Skip to main content
Log in

Evolution of Spectral Distributions in Deep-Water Constant Vorticity Flows

  • Original Article
  • Published:
Water Waves Aims and scope Submit manuscript

Abstract

A central question in sea-state modeling is the role that various physical effects have on the evolution of the statistical properties of random sea states. This becomes a critical issue when one is concerned with the likelihood of rare events such as rogue, or freak, waves which can have significant destructive potential on deep sea ships and other offshore structures. In this paper then, using a recently derived higher-order model of deep water nonlinear waves, we examine the impact of constant vorticity currents on the statistical properties of nonlinearly evolving random sea states. As we show, these currents can both decrease and increase the kurtosis of the affiliated distributions of the sea states, thereby diminishing or enhancing the likelihood of rare events. We likewise numerically study the relationship between the kurtosis and a non-dimensional parameter, the Benjamin–Feir Index, which has proven to be a useful measure of when rare events are likely in oceanographic application.

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

Similar content being viewed by others

References

  1. Dysthe, K.B., Trulsen, K., Krogstad, H.E., Socquet-Juglard, H.: Evolution of a narrow-band spectrum of random surface gravity waves. J. Fluid Mech. 478, 1–10 (2003)

    Article  MathSciNet  Google Scholar 

  2. Socquet-Juglard, H., Dysthe, K.B., Trulsen, K., Krogstad, H.E.: Probablity distributions of surface gravity waves during spectral changes. J. Fluid Mech. 510, 195–216 (2005)

    Article  Google Scholar 

  3. Onorato, M., Osborne, A.R., Serio, M., Cavaleri, L., Brandini, C., Stansberg, C.T.: Extreme waves, modulational instability and second order theory: wave flume experiments on irregular waves. Eur. J. Mech. B Fluids 22, 586–601 (2006)

    Article  MathSciNet  Google Scholar 

  4. Thomas, R., Kharif, C., Manna, M.: A nonlinear Schrödinger equation for water waves on finite depth with constant vorticity. Phys. Fluids 24(12), 127102 (2012)

    Article  Google Scholar 

  5. Janssen, P.A.E.M.: Nonlinear four-wave interactions and freak waves. J. Phys. Ocean 33, 863–884 (2003)

    Article  MathSciNet  Google Scholar 

  6. Hasselmann, K.: On the non-linear energy transfer in a gravity-wave spectrum. Part 1. General theory. J. Fluid Mech. 12, 481–500 (1962)

    Article  MathSciNet  Google Scholar 

  7. Holthuijsen, L.: Waves in Oceanic and Coastal Waters. Cambridge University Press, Cambridge (2007)

    Book  Google Scholar 

  8. Shemer, L., Sergeeva, A.: An experimental study of spatial evolution of statistical parameters in a unidirectional narrow-banded random wavefield. J. Geo. Res. 114, C01015 (2009)

    Article  Google Scholar 

  9. Shemer, L., Slunyaev, A., Sergeeva, A.: Applicability of envelope model equations for simulation of narrow-spectrum unidirectional random field evolution: Experimental validation. Phys. Fluids 22, 016601 (2010)

    Article  Google Scholar 

  10. Shemer, L., Sergeeva, A., Liberzon, D.: Effect of the initial spectrum on the spatial evoluion of statistics of unidirectional nonlinear random waves. J. Geo. Res. 115, C12039 (2010)

    Article  Google Scholar 

  11. Slunyaev, A.V., Sergeeva, A.V.: Stochastic simulation of unidirectional intense waves in deep water applied to rogue waves. JETP Lett. 94, 779–786 (2011)

    Article  Google Scholar 

  12. Fedele, F., Brennan, J., De León, S.P., Dudley, J., Dias, F.: Real world ocean rogue waves explained without the modulational instability. Sci. Rep. 6, 2045–2322 (2016)

    Article  Google Scholar 

  13. Alber, I.E.: The effects of randomness on the stability of two-dimensional surface wavetrains. Proc. Roc. Soc. A 363, 525–546 (1978)

    MathSciNet  MATH  Google Scholar 

  14. Alber, I.E., Saffman, P.G.: Stability of random nonlinear deep water waves with finite bandwidth spectra. Technical report, TRW Defense and Space System Group (1978)

  15. Slunyaev, A., Sergeeva, A., Pelinovsky, E.: Wave amplification in the framework of forced nonlinear Schrödinger equation: the rogue wave context. Phys. D 303, 18–27 (2015)

    Article  MathSciNet  Google Scholar 

  16. Eeltink, D., Armaroli, A., Ducimetière, Y.M., Kasparian, J., Brunetti, M.: Single-spectrum prediction of kurtosis of water waves in a nonconservative model. Phys. Rev. E 100(1), 013102 (2019)

    Article  Google Scholar 

  17. Curtis, C.W., Carter, J.C., Kalisch, H.: Particle paths in nonlinear Schrödinger models in the presence of linear shear currents. J. Fluid Mech. 255, 322–350 (2018)

    Article  Google Scholar 

  18. Touboul, J., Kharif, C.: Effect of viscosity on the generation of rouge waves due to dispersive focusing. Nat. Hazards 84, 585–598 (2016)

    Article  Google Scholar 

  19. Ashton, A.C.L., Fokas, A.S.: A non-local formulation of rotational water waves. J. Fluid Mech. 689, 129–148 (2011)

    Article  MathSciNet  Google Scholar 

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

    MATH  Google Scholar 

Download references

Acknowledgements

The authors would like to thank the support of the NSF through DMS-1715039. Likewise, we appreciate the many conversations and constructive suggestions of John Carter. We likewise thank the anonymous reviewers for their constructive questions and suggestions.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Christopher W. Curtis.

Ethics declarations

Conflict of interest

On behalf of all authors, the corresponding author states that there is no conflict of interest.

Additional information

Publisher's Note

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

Appendix

Appendix

1.1 Shear Profiles over Infinitely Deep Fluids

As noted in the text, the assumption that the fluid velocity is to leading order given by

$$\begin{aligned} \mathbf{u} \approx \omega z \hat{\mathbf{i}} \end{aligned}$$

is clearly unrealistic insofar as it leads to currents of infinite speed as one descends through the fluid. A more realistic, though also more complicated, ansatz is to suppose that to leading order we have that \(\mathbf{u}\approx u(z)\hat{\mathbf{i}}\) where

$$\begin{aligned} u(z) = \left\{ \begin{array}{rl} \omega z, &{} -h_{0} \le z< 0 \\ \frac{-\omega h_{0}}{h_{1}-h_{0}} (z+h_{1}), &{} -h_{1} \le z< -h_{0} \\ 0, &{} z < -h_{1}, \end{array} \right. \end{aligned}$$

where \(h_{1}>h_{0}\gg 1\), so that we are looking at a deep, continuous shear profile which is zero at or near the surface \(z=\epsilon \eta (x,t)\) and at the depth \(z=-\,h_{1}\), after which the fluid is to leading order quiescent. Note, we can also see this profile as satisfying to leading order two ‘no-slip’ conditions, one near the free surface at \(z=0\) and one near \(z=-\,h_{1}\).

While a full, nonlinear description of the above shear profile would require two more freely evolving interfaces, and thus is beyond the scope of this paper, we can readily find the dispersion relationship affiliated with this profile. This then allows us to provide some analytic argument for why we study an otherwise unphysical velocity profile in the main body of the text. Likewise, information from the dispersion relationship provides us with a better understanding of how depth varying shear profiles induce both surface and internal waves.

Following relatively classical approaches, we introduce three fluid velocities:

$$\begin{aligned} \mathbf{u}_{1}&= \omega z \hat{\mathbf{i}} + \epsilon \nabla \phi _{1},\qquad \qquad \qquad \qquad \,\, ~-h_{0}+\epsilon \eta _{2}< z< \epsilon \eta _{1}\\ \mathbf{u}_{2}&= \frac{-\omega h_{0}}{h_{1}-h_{0}} (z+h_{1}) \hat{\mathbf{i}} + \epsilon \nabla \phi _{2},\quad ~-h_{1}+\epsilon \eta _{3}< z< -h_{0}+\epsilon \eta _{2}\\ \mathbf{u}_{3}&= \epsilon \nabla \phi _{3},\qquad \qquad \qquad \qquad \qquad \qquad \,\, ~z < -h_{1} + \epsilon \eta _{3} \end{aligned}$$

so that after linearizing around the small disturbances, we derive the dispersion relationship

$$\begin{aligned} \frac{gk}{\Omega } {+} \omega -\Omega \tanh (kh_{0}) {+} \left( \frac{1}{\Omega }\left( \frac{gk}{\Omega } {+} \omega \right) \tanh (kh_{0}) - 1\right) \left( \frac{\omega h_{1}}{\delta h} {+} \Omega \Omega _{1}(k\delta h)\right) = 0, \end{aligned}$$

where

$$\begin{aligned} \Omega _{1}(k\delta h) = \frac{s\left( \Omega -\frac{\omega h_{0}}{\delta h}s\right) +\Omega \tanh (k\delta h)}{\Omega +s\left( \Omega - \frac{\omega h_{0}}{\delta h}s\right) \tanh (k\delta h)}, \end{aligned}$$

and where \(\delta h = h_{1}-h_{0}\) and \(s=\text{ sgn }(k)\). While in general we would have to find the roots of a fifth-order polynomial to determine the values of \(\Omega \), we see for \(h_{1}\gg h_{0}\) that \(\Omega _{1}\sim s\), and thus we get that the dispersion relationship simplifies to

$$\begin{aligned} \frac{gk}{\Omega } + \omega -\Omega \tanh (kh_{0}) + \left( \frac{1}{\Omega }\left( \frac{gk}{\Omega } + \omega \right) \tanh (kh_{0}) - 1\right) \left( \tilde{\omega } + s\Omega \right) \sim 0, \end{aligned}$$

where

$$\begin{aligned} \tilde{\omega } = \frac{\omega h_{1}}{\delta h}. \end{aligned}$$

Since \(\tilde{\omega }\) only approaches \(\omega \) at an algebraic rate, it seems appropriate to keep it included in the analysis. Letting

$$\begin{aligned} \left| \tanh (kh_{0}) \right| = 1 - \tilde{\epsilon }, \end{aligned}$$

and noting that \(\tilde{\epsilon }\) vanishes to zero exponentially fast as we increase \(h_{0}\), we see the reduced dispersion relationship factors into the form

$$\begin{aligned} \left( 2\Omega + s\tilde{\omega } \right) \left( g\left| k\right| + s\omega \Omega - \Omega ^{2} \right) + \tilde{\epsilon }\left( \Omega ^{3} - (s\tilde{\omega } + \Omega )(g|k|+\omega s\Omega ) \right) \sim 0. \end{aligned}$$

Thus, we have a completely regular perturbation problem for the roots, which we readily see are given by

$$\begin{aligned} \Omega \sim \frac{1}{2}\left( -\,s\omega \pm \sqrt{\omega ^{2}+4g|k|} \right) , ~ -\frac{s\tilde{\omega }}{2}. \end{aligned}$$

Thus, of the three roots we find, two give the dispersion relationship we find in the body of the text using \(\mathbf{u}\sim \omega z \hat{\mathbf{i}}\). Looking at the affiliated disturbances of the relevant free surface and internal wave, we find that

$$\begin{aligned} \eta _{1} \sim \alpha _{12}\frac{ik}{\Omega }e^{i\theta } + \text{ cc }, ~ \eta _{2} \sim \alpha _{12}\frac{ik}{\Omega ^{3}}\cosh (kh_{0})\left( \Omega ^{2}-(1-\tilde{\epsilon })\left( g|k|+s\omega \Omega \right) \right) e^{i\theta }+ \text{ cc }. \end{aligned}$$

Thus, if we choose \(\Omega \) so that \(\Omega ^{2}-\left( g|k|+s\omega \Omega \right) = 0\), the magnitude of the internal wave essentially vanishes, thereby localizing dynamics along the free surface near \(z=0\). Therefore, while not necessarily physically justifiable throughout the bulk of the fluid, our simplified shear profile assumption produces results which are asymptotically consistent with a more sophisticated treatment of the shear profile.

We note though that the more physically realistic shear profile shows that there is a choice of \(\Omega \) which corresponds to a non-trivial internal mode near \(z=-h_{0}\). We also note that this choice of \(\Omega \) makes many of the terms in our modulation theory singular, thus showing that this case would require a markedly different treatment. While interesting, this issue is beyond the scope of the current paper and will be addressed in future research.

1.2 Higher Order-Dispersive Corrections to the VDE

To better understand why we must include the higher order corrections to the dispersion used in the VDE, we examine how such terms are found from first principles. To leading order from Eq. (1) over an infinitely deep fluid, we have that

$$\begin{aligned} \eta _{t} = -\mathcal {H}Q, \end{aligned}$$

where \(q(x,t) = \phi (x,\eta (x,t),t)\), \(Q = q_{x}\), and \(\mathcal {H}\) is the Hilbert transform. Likewise, from Eq. (2) we have to leading order that

$$\begin{aligned} Q _{t}+ \omega \eta _{t} + \eta _{x}= 0, \end{aligned}$$

so that by combining the two expressions we have the wave equation in \(\eta \) alone given by

$$\begin{aligned} \eta _{tt}- \omega \mathcal {H}\eta _{t} - \mathcal {H}\eta _{x}= 0. \end{aligned}$$

Using Fourier transforms, we can readily write a solution to the affiliated initial-value problem in the form

$$\begin{aligned} \eta (x,t) = \frac{1}{2\pi }\int _{\mathbb {R}}\hat{\eta }_{0}(k) e^{ikx+i\Omega _{+}(k)t} \mathrm{d}k + \frac{1}{2\pi }\int _{\mathbb {R}}\hat{\eta }_{0}(k) e^{ikx+i\Omega _{-}(k)t} \mathrm{d}k, \end{aligned}$$

where

$$\begin{aligned} \Omega _{\pm }(k) = \frac{1}{2}\left( \omega s \pm \sqrt{\omega ^{2}+4|k|} \right) . \end{aligned}$$

If we assume, as per usual when deriving NLS-type models, that

$$\begin{aligned} \hat{\eta }_{0}(k) = \frac{1}{\epsilon }\hat{A}\left( \frac{k-k_{0}}{\epsilon } \right) , \end{aligned}$$

representing the assumption that the initial conditions are narrowly banded around a carrier wave number \(k_{0}\), then we see that, taking only the plus branch of \(\Omega _{\pm }\),

$$\begin{aligned} \eta (x,t) =&\frac{e^{i\theta (x,t,k_{0})}}{2\pi }\int _{\mathbb {R}}\hat{A}(\tilde{k})e^{i\tilde{k}\xi } e^{i\tau \tilde{\Omega }(k_{0},\tilde{k})}\mathrm{d}\tilde{k}\\ \sim&\frac{e^{i\theta (x,t,k_{0})}}{2\pi }\int _{\mathbb {R}}\hat{A}(\tilde{k})e^{i\tilde{k}\xi }e^{i\tau \left( \frac{\Omega ''(k_{0})}{2}\tilde{k}^{2} + \frac{\epsilon }{6}\Omega '''(k_{0})\tilde{k}^{3} + \frac{\epsilon ^{2}}{24}\Omega ''''(k_{0})\tilde{k}^{4}\right) }\mathrm{d}\tilde{k}, \end{aligned}$$

where

$$\begin{aligned} \theta (x,t,k_{0}) = k_{0}x + \Omega (k_{0})t, ~c_{g}(k_{0}) = \Omega '(k_{0}), ~ \xi = \epsilon (x+c_{g}t), ~ \tau = \epsilon ^{2}t, \end{aligned}$$

and where

$$\begin{aligned} \tilde{\Omega }(k_{0},\tilde{k}) = \frac{1}{\epsilon ^{2}}\left( \Omega (k_{0}+\epsilon \tilde{k})-\Omega (k_{0})-\Omega '(k_{0})\epsilon \tilde{k}\right) \end{aligned}$$

and where we have expanded the dispersion relationship up to \(\mathcal {O}(\epsilon ^{3})\) as is done in the main body of the text. This expansion corresponds to the affiliated linear-evolution equation for the slowly evolving envelope \(A(\xi ,\tau )\)

$$\begin{aligned} i\partial _{\tau }A = \left( \frac{\Omega ''(k_{0})}{2}\partial _{\xi }^{2}-\frac{i\epsilon }{6}\Omega '''(k_{0})\partial _{\xi }^{3} - \frac{\epsilon ^{2}}{24}\Omega ''''(k_{0})\partial _{\xi }^{4}\right) A. \end{aligned}$$

As to the issue of why to include the \(\epsilon ^{2}\) term, we note that the affiliated group velocity of the slowly evolving envelope is given by

$$\begin{aligned} \frac{\mathrm{d}}{\mathrm{d}\tilde{k}}\tilde{\Omega } =&\frac{1}{\epsilon }\left( \Omega '(k_{0}+\epsilon \tilde{k}) - \Omega '(k_{0})\right) \\ \sim&\Omega ''(k_{0})\tilde{k}\left( 1 - \frac{3\epsilon c_{g}^{2}}{s}\tilde{k} + 10 \epsilon ^{2}c_{g}^{4}\tilde{k}^{2}\right) . \end{aligned}$$

We see that the full dispersion relationship has exactly one root at \(\tilde{k}=0\), so that this root is the only point of stationary phase in the corresponding integral representation of the solution \(\eta (x,t)\). However, using our Taylor series expansion, were we to ignore the \(\epsilon ^{2}\) term, we would introduce a new stationary-phase point at \(\tilde{k}_{*}\), where

$$\begin{aligned} \tilde{k}_{*} = \frac{s}{3\epsilon c_{g}^{2}}. \end{aligned}$$

The presence of this stationary-phase point allows for the accumulation of energy at high frequencies, which without a corresponding band-limiting requirement introduces non-physical effects into the model. By including the \(\epsilon ^{2}\) term though, we remove this spurious high-frequency stationary-phase point, thereby avoiding the inclusion of an otherwise unnecessary bandwidth condition on the slowly evolving envelope.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Curtis, C.W., Murphy, M. Evolution of Spectral Distributions in Deep-Water Constant Vorticity Flows. Water Waves 2, 361–380 (2020). https://doi.org/10.1007/s42286-020-00033-x

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s42286-020-00033-x

Keywords

Navigation