Skip to main content
Log in

Some aspects of the pressure field preceding the onset of critical jets in a breaking wave

  • Published:
Journal of Engineering Mathematics Aims and scope Submit manuscript

Abstract

We consider in this article two-dimensional nonlinear free surface motions in potential theory. There are situations where the onset of critical jet along a free surface can be predicted by analyzing the temporal and spatial variation of the pressure itself and its successive derivatives. In particular, when a local maximum of pressure appears close enough to a free surface where the pressure vanishes, it is expected a great pressure gradient at the free surface and consequently a large Lagrangian acceleration. The present analysis examines this phenomenon. This analysis is facilitated after identifying a line along which the pressure gradient is parallel to one of the eigenvectors of the Hessian matrix of the pressure. In a standard overturning crest, this line connects the tip of the crest to the bottom of the tank in which the flow is simulated. The analysis of its time evolution gives much information into the onset of critical jets. Other types of critical jet can also appear when a region of positive Gaussian curvature of the pressure suddenly grows in the vicinity of the free surface.

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

Similar content being viewed by others

References

  1. Longuet-Higgins MS (1980) A Technique for Time-Dependent Free-Surface Flows. Proc R Soc Lond A 371:441–451

    Article  MathSciNet  MATH  Google Scholar 

  2. Cooker MJ, Peregrine DH (1990) A model of breaking wave impact pressures. In: Proceedings of the 22nd Conference in Coastal Engineering, Holland ASCE, pp 1473-1486

  3. Brosset L, Mravak Z, Kaminski M Collins S, Finnigan T (2009) Overview of SLOSHEL project, In: Proceedings of the ISOPE, Osaka, Japan

  4. Chan ES, Melville WK (1988) Deep water plunging-wave pressures on a vertical plane wall. Proc R Soc Lond A 417:95–131

    Article  Google Scholar 

  5. Cooker MJ (2010) The flip-through of a plane inviscid jet with a free surface. J. Eng. Math 67:137–152

    Article  MathSciNet  MATH  Google Scholar 

  6. Peregrine DH, Prentice PR (1994) Jet formation at a free surface. In: Blake, JR, Boulton-Stone, JM, Thomas, NH (eds) Bubble Dynamics and Interface Phenomena. Fluid Mechanics and Its Applications, vol 23. Springer, Dordrecht, 23:397-404

  7. Cooker MJ (2001) Unsteady pressure fields which precede the launch of free-surface liquid jets. Proc R Soc Lond A 458:473–488

    Article  MathSciNet  MATH  Google Scholar 

  8. Longuet-Higgins MS, Og̃uz HN (1995) Critical microjets in collapsing cavities. J Fluid Mech 290:183–201

    Article  MathSciNet  MATH  Google Scholar 

  9. Zeff BW, Kleber B, Fineberg J, Lathrop DP (2000) Singularity dynamics in curvature collapse and jet eruption on a fluid surface. Nature 403:401–404

    Article  Google Scholar 

  10. Longuet-Higgins MS (1993) Highly accelerated free-surface flows. J Fluid Mech 248:449–475

    Article  MathSciNet  MATH  Google Scholar 

  11. Scolan Y-M, Etienne S (2021) (2021) Properties of the pressure field in highly nonlinear free surface flows with critical jet. J Eng Math 128:4. https://doi.org/10.1007/s10665-021-10116-4

  12. Longuet-Higgins MS (2001) Vertical jets from standing waves; the bazooka effect. In: King AC, Shikhmurzaev YD (eds) IUTAM symposium on free surface flows. Kluwer Academic Publishers, Dordrecht, pp 195–203

    Chapter  Google Scholar 

  13. Karimi MR, Brosset L, Ghidaglia J-M, Kaminski ML (2016) Effect of ullage gas on sloshing, Part II: Local effects of gas-liquid density ratio. Eur J Mech B 57:82–100

    Article  Google Scholar 

  14. Scolan Y-M (2010) Some aspects of the flip-through phenomenon: A numerical study based on the desingularized technique. J Fluid Struc 26(6):918–953

    Article  Google Scholar 

  15. Scolan Y-M (2018) Critical Free Surface Flows in a Sloshing Tank. J Adv Res Ocean Eng 4(4):163–173

    Google Scholar 

  16. Longuet-Higgins MS, Dommermuth DG (2001) Vertical jets from standing waves. II. Proc R Soc Lond A 457:2137–2149

    Article  MATH  Google Scholar 

  17. Stansby P, Chechini A, Barnes T (1998) The initial stages of dam-break flow. J Fluid Mech 374:407–424

    Article  MathSciNet  MATH  Google Scholar 

  18. Scolan Y-M, Etienne S (2020) Analysis of the pressure field in highly accelerated free surface flows (in french). In: Proceedings of \(17^{th}\) Journées de l’Hydrodynamique, Cherbourg, France

  19. Scolan Y-M (2022) Analysis of the pressure field in a breaking wave. In: Proceedings of the \(37^{th}\) International Workshop on Water Waves and Floating Bodies, Roma, Italia, pp 122-125

  20. Peregrine DH (2003) Water-wave impact on walls. Annu Rev Fluid Mech 35:23–43

    Article  MathSciNet  MATH  Google Scholar 

  21. Cooker MJ (2009) The flip-through of a plane inviscid jet with a free surface. J Eng Math 67(1):137–152

    MathSciNet  MATH  Google Scholar 

  22. Hay A, Yu KR, Etienne S, Garon A, Pelletier D (2014) High-order temporal accuracy for 3D finite-element ALE flow simulations. Computers and Fluids 100(1):204–217

    Article  MathSciNet  MATH  Google Scholar 

  23. Hay A, Etienne S, Pelletier D, Garon A (2015) Time-integration for ALE simulations of Fluid-Structure Interaction problems: Stepsize and order selection based on the BDF. Computer Methods in Applied Mechanics and Engineering 295:172–195

    Article  MathSciNet  MATH  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Yves-Marie Scolan.

Additional information

Publisher's Note

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

Appendices

Appendix A: FSID

The software FSID solves the nonlinear free surface equations in potential theory. The configuration is two dimensional. A sketch of the flow configuration is shown in Fig. 11. The boundaries of the fluid domain are the walls of a tank and the free surface. For the present applications, the tank is rectangular and the fluid is hence contained in a half infinite strip. The velocity potential is decomposed as a finite sum of Rankine sources. These singularities are located outside the fluid domain so that the solution is said desingularized. By using a conformal mapping the fluid domain is described in the upper half space where the images of the walls are the horizontal real axis. By using mirror images of the singularities, the Rankine Green function easily accounts for the impermeability of the walls. Lagrangian markers are located along the free surface, they are transported over time with their local velocity. This velocity is obtained by updating the velocity potential at the free surface. The resulting time differential system is solved by using a fourth order Runge–Kutta algorithm. The distance of desingularization, the number of markers and the time step are chosen so that the relative errors on mass and energy conservation are ensured with a maximum relative errors of \(10^{-4.5}\) and \(10^{-3}\) respectively. In the present computations, the time step is decreased when the free surface deformation becomes faster and faster. Some numerical improvements could be achieved by choosing a more precise time marching scheme.

Fig. 11
figure 11

Configuration of the flow in the rectangular tank. The fluid motion is described in a coordinate system attached to the tank. The length of the tank is L. The water depth at rest is h. The arclength \(\sigma \) is measured positively from the left wall to the right wall

The conformal function is denoted \(w=g(z)\) and z is a complex coordinate of a point in the physical plane. The fluid motion is described in a coordinate system attached to the tank. It is represented by a complex velocity potential and it is calculated at time t as

$$\begin{aligned} F(z,t)= \sum _{j=1}^{N} q_j(t) G(w,W_j(t)) \end{aligned}$$
(12)

at both complex coordinates z and w, provided that \(w=g(z)\). The intensity of the source # j is \(q_j\) and \(W_j=g(Z_j)\) is the complex coordinate of the source # j in the transformed plane w. G is the complex potential of this source plus its mirror image with respect to the axis \(\mathfrak {I}(w)=0\),

$$\begin{aligned} G(w,W_j) = G_j = \log ( w - W_j ) + \log ( w - \overline{W_j} ). \end{aligned}$$
(13)

The overline denotes the conjugate of the complex variable. The following formulæ give the first derivative in time and space of F.

$$\begin{aligned}{} & {} F_{z}= \frac{1}{J} F_{w}, \quad J= \frac{\textrm{d}z}{\textrm{d}w}, \end{aligned}$$
(14)
$$\begin{aligned}{} & {} F_{t} = \sum _{j=1}^{N} \dot{q}_j G_j + q_j G_{j,t}, \quad G_{j,t} = - \left( \frac{\dot{Z}_j}{J_j[w - W_j]} + \frac{\dot{\overline{Z}}_j }{\overline{J}_j[w - \overline{W}_j]} \right) , \quad J_j= \frac{\textrm{d}z}{\textrm{d}w}(W_j). \nonumber \\ \end{aligned}$$
(15)

The dot denotes the time derivative. The successive derivatives in (1), (4), (5) and (6) are calculated according to the same principle.

Fig. 12
figure 12

Oscillation cycle of the horizontal motion of the tank. Temporal variation of the displacement (a), velocity (b) and acceleration (c). The reference signal is in black \(X=1\). Only the signals obtained with the amplification factors \(X=1.09\) and 1.19 are plotted

Appendix B: Description of the jets appear computations of [18]

As for the configuration studied in Sect. (3), the flow is generated in a rectangular tank whose motion consists of an oscillating horizontal motion. A parametric analysis is carried out in terms of an amplification factor (denoted X) that affects the amplitude of the motion, and consequently the amplitudes of the velocity and acceleration. Figure 12 illustrate the time variations of the kinematic of the tank. The reference motion (factor \(X=1\)) is generated with a given amplitude and period of the cycle. In Fig. 12, the time signals for two factors (\(X=1.09\) and \(X=1.19\)) are superimposed. For the parametric analysis, the factor X varies in the range \(X\in [1.0\) to 1.19] by steps of 0.01.

A common feature of all the numerical tests can be summarized as three successive stages: (1) beginning of the overturning of a crest, (2) this is stopped with the formation of a jet in the middle of the nascent barrel (called here pre-bazooka jet), and (3) the development of this jet is finally interrupted by the onset of a flip-through jet which runs up along the vertical wall. Figure 13 shows the free surface profiles for a given amplification factor \(X=1.16\). One instant (\(t=1.7024\,\textrm{s}\)) just before the onset of the pre-bazooka jet is plotted: it corresponds to an almost vertical-free surface profile in the barrel below the wave crest. At later instants, ranged in the interval \(t\in [1.728\,\textrm{s}:1.732\,\textrm{s}]\), the excrescence reaches a significant size. For all the other amplification factors X, the profiles have more or less the same features. It should be noted that the parabolic shape of the free surface at the wall announces that a flip through is likely to occur (see [21]).

Fig. 13
figure 13

Successive profiles of the free surface for the amplification factor \(X=1.16\). The instants of plots are ranged in the interval \(t\in [1.728\,\textrm{s}:1.732\,\textrm{s}]\) and at one instant before the onset of the bazooka jet \(t=1.7024\,\textrm{s}\) (thick line). For the last profiles, the time interval between each profile is \(2\,\textrm{ms}\). Units along the horizontal and vertical axis: x/L and y/L with \(L=1\,\textrm{m}\) the length of the rectangular tank and \(h=0.25\,\textrm{m}\) the water depth at rest

In [18], comparisons are made with a Navier–Stokes solver CADYF developed in Polytechnique Montréal (see [22, 23]). The corresponding computations yield identical time variations of both the free surface profiles and energy distribution. Significant differences appear during the very last stages of the flow when flip through occurs. That is understandable since a Navier–Stokes solver may have numerical difficulties to capture extremely great acceleration along impermeable walls.

The parametric analysis achieved in [18] concludes that a power law (see Eq. 11) can be fitted to the time variation of the acceleration of the flip-through jet. The exponent is close to 1.

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Scolan, YM. Some aspects of the pressure field preceding the onset of critical jets in a breaking wave. J Eng Math 138, 1 (2023). https://doi.org/10.1007/s10665-022-10246-3

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s10665-022-10246-3

Keywords

Navigation