Skip to main content
Log in

Particle motions beneath irrotational water waves

  • Published:
Ocean Dynamics Aims and scope Submit manuscript

Abstract

Neutral and buoyant particle motions in an irrotational flow are investigated under the passage of linear, nonlinear gravity, and weakly nonlinear solitary waves at a constant water depth. The developed numerical models for the particle trajectories in a non-turbulent flow incorporate particle momentum, size, and mass (i.e., inertial particles) under the influence of various surface waves such as Korteweg-de Vries waves which admit a three parameter family of periodic cnoidal wave solutions. We then formulate expressions of mass-transport velocities for the neutral and buoyant particles. A series of test cases suggests that the inertial particles possess a combined horizontal and vertical drifts from the locations of their release, with a fall velocity as a function of particle material properties, ambient flow, and wave parameters. The estimated solutions exhibit good agreement with previously explained particle behavior beneath progressive surface gravity waves. We further investigate the response of a neutrally buoyant water parcel trajectories in a rotating fluid when subjected to a series of wind and wave events. The results confirm the importance of the wave-induced Coriolis-Stokes force effect in both amplifying (destroying) the pre-existing inertial oscillations and in modulating the direction of the flow particles. Although this work has mainly focused on wave-current-particle interaction in the absence of turbulence stochastic forcing effects, the exercise of the suggested numerical models provides additional insights into the mechanisms of wave effects on the passive trajectories for both living and nonliving particles such as swimming trajectories of plankton in non-turbulent flows.

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
Fig. 18

Similar content being viewed by others

References

  • Bakhoday-Paskyabi M (2014) Small-scale turbulence dynamics under sea surface gravity wave. University of Bergen, Geophysical Institute, http://hdl.handle.net/1956/8692

  • Bakhoday-Paskyabi M, Fer I (2014) Turbulence structure in the upper ocean: a comparative study of observations and modelling. Ocean Dyn. doi:10.1007/s10236--014--0697--6

  • Bakhoday-Paskyabi M, Fer I, Jenkins AD (2012) Surface gravity wave effects on the upper ocean boundary layer: Modification of a one–dimensional vertical mixing model. Cont Shelf Res 38:63–78

    Article  Google Scholar 

  • Borluk H, Kalisch H (2012) Particle dynamics in the KdV approximation. Wave Mot 49:691–709

    Article  Google Scholar 

  • Carter DJT (1982) Prediction of wave height and period for a constant wind velocity using the JONSWAP results. Ocean Eng 9:17–33

    Article  Google Scholar 

  • Constantin A (2006) The trajectories of particles in Stokes waves. Invent math 166:523–535

    Article  Google Scholar 

  • Constantin A, Escher J (2007) Particle trajectories in solitary water waves. Bull Amer Math Soc (N.S.) 44:423–431

    Article  Google Scholar 

  • Constantin A, Strauss W (2010) Pressure beneath a Stokes wave. Comm Pure Appl Math 63:533–557

    Google Scholar 

  • Craik ADD, Leibovich S (1976) A rational model for Langmuir circulations. J Fluid Mech 73:401–426

    Article  Google Scholar 

  • D’Asaro EA (2001) Turbulent vertical kinetic energy in the ocean mixed layer. J Phys Oceanogr 31:3530–3537

    Article  Google Scholar 

  • Dean RG, Dalrymple RA (1992) Water wave mechanics for engineers and scientists. Prentice-Hall Inc, Englewood Cliffs

  • Eames I (2008) Settling of particles beneath water waves. J Phys Oceanogr 38:2846–2008

    Article  Google Scholar 

  • Gerstner FV (1802) Theory of waves. Abhandlungen der Koenigl, boehmischen Gesellschaft der Wissenschaften zu Prag

  • Hasselmann K (1970) Wave-driven inertial oscillations. Geophys Fluid Dyn 1:463–502

    Article  Google Scholar 

  • Longuet-Higgins M (1953) Mass transport in water waves. Philos Trans Roy Soc London 245:535–581

    Article  Google Scholar 

  • Longuet-Higgins M (1986) Eulerian and lagrangian aspects of surface waves. J Fluid Mech 173:683–707

    Article  Google Scholar 

  • Makin VK, Kudryavtsev VN (1999) Coupled sea surface-atmosphere model 1. Wind over waves coupling. J Geophy Res 15:7613– 7623

    Article  Google Scholar 

  • Maxey MR, Riley JJ (1983) Equation of motion for a small rigid sphere in a nonuniform flow. Phys. Fluids 26.4:883–889

    Article  Google Scholar 

  • McWilliams JC, Restrepo JM (1999) The wave driven ocean circulation. J Phys Oceanogr 29:2523–2540

    Article  Google Scholar 

  • Melville WK, Matusov P (2002) Distribution of breaking waves at the ocean surface. Nature 417:58–63

    Article  Google Scholar 

  • Osborn TR, Farmer DM, Vagle S, Thorpe SA, Cure M (1992) Measurements of bubble plumes and turbulence from a submarine. Atmos Ocean 30:419–440

    Article  Google Scholar 

  • Osborne AR, Kirwan AD, Provenzale A, Bergamasco L (1986) The Korteweg-de Vries equation in Lagrangian coordinates. Phys Fluids 29:656–660

    Article  Google Scholar 

  • Phillips OM (1958) The equilibrium range in the spectrum of wind-generated waves. J Fluid Mech 4:426–434

    Article  Google Scholar 

  • Santamaria F, Boffetta G, Afonso MM, Mazzino A, Onorato M, Pugliese D (2013) Stokes drift for inertial particles transported by water waves. EPL 102. doi:10.1209/0295--5075/102/14003

  • Stokes G (1847) On the theory of oscillating waves. Trans Cambridge Philos Soc 8:441–455

    Google Scholar 

  • Sullivan PP, McWilliams JC (2010) Dynamics of winds and currents coupled to surface waves. Annu Rev Fluid Mech 42:19– 42

    Article  Google Scholar 

  • Thorpe SA, Humphries P (1980) Bubbles and breaking waves. Nature 283:463–465

    Article  Google Scholar 

  • Umeyama M (2012) Eulerian-Lagrangian analysis for particle velocities and trajectories in a pure wave motion using particle image velocimetry. Philos Trans R Soc Lond Ser A Math Phys Eng Sci 370:1687–1702

    Article  Google Scholar 

  • Whitham GB (1974) Linear and Nonlinear Waves. Wiley, New York

    Google Scholar 

  • Wiegel RL (1960) A presentation of cnoidal wave theory for practical application. J Fluid Mech 7:273–286

    Article  Google Scholar 

Download references

Acknowledgments

This work has been performed as part of the Norwegian Center for Offshore Wind Energy (NORCOWE) funded by the Research Council of Norway (RCN 1938211560) and by OBLO (Offshore Boundary Layer Observatory) project (RCN: 227777). I also express my gratitude to Hans Burchard and two anonymous reviewers for their insightful comments.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Mostafa Bakhoday-Paskyabi.

Additional information

Responsible Editor: Bruno Castelle

Appendices

Appendix A: Derivation of the KdV Eq. 10

The KdV equation describes the propagation of weakly nonlinear water waves in a dispersive media. The equation is derived from the Euler equations under the assumptions that the amplitude is small compared to the depth and the depth is small compared to the wavelength. Although rich mathematical studies on different aspects of this equation have been carried out, for the matter of completeness, we present the derivation of KdV equation using the approach proposed by Whitham (1974).

To provide governing equations more tractable, and to split them into different regions of interest, we take the standard step of nondimensionalization as follows

$$ x = L\hat{x}; z = H\hat{z}; t =\frac{L}{c_{0}}\hat{t}; \varphi=aL\sqrt{\frac{g}{H}}\hat{\varphi}; \eta=a\hat{\eta}\-, $$
(29)

where \(c_{0}=\sqrt {gH}\), a is wave amplitude, and L denotes the characteristic wavelength. Using the definition of a potential function, and substituting Eq. 29 into the Euler Eqs. 13, and then dropping the hats, we obtain the following nondimensionalized equations:

$$\begin{array}{@{}rcl@{}} \delta \frac{\partial^{2} \varphi}{\partial x^{2}} + \frac{\partial^{2} \varphi}{\partial z^{2}}&=& 0, z\in (-1,1+\varepsilon \eta)\-, \end{array} $$
(30)
$$\begin{array}{@{}rcl@{}} \eta_{t}+\varepsilon\varphi-\frac{1}{\delta}\frac{\partial \varphi}{\partial z} &=& 0, z=1+\varepsilon\eta\-, \end{array} $$
(31)
$$\begin{array}{@{}rcl@{}} \eta+\frac{\partial \varphi}{\partial z}+\frac{1}{2}\varepsilon\frac{\varepsilon}{\delta}\frac{\partial^{2} \varphi}{\partial z^{2}} &=& 0, z=1+\varepsilon\eta\-, \end{array} $$
(32)
$$\begin{array}{@{}rcl@{}} \frac{\partial \varphi}{\partial z} &=& 0, z=-1\-. \end{array} $$
(33)

Without loss of generality, we neglected the effect of pressure at the free surface in Eqs. 3132. We also applied the following dimensionless parameters:

$$ \varepsilon=\frac{a}{H}, \text{and}~\delta=\frac{H^{2}}{L^{2}}\-, $$
(34)

where ε gives the measure of nonlinearity, and δ describes the limit of shallowness. For weak nonlinearity assumption, we should have ε≪1 (but not zero), and for shallowness assumption, we should have δ≪1. In the case that the depth is small compared to the wave characteristic length scale, we can expand the nondimensional potential function φ in terms of δ without imposing any assumption on ε:

$$\varphi=\sum\limits_{n} \varphi_{n}(x,t)(z\delta)^{n}\-. $$

By substituting this series into Eq. 33, and since velocity vanishes at z=−1, we get φ 1=0. By recurrence, it results in φ 2n+1=0. Consequently,

$$ \varphi=\sum\limits_{n} (-1)^{n} \frac{\delta^{n}z^{2n}}{(2n)!}\frac{\partial^{2n}\varphi_{0}(x,t)}{\partial x^{2n}}\-. $$
(35)

By substituting Eq. 35 into the Eqs. 3132, we obtain the following estimates up to order \(\mathcal {O} (\varepsilon ^{2},\delta \varepsilon ,\delta ^{2})\):

$$ \eta_{t}+\frac{\partial^{2} \phi_{0}}{\partial x^{2}}+\varepsilon \left( \eta\frac{\partial\phi_{0}}{\partial x}\right)_{x}-\frac{1}{6}\delta\left( \frac{\partial \phi_{0}}{\partial x}\right)_{xxx}= 0\-, $$
(36)
$$ \eta+\frac{\partial \phi_{0}}{\partial t}+\frac{1}{2}\delta\left( \frac{\partial \phi_{0}}{\partial x}\right)^{2}-\frac{1}{2}\delta\frac{\partial^{4} \phi_{0}}{\partial x^{3}\partial t}= 0\-. $$
(37)

Furthermore, using Eq. 33, the lowest order term in the power series satisfies (φ 0) z =0, for all z, meaning that the horizontal velocity components at the bottom (z=−1) can be determined independent of the vertical coordinate z. Therefore, using prior relations and defining f(x,t)= x φ 0, the horizontal and vertical components of velocity at the bottom are determined as follows

$$\begin{array}{@{}rcl@{}} u(x,t)&=&f(x,t)-\frac{1}{2}z^{2}\delta f_{xx}(x,t)\-, \end{array} $$
(38)
$$\begin{array}{@{}rcl@{}} w(x,t)&=&-zf_{x}(x,t)+\frac{1}{2}z^{3} f_{xx}(x,t)\-. \end{array} $$
(39)

Following Whitham (1974), f can be estimated as a function of η:

$$f(x,t)=\eta-\frac{1}{4}\varepsilon\eta^{4}+\frac{1}{4}\varepsilon\eta_{xx}+\mathcal{O} (\varepsilon^{2},\delta\varepsilon,\delta^{2})\-. $$

Hence, the nondimentionlized KdV equation, Eq. 36, can be written using above relations as

$$\eta_{t}+\eta_{x}+\frac{3}{2}\varepsilon\eta\eta_{x}+\frac{1}{6}\delta\eta_{xxx}+\mathcal{O} (\varepsilon^{2},\delta\varepsilon,\delta^{2})\-. $$

This expression is then converted to the physical dimensional space:

$$\eta_{t}+c_{0}\left( 1+\frac{3}{2}\frac{\eta}{H}\right)\eta_{x}+\frac{1}{6}c_{0}H^{2}\eta_{xxx}=0\-. $$

To evaluate the horizontal and vertical velocity components at level 𝜃 H, where 0≤𝜃≤1 (1314), we use Taylor expansion about 𝜃 H up to third order as \(f(x,t)=f^{\theta }+\frac {1}{2}(\theta H)^{2}f_{xx}\) by using Eqs. 38 and 39.

Appendix B: Estimation of the Jacobian elliptic parameter

Determination of the Jacobian elliptic parameter is a crucial step in cnoidal wave theory. Here, we briefly give a recipe on how to determine m accurately.

In cnoidal waves, the relationships between associated parameters are given by

$$\begin{array}{@{}rcl@{}} \lambda &=& H\sqrt{\frac{16mH}{3\mathcal{A}}} K\-, \end{array} $$
(40)
$$\begin{array}{@{}rcl@{}} T^{2} &=&\frac{16H^{3}m^{2}K^{2}}{3g\mathcal{A}\left[mH+\mathcal{A}\left( 2-m-\frac{3E}{K}\right)\right]}\-, \end{array} $$
(41)

where K is the complete elliptic integral of second kind, and \(\mathcal {A}\) is the cnoidal wave amplitude. Using Eqs. 12 and 4041, the elliptic parameter m is related to the wavelength, λ, and cnoidal wave period, T, through the following relation:

$$c-\frac{\lambda}{T}=0\-. $$

Here, c is the cnoidal wave phase speed. The root of this equation, which varies from 0 to 1, is called cnoidal elliptic parameter. Here, we use the bisection technique to find modulus m.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Bakhoday-Paskyabi, M. Particle motions beneath irrotational water waves. Ocean Dynamics 65, 1063–1078 (2015). https://doi.org/10.1007/s10236-015-0856-4

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10236-015-0856-4

Keywords

Navigation