, the method of multiple scales is used for the derivation of a nonlinear Schrödinger equation resulting from forced Euler Eq. (21). Following Dawey and Stewardson [72] and Hasimoto and Ono [73], the scalings

$$\begin{aligned} \xi =\varepsilon (x-c_{\mathrm{g}}\,t),\quad \tau =\varepsilon ^2\,t \end{aligned}$$
(22)

are used, whereby the small parameter \(\varepsilon \) represents the wave steepness, \(c_{\mathrm{g}}=\dfrac{g}{2\omega }(\tanh kh+kh(1-\tanh ^2kh))\) is the group velocity, \(\omega \) is the frequency, and k is the wave number of the considered carrier wave. The velocity potential \(\phi \) and the surface elevation \(\eta \) are expanded in series of the form

$$\begin{aligned} \phi (x,z,t)&=\sum _{n=1}^\infty \varepsilon ^n\sum _{m=-n}^n \phi ^{n,m}(\xi ,z,\tau )\,E^m, \end{aligned}$$
(23)
$$\begin{aligned} \eta (x,t)&=\sum _{n=1}^\infty \varepsilon ^n\sum _{m=-n}^n \eta ^{n,m}(\xi ,\tau )\,E^m, \end{aligned}$$
(24)

where

$$\begin{aligned}&E=\exp \left( \mathrm {i}(kx-\omega t) \right) ,\quad \phi ^{(n,-m)}={\bar{\phi }}^{(n,m)},\\&\eta ^{(n,-m)} ={\bar{\eta }}^{(n,m)}, \end{aligned}$$

and a bar denotes the complex conjugate. As a next step, the velocity potential at the free surface \(\eta \) is expanded in a Taylor series around \(z=0\)

$$\begin{aligned} \phi (x,\eta ,t)=\sum _{j=0}^\infty \frac{ \eta ^j}{j!} \frac{\partial ^j}{\partial z^j} \phi \bigg |_{z=0} \end{aligned}$$
(25)

Then expansions (23)–(25) are substituted into Eq. (21).

As in [41, 42] the wind-induced pressure \(P_a\) is assumed to be of order \({\mathcal {O}}(\varepsilon ^3)\). Therefore, the wind-induced pressure evaluated at \(z=0\) is expanded as

$$\begin{aligned} P_a(x,t)=\sum _{n=1}^\infty \varepsilon ^{n-1}\sum _{m=-n}^n p^{n,m}(\xi ,\tau )\,E^m. \end{aligned}$$
(26)

and substituted into boundary condition (21c). Then, terms of linear and quadratic order are not affected by wind forcing and the well-known results for \(\phi ^{m,n}(\xi ,z,\tau )\) with \(n\le 2\) can be used [72, 73]. These are given as

$$\begin{aligned} \begin{aligned}&\phi ^{1,1}=\psi \frac{\cosh k(z+h)}{\cosh kh},\\&\phi ^{1,0}_z=0,\quad \phi ^{2,0}_z=0,\\&\phi ^{2,1}=D\frac{\cosh k(z+h)}{\cosh kh}-\mathrm {i} \beta _1 \psi _{\xi }, \\&\phi ^{2,2}=\mathrm {i}\beta _2\frac{\cosh 2k(z+h)}{\cosh (2kh)},\quad \eta ^{1,0}=0, \\&g\eta ^{1,1}= \mathrm {i} \omega \psi , \\&g\eta ^{2,0}=c_{\mathrm{g}} \phi ^{1,0}_{\xi }-k^2(1-\theta ^2)|\psi |^2, \\&g\eta ^{2,1}=\mathrm {i}\omega D+ c_{\mathrm{g}}\psi _{\xi }, \quad g\eta ^{2,2}=\beta _3 \psi ^2, \end{aligned} \end{aligned}$$
(27)

where

$$\begin{aligned} \begin{aligned}&\beta _1= -\frac{(z+h)\sinh (k(z+h))-h\theta \cosh (k(z+h))}{\cosh (kh)},\\&\beta _2= \frac{3k^2(1-\theta ^4(kh))}{4\omega \theta ^2},\; \beta _3=\frac{k^2(\theta ^2-3)}{2\theta ^2},\\&\theta =\tanh (kh), \end{aligned} \end{aligned}$$
(28)

\(\omega \) fulfills the dispersion relation \(\omega =\sqrt{gk\theta }\), and \(\psi \) and D are functions of the slow variables \(\xi \), \(\tau \).

Terms of the velocity potential \(\phi \) of order \(n=3\) and harmonic \(m=0\) in boundary condition (21b) do not involve the wind-induced pressure \(P_a\) or the viscosity \(\nu \) and lead similarly as in [72] to

$$\begin{aligned} \phi _{\xi }^{1,0}= & {} \beta _4|\psi |^2,\; \mathrm {where}\nonumber \\ \beta _4= & {} -k^2\frac{ 2c_p+c_{\mathrm{g}}(1-\theta ^2) }{gh-c_{\mathrm{g}}^2},\quad \phi _{\xi z}^{1,0}=0, \end{aligned}$$
(29)

and \(c_p=\omega /k\) is the phase velocity. From Laplace Eq. (21a) together with bottom boundary condition (21d), \({\phi }^{3,1}\) is obtained as

$$\begin{aligned} \begin{aligned} {\phi }^{3,1}&= \frac{\cosh k(z+h)}{\cosh kh}G+\frac{\beta _1}{2k}\left( 2kh\theta ^2\psi _{\xi \xi }-2\mathrm {i}k D_{\xi } \right) \\&\quad -\,\frac{\left( (z+h)^2-h^2\right) \cosh (k(z+h)}{2 \cosh (kh)}\psi _{\xi \xi }, \end{aligned} \end{aligned}$$
(30)

whereby G is a function of \(\xi \), \(\tau \). This is the same result as in [74] (without surface tension), since no wind-induced pressure is involved in the Laplace equation and the bottom boundary condition.

Collecting terms of the velocity potential \(\phi \) of order \(n=3\) and harmonic \(m=1\) in boundary condition (21b), the equation

$$\begin{aligned} \begin{aligned}&\phi _{z}^{3,1}+\mathrm {i}\omega {\eta }^{3,1}\\&=-\,c_{g}\,\eta _{\xi }^{2,1}+\eta _{\tau }^{1,1}+\mathrm {i}k{ \eta }^{1,1}\phi _{\xi }^{1,0}\\&\quad +\,2\,\nu \,k^2\,{\eta }^{1,1}-{\eta }^{2,0}\phi _{ zz}^{1,1}-{\bar{\eta }}^{1,1}\phi _{zz}^{2,2}\\&\quad -\,\eta ^{2,2}{\bar{\phi }}_{zz}^{1,1}+\left( \eta ^{1,1}\right) ^2 \left( k^2 {\bar{\phi }}_{zz}^{1,1}-\frac{1}{2} {\bar{\phi }}_{zzz}^{1,1}\right) \\&\quad -\,\eta ^{1,1}{\bar{\eta }}^{1,1} \phi _{zzz}^{1,1} \\&\quad +\,2 k^2 {\bar{\eta }}^{1,1}\phi _{z}^{2,2}+2 k^2 {\eta }^{2,2}{\bar{\phi }}_{z}^{1,1} \end{aligned} \end{aligned}$$
(31)

is obtained. Thereby, terms that are zero in Eq. (27) have already been omitted. The corresponding equation for \(n=3\) and \(m=1\) in boundary condition (21c) is given by

$$\begin{aligned} \begin{aligned}&-\mathrm {i}{\phi }^{3,1}\omega +g{\eta }^{3,1}\\&=-{\frac{1}{\rho _{{\mathrm{w}}}}}{p}^{1,1}-2\,\nu \,{\phi }^{1,1}_{zz}\\&\quad -\,\frac{1}{2}\mathrm {i}\omega \,{\eta }^{2,2}{{\bar{\phi }}_{{{ zz}}}}^{1,1} -c_{g}\,{\phi }^{2,1}_{\xi }-{\phi }^{ 1,1}_{\tau }\\&\quad -\,\mathrm {i}k{\phi }^{1,1}\phi _{\xi }^{1,0}-\mathrm {i}\omega \,{\eta }^{2,2}{ {\bar{\phi }}_{{z}}}^{1,1}\\&\quad -\,2\,{k}^{2}{\bar{\phi }}^{1,1}{ \phi }^{2,2} -{k}^{2}{\eta }^{1,1}{\phi _{{z}}}^{1,1}{{\bar{\phi }}}^{1,1}\\&\quad -\,{k}^{2}{\eta }^{1,1}{{\bar{\phi }}_{{z}}}^{1,1}{\phi }^{1,1 }+{k}^{2}{{\bar{\eta }}}^{1,1}\phi _{{z}}^{1,1}{\phi }^{1,1}\\&\quad +\,2\,\mathrm {i} \omega \,{{\bar{\eta }}}^{1,1}\phi _{{z}}^{2,2}\\&\quad +\,\mathrm {i}\omega \,{\eta _{}}^{ 1,1}{{\bar{\eta }}}^{1,1}\phi _{{{ zz}}}^{1,1}-\phi _{{z}}^{ 2,2}{{\bar{\phi }}_{{z}}}^{1,1}\\&\quad +\,\mathrm {i}\omega \,{\eta }^{2,0}\phi _{z}^{1,1}\\&\quad -\,{\eta }^{1,1}\phi _{{z}}^{1,1}{{\bar{\phi }}_{{{ zz}}}}^{1,1}-{\eta }^{1,1}{{\bar{\phi }}_{{z}}}^{1,1}\phi _{{{ zz}}}^{1,1}\\&\quad -\,{{\bar{\eta }}}^{1,1}\phi _{{z}}^{1,1}\phi _{{{ zz}}}^{1,1}, \end{aligned} \end{aligned}$$
(32)

and includes the leading-order time- and space-varying wind-induced pressure term \(p^{1,1}(\xi ,\tau )\). It is also possible to assume \(P_a\) to be of order \({\mathcal {O}}(\varepsilon ^2)\), which involves additional terms due to the wind-induced pressure, as can be seen in [44], where a time- and space-invariant wind velocity is considered. Subtracting (31) from (32) such that the unknown \({\eta }^{3,1}\) vanishes leads finally to the evolution equation for the wave envelope \(\psi (\xi ,\tau )\), which is subjected to a time- and space-variant wind-induced pressure. This equation is given by

$$\begin{aligned} \mathrm {i}\,\psi _{\tau }+\mu \,\psi _{\xi \xi }-\gamma \,|\psi |^2 \psi =-\mathrm {i}\,\frac{p^{1,1}}{2\rho _{{\mathrm{w}}}}-\mathrm {i}\,2\,\nu \,k^2\,\psi , \end{aligned}$$
(33)

where \(\mu =\frac{1}{2}\frac{\partial ^2}{\partial k^2} \omega (k)\), and

$$\begin{aligned} \begin{aligned} \gamma&=\frac{k^4}{4\omega }\left( \frac{9}{\theta ^2}-12+13 \theta ^2-2\theta ^4\right. \\&\quad -\,\frac{2}{gh-c_{\mathrm{g}}^2}\left( 4c_p^2+4c_p c_{\mathrm{g}}(1-\theta ^2)\right. \\&\quad \left. \left. +\,c_{\mathrm{g}}^2(1-\theta ^2)^2 \right) \right) . \end{aligned} \end{aligned}$$

The dissipative term \(-\mathrm {i}\,2\,\nu \,k^2\,\psi \) in Eq. (33) is the same as was for example previously obtained in [75]. It is noted that only the space- and time-variant pressure component \(p^{1,1}(\xi ,\tau )\) from pressure expansion (26) is involved in this equation. For deep water (\(h\rightarrow \infty \)) we get \(\gamma =\frac{2k^4}{\omega }\) and (33) reduces with \(\theta \rightarrow 1\), \(c_{\mathrm{g}} \rightarrow \frac{\omega }{2k}\) and a scaling of \(\psi \rightarrow \frac{\omega }{2k} \psi \) to

$$\begin{aligned}&\mathrm {i}\frac{\omega }{2k}\,\psi _{\tau }-\frac{\omega }{8 k^2}\frac{\omega }{2k}\,\psi _{\xi \xi } -\frac{2k^4}{\omega }\frac{\omega ^3}{8k^3} \,|\psi |^2\psi \nonumber \\&\quad =-\mathrm {i}\,\frac{ p^{1,1}}{2\rho _{{\mathrm{w}}}}-\mathrm {i}\,2\,\nu \,k^2\frac{\omega }{2k}\,\psi \nonumber \\&\quad \Longleftrightarrow \mathrm {i}\,\psi _{\tau }-\frac{\omega }{8 k^2}\,\psi _{\xi \xi }-\frac{1}{2}\omega k^2\,|\psi |^2\psi \nonumber \\&\quad =-\mathrm {i}\,k\,\frac{ p^{1,1}}{\omega \rho _{{\mathrm{w}}}}-\mathrm {i}\,2\,\nu \,k^2\,\psi . \end{aligned}$$
(34)

Appendix 2: Relaxation method for nonlinear gravity waves

Using the relaxation scheme, deterministic NLS (6) is discretized by [68, 69]

$$\begin{aligned} \begin{aligned} \frac{\phi ^{n+1/2}+\phi ^{n-1/2}}{2}&=\frac{1}{2}\omega \,k^2|\psi ^n|^2,\\ \mathrm {i}\frac{\psi ^{n+1}-\psi ^n}{\delta t}&=\frac{\omega }{8k^2} \, \left( \frac{\psi ^{n+1}+\psi ^n}{2}\right) _{xx}\\&\quad +\,\phi ^{n+1/2} \frac{\psi ^{n+1}+\psi ^n}{2}\\&\quad +\,\frac{\Gamma ^{n+1}\,\psi ^{n+1}+\Gamma ^n\,\psi ^n}{2}. \end{aligned} \end{aligned}$$
(35)

Moreover, the space can be discretized by a pseudo-spectral approximation, which leads to a high accuracy, see for example [69]. For the white noise case, where the excitation \(\zeta \) of stochastic NLS (8) is given by a scalar time-varying white noise \(\zeta =\zeta (t)\) as specified in Definition 1, the stochastic integral is approximated in the sense of Stratonovich by

$$\begin{aligned}&\int _{t_n}^{t_{n+1}}\psi (s,x)\circ \mathrm {d}W(s) \nonumber \\&\quad \approx \frac{\psi (t_{n+1},x)+\psi (t_{n},x)}{2}\, ( W_{t_{n+1}}- W_{t_n}), \end{aligned}$$
(36)

where \(\mathrm {d}W(t)=\zeta (t)\, \mathrm {d}t\) is the increment of the standard Wiener process \(W_t\).

We set \({\chi ^{n+1/2}:=( W_{t_{n+1}}- W_{t_n})/\sqrt{\delta t}}\), which are distributed according to the normal distribution \({\mathcal {N}}(0,1)\). Then the relaxation scheme for the NLS, which is excited by white noise in time, reads

$$\begin{aligned} \begin{aligned} \frac{\phi ^{n+1/2}+\phi ^{n-1/2}}{2}&=\frac{1}{2}\omega \,k^2|\psi ^n|^2,\\ \mathrm {i}\frac{\psi ^{n+1}-\psi ^n}{\delta t}&=\frac{\omega }{8k^2} \, \left( \frac{\psi ^{n+1}+\psi ^n}{2}\right) _{xx}\\&\quad +\,\phi ^{n+1/2} \frac{\psi ^{n+1}+\psi ^n}{2}\\&\quad +\,\frac{1}{2\sqrt{\delta t}}\chi ^{n+1/2}(\psi ^{n+1}+\psi ^n). \end{aligned} \end{aligned}$$
(37)

For the above numerical calculations, the GPELab toolbox can be used [68]. In the case, where the NLS is excited by a random wind process in time, we first have to generate the nonwhite wind velocity process \(U(z,t_n)\) by means of a CARMA process, as given in Definition 1. This can be done using for example the Euler–Maruyama scheme [76]. From the randomly time-varying wind velocity process \(U(z,t_n)\) the friction velocity \(u_*(t_n)\) is calculated by means of a fix point iteration at each instant of time using Eq. (9). The resulting random friction velocity \(u_*(t)\) defines the stochastic process \(\Gamma ^n\), which is then substituted into Eq. (35) where

$$\begin{aligned} \Gamma ^n:=\frac{k\omega }{2g}\frac{\rho _{\mathrm{a}}}{\rho _{{\mathrm{w}}}} \,\beta \,\left( \frac{u_*(t_n)}{\kappa }\right) ^2 -2\,\nu \,k^2. \end{aligned}$$
(38)

Appendix 3: CARMA process

A CARMA process is defined as follows, cf. [2, 3].

Definition 1

(CARMA(p,q) process) A CARMA(p,q) process y(t), \(0\le q< p\), is defined as the stationary solution of

$$\begin{aligned} y={\mathbf {c}}\;{\mathbf {u}}(t), \end{aligned}$$
(39)

with the linear differential equation for the state vector \({\mathbf {u}}(t)\in \mathbb {R}^p\)

$$\begin{aligned} {\dot{\mathbf{u}}}(t)=\mathrm {{\mathbf {A}}}\;{\mathbf {u}}(t)+{\mathbf {b}}\; \xi (t), \end{aligned}$$
(40)

where \(\xi (t)\) is white noise with \(E\{\xi (t)\}=0\) and \(E\{\xi (t)\xi (t+\tau )\}=\sigma ^2\delta (\tau )\), \(\;\sigma \in \mathbb {R}\), \(\delta (\cdot )\) is the Dirac function,

$$\begin{aligned} \begin{aligned} \mathrm {{\mathbf {A}}}&=\left( \begin{matrix} -a_1&{}\quad 1 &{}\quad 0 &{}\quad \cdots &{}\quad 0\\ -a_2&{}\quad 0 &{}\quad 1 &{}\quad \ddots &{}\quad \vdots \\ \vdots &{}\quad \vdots &{}\quad \ddots &{}\quad \ddots &{}\quad 0\\ -a_{p-1}&{}\quad 0 &{} \quad \cdots &{}\quad 0&{}\quad 1 \\ -a_p&{}\quad 0 &{}\quad \cdots &{}\quad \cdots &{}\quad 0 \end{matrix}\right) ,\\ {\mathbf {b}}&=\left( \begin{array}{c} b_{p-1}\\ b_{p-2}\\ \vdots \\ b_1\\ b_0\\ \end{array}\right) ,\;\;\; {\mathbf {c}}= \left( \begin{array}{c} 1\\ 0\\ \vdots \\ 0\\ \end{array}\right) , \end{aligned} \end{aligned}$$
(41)

and \(b_j=0\) if \(j>q\).

Equation (40) can be expressed as the stochastic differential equation

$$\begin{aligned} \mathrm {d}{\mathbf {u}}(t)=\mathrm {{\mathbf {A}}}\,{\mathbf {u}}(t) \,\mathrm {d}t+{\mathbf {b}}\, \mathrm {d}W(t), \end{aligned}$$
(42)

using the relation \(\mathrm {d}W(t)=\xi (t)\, \mathrm {d}t \) between the white noise \(\xi (t)\) and the increment of the Wiener process \(\mathrm {d}W(t)\). The above-defined CARMA(p,q) process has the transfer function

$$\begin{aligned} H_{\mathrm {carma}}(s)=\frac{b_0+b_1s+\cdots +b_qs^q}{s^p+a_1 s^{p-1}+\cdots +a_p}, \end{aligned}$$
(43)

where \(s:=\mathrm {i}\omega \), and the spectral density

$$\begin{aligned} S_{\mathrm {carma}}(\omega )=\sigma ^2\frac{|b_0+b_1s+\cdots +b_qs^q|^2}{|s^p+a_1 s^{p-1}+\cdots +a_p|^2}. \end{aligned}$$
(44)

Thereby, \(\omega \) is the angular frequency. It should be noted that the state space representation in Definition 1 is not unique.

Appendix 4: Forces on a cylindrical pile

In order to compute the wave-induced hydrodynamic force on a vertical cylindrical pile, the velocity and acceleration of water particles have to be computed. This can be done by determining the corresponding velocity potential. Given the complex-valued wave envelope \(\psi =\psi (x,t)\), obtained from the solution of NLS (6), the velocity potential at a given depth z was computed by Carter et al. [47]

$$\begin{aligned} \begin{aligned} \phi (x,z,t)&=\biggl \lbrace \frac{\mathrm {i} \varepsilon \omega }{k} {\hat{B}} +\frac{\varepsilon ^2 \omega }{2k^2} {\hat{B}}_X + \varepsilon ^3 \\&\quad \left( -\frac{\mathrm {i} k \omega }{2} |{\hat{B}}|^2 {\hat{B}} - \frac{3 \mathrm {i} \omega }{8 k^3} {\hat{B}}_{XX} \right) \biggl \rbrace e ^{\mathrm {i} \omega t - \mathrm {i} kx + kz} \\&\quad +\, {\mathcal {O}}(\varepsilon ^4) + c.c., \end{aligned} \end{aligned}$$
(45)

where \({\hat{B}}=B(X+\mathrm {i}Z,T)\), \(X=\varepsilon x\), \(Z=\varepsilon z\), \(T=\varepsilon ^2 t\) and c.c. represents the complex conjugate. The velocity u in x-direction, velocity v in z-direction, and acceleration a in x-direction can then be computed by

$$\begin{aligned} u(x,z,t)= & {} \frac{\partial \phi }{\partial x} (x,z,t),\quad v(x,z,t)=\frac{\partial \phi }{\partial z} (x,z,t),\nonumber \\ a(u,z,t)= & {} \frac{\partial u}{\partial t} (x,z,t). \end{aligned}$$
(46)

With that, the force \(d F_x\) acting on a part \(d z\) of a vertical cylindrical pile with diameter D can be computed by means of the Morison equation and reads [48]

$$\begin{aligned} d F_x=C_{\mathrm {m}}\rho _{\mathrm{w}} \frac{\pi D^2}{4} d z a + C_{\mathrm{d}} \frac{\rho _{\mathrm{w}}}{2} D d z |u|u, \end{aligned}$$
(47)

whereby \(C_{\mathrm {m}}\) is the inertia coefficient, \(C_{\mathrm{d}}\) is the drag coefficient, and \(\rho _{\mathrm{w}}\) is the density of the incompressible fluid.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Dostal, L., Hollm, M. & Kreuzer, E. Study on the behavior of weakly nonlinear water waves in the presence of random wind forcing. Nonlinear Dyn 99, 2319–2338 (2020). https://doi.org/10.1007/s11071-019-05416-5

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11071-019-05416-5

Keywords

Navigation