Next Article in Journal
Large-Scale, Multidisciplinary Laboratory Teaching of Fluid Mechanics
Next Article in Special Issue
Wave Patterns of Gravity–Capillary Waves from Moving Localized Sources
Previous Article in Journal
Simplified Modelling of Inclined Turbulent Dense Jets
Previous Article in Special Issue
Faraday Waves in a Square Cell Network: The Effects of Varying the Cell Size
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Interaction Features of Internal Wave Breathers in a Stratified Ocean

by
Ekaterina Didenkulova
1,* and
Efim Pelinovsky
1,2,3
1
Faculty of Informatics, Mathematics, and Computer Science, National Research University Higher School of Economics, 25/12, Bolshaya Pecherskaya str., 603155 Nizhny Novgorod, Russia
2
Laboratory of Modeling of Natural and Anthropogenic Disasters, Nizhny Novgorod State Technical University n.a. R.E. Alekseev, 24, Minina str., 603155 Nizhny Novgorod, Russia
3
Department of Nonlinear Geophysical Processes, Institute of Applied Physics of the Russian Academy of Science, 46, Ul’yanova str., 603155 Nizhny Novgorod, Russia
*
Author to whom correspondence should be addressed.
Fluids 2020, 5(4), 205; https://doi.org/10.3390/fluids5040205
Submission received: 17 October 2020 / Revised: 6 November 2020 / Accepted: 8 November 2020 / Published: 10 November 2020
(This article belongs to the Special Issue Recent Advances in Free Surface Hydrodynamics)

Abstract

:
Oscillating wave packets (breathers) are a significant part of the dynamics of internal gravity waves in a stratified ocean. The formation of these waves can be provoked, in particular, by the decay of long internal tidal waves. Breather interactions can significantly change the dynamics of the wave fields. In the present study, a series of numerical experiments on the interaction of breathers in the frameworks of the etalon equation of internal waves—the modified Korteweg–de Vries equation (mKdV)—were conducted. Wave field extrema, spectra, and statistical moments up to the fourth order were calculated.

1. Introduction

Localized wave formations (solitons or breathers) play a key role in the dynamics of different nonlinear integrable systems (the nonlinear Schrödinger equation, the Korteweg–de Vries (KdV) equation, etc.), which describe physical phenomena in the environment, including surface and internal water waves and wave movements in optics and plasma. The classical problem of soliton theory about their interaction with each other began to develop back in the last century [1], but is still relevant [2,3,4,5], also for non-integrable equations [6,7]. Multiple soliton interactions lead to so-called soliton turbulence or soliton gas [8,9,10,11,12,13,14], and such interactions are blamed for the formation of abnormally large (freak or rogue) waves [15,16,17,18]. The impact of pairwise soliton interactions and multiple soliton interactions on soliton gas dynamics and statistics has been considered in [19,20,21,22,23,24].
Like solitons, breathers also have particle-like behavior in some sense, but they have been studied much less. The existence of breathers has been proved in many experiments, and some recent studies have been devoted to the dynamics of breathers on the surface of deep water [25,26,27] and inside the ocean [28,29,30,31]. The most common models to describe the dynamics of internal waves in a stratified ocean are the extended KdV (the Gardner equation) when both cubic and quadratic nonlinearities are essential, and the modified Korteweg-de Vries equation, when quadratic nonlinearity can be neglected. The existence of solitons and breathers within these models is well predicted by weakly nonlinear theory [32,33]. These waves are often observed in the ocean [34,35]. Such waves complicate human economic activity in the shelf area, affect the propagation of acoustic signals, the motion of submersibles and mixing stratified water, and the transfer of admixtures and pollution; they lead to soil erosion under oil and gas platforms and change the plankton distribution. Interactions of energy-carrying waves play a special role in these processes. Some features of the interaction between a soliton and a breather have been considered in [36] using the Gardner equation and in [37] using the mKdV equation. Breather interactions are a complex physical process, the results of which depend on several wave parameters: amplitudes, positions of waves (initial phases) and the number of waves in the oscillating packets. Such a variation of parameters leads to a wide variety of resulting impulses at the time of interaction. Pairwise breather interactions using the Sine-Gordon equations have been considered in [38,39,40]. The interaction of two breathers within the Gardner equation has been considered in [41]. In the present work, the study of interaction features between two breathers has been carried out using the mKdV equation.

2. The mKdV Equation and Its Breather Solution

The mKdV equation can be derived using the asymptotic technique for waves in a liquid medium with a specific stratification (both continuous and multilayer), and in a moving frame of reference, it takes the form:
u t   + α 1 u 2 u x + β 3 u x 3   =   0 ,
where α 1 is the coefficient of the cubic nonlinear term, and β is the coefficient of dispersion. In the Boussinesq approximation, the coefficients are determined in terms of given vertical distributions of the field [42]:
β   =   1 2 0 H ( c U ) 2 Φ 2 d z 0 H ( c U ) ( d Φ / dz ) 2 d z ,   α 1   =   3 2 0 H ( c U ) 2 [ 2 ( d Φ dz ) 4 3 ( d T d z ) ( d Φ dz ) 2 ] d z 0 H ( c U ) ( d Φ dz ) 2 d z
Here, c is the phase velocity of long waves, U is a shear flow, and the function T is a nonlinear correction to the modal function Φ, determined by the equation:
d d z [ ( c U ) 2 d T d z ] + N 2 T   =   3 2 d d z [ ( c U ) 2 ( d Φ d z ) 2 ]
with zero boundary conditions T(0) = T(H) = 0 and normalization conditions T(zmax) = 0, where zmax is the coordinate of the maximum point of the modal function Φ(zmax) = 1, N(z) is the Brunt–Vaisala frequency.
In the particular case of a three-layer model, the coefficients in Equation (1) have been found in [42].
After scaling, the coefficients are: α1 = 6 and β = 1. Equation (1) with such coefficients is canonical and completely integrable by means of the inverse scattering method.
The single-breather solution of the mKdV equation is:
u ( x , t )   =   4 q sec h ( R ) [ cos ( f )     q p sin ( f ) tanh ( R ) 1   +   ( q p ) 2 sin 2 ( f ) sec h 2 ( R ) ] ,
where
R   =   2 q x + 8 q ( 3 p 2 q 2 ) t + R 0 ,  
f   =   2 p x + 8 p ( p 2 3 q 2 ) t + f 0 ,
p ,   q ,   R 0 ,   f 0   are free parameters: p affects the number of waves in a packet, and q determines the breather amplitude; constants   R 0 and   f 0 determine the initial positions of the envelope and carrier, and their meaning is obvious in the case of a linear wave packet. If q >> p, a breather has few cycles and resembles the superposition of the two mKdV-solitons. A breather containing many oscillations (p >> q) can be interpreted as an envelope soliton of the nonlinear Schrödinger equation [43].
Breathers are sometimes called pulsating solitons because they propagate as isolated disturbances without losses and have an additional internal “oscillatory” degree of freedom. In contrast to solitons, all breather solutions (4) have zero mass. According to the relation between p and q, a breather may propagate in any direction.

3. Pairwise Breather Interaction

The features of two-breather interactions are studied as an elementary act of soliton-breather turbulence for a further understanding of their impact on multi-wave dynamics. A series of 150 numerical experiments with different breather phases was conducted. The initial conditions corresponded to the sum of two breathers (4) with parameters f01, R01, p1, q1 and f02, R02, p2, q2, respectively. In each experiment, breathers were initially separated by the level 10−5, and the calculations stopped when the distance between waves after their interaction became the same as at t = 0. Equation (1) is solved numerically by the pseudo-spectral method with periodic boundary conditions. This method replaces the initial partial differential equations with ordinary differential equations for the coefficients of the expansion of the desired functions for a certain basis. At each time, the coefficients allow one to retrieve the desired solution with the help of a fast Fourier transform. This method is described in more detail in [44]. Numerical simulations are controlled by retaining the first and second moments with precisions of 10−14 and 10−7, respectively.
In each experiment there are always two breathers with fixed parameters f01 = 0; p01 = 0.05; q01 = 0.25 and f02 = 0; R02 = 0; p02 = 2; q02 = 0.25. The first breather has an N-shape (a pair of coupled solitons), and the second one contains many oscillations. The only parameter which is changed from one experiment to another is the position of the first breather, R01. In the experiments, it changes from 30 to 61.5, and this range covers all possible shapes of the first breather (during one period). Several examples of initial conditions are shown in Figure 1. Red curves correspond to the boundary positions of the first breather. The initial form and phase of the second breather remain the same.
The chosen parameters p01, q01, p02, q02 correspond to the counter propagation of the breathers. The interaction from a single experiment is demonstrated in Figure 2. Breathers retain their identity after the interaction; there is no energy loss after the interaction. A spatiotemporal diagram of the interaction shows the difference in the velocities of the breathers, which remain constant before and after the interaction (Figure 3). The first breather is much slower than the second one. There are phase shifts, which are observed in Figure 3.

4. Wave Field Extrema

Breathers “pulsate” during their propagation, and the internal oscillations can greatly change their amplitude. These changes intensify as the number of oscillations in the wave packet decreases. Both breathers participating in the interaction reach the maximum positive and negative amplitude equal to 1 during the propagation. For optimal phases of interacting breathers, the amplification of the wave field is equal to the sum of the maximum amplitudes of the breathers [23], i.e., 2 in this case. However, such situations are extremely rare in real physical systems, and non-optimal interactions without maximum wave amplification often occur. In Figure 4, the maximum maximorums and minimum minimorums are plotted for each numerical experiment (for each R01), up to 64.5, to demonstrate the periodicity of the functions. Because both breathers oscillate during the propagation, changing their shape, the number of unique forms of the resulting impulses at the time of the interaction is extremely large. The series of experiments covering all possible initial forms of the first breather was considered. Although this does not give all the possible interactions, these experiments show a typical picture of the interaction between breathers with given parameters, which may undergo small quantitative but not qualitative changes. Figure 4 demonstrates the change in the maximum field values from 1.2 to close to 2, and the minimum values from −2 to −1.2; the maxima and minima on both graphs relate to each other.
The pulse shapes at the instant of interaction with the largest positive and negative amplification of the wave field are shown in Figure 5. Both figures correspond to near-optimal focusing, and their shapes are almost symmetrical about the vertical axis. In the case of optimal focusing, the resulting pulse is asymmetric about the horizontal axis: it is significantly extended towards the amplification of the wave field (in contrast to non-optimal focusing, see the middle plot in Figure 2).

5. Moments and Spectra of the Wave Fields

A statistical approach, when the probability distribution functions of the wave parameters, statistical moments, etc., are calculated to describe the stochastic wave dynamics, is often used in the theory of turbulence. This section examines the behavior of the moments of the wave fields consisting of the two breathers discussed in the previous sections. The moments of the wave fields corresponding to the mean field, variance, skewness and kurtosis, see, for example, [45], are calculated by:
M n ( t )   =   + u n ( x , t ) d x ,   n   =   1 , 2 , 3 , 4 .
The first two of them correspond to invariants (conservation laws of the mKdV equation); therefore, they are preserved in the process of breather interaction. The most interesting is the third moment, which is responsible for the asymmetry of the wave field and the fourth, which is responsible for the amplification of the wave field. Figure 6 shows the graphs of changes in M3 and M4 for two experiments corresponding to the largest negative and positive field amplification during the interaction of breathers: R01 = 31.3 and R01 = 47. First, when breathers are separated, the left (the slowest) breather gives the largest impact on M3 and M4 (because the moment changes of the right breather are close to constant; see [43]). During the considered time equals to 3, the form of the left breather changes very little, so it gives small changes in the “background” in Figure 6. At the same time, we are interested in the interaction process, which happens very quickly because of the high speed of the second breather. For longer times, this function will be periodic. Second, we see a pronounced stage of the wave interaction on these graphs. The third moment increases significantly at R01 = 47, signaling a positive amplification of the wave field, and significantly decreases at R01 = 31.3, signaling a negative amplification of the wave field. The fourth moment behaves in approximately the same way during the positive and negative wave amplification.
The Fourier spectra of the wave field, corresponded to the positive “optimal” interaction (R01 = 47), before interaction (t = 0), at the moment of the strongest breather interaction (t = t*) and after the interaction (t = tend) shown in Figure 7a. At t = t*, the spectrum is smoother, and it decreases slower; there are two pronounced depressions. Before and after the interaction, the spectrum is rougher. Different relative breather phases influence the shape of two depressions at the moment of breather interaction (Figure 7b).

6. Conclusions

A series of 150 numerical experiments on the interaction of two breathers using the mKdV equation was carried out. In each experiment, one breather was N-shaped, and the other had several oscillations. In each experiment, the phase of the first breather was changed. The considered series of experiments covers all possible initial forms of the first breather. Although this does not give all the possible variants of interaction, these experiments show a typical picture of the interaction of breathers with given parameters, which may undergo small quantitative changes, but not qualitative ones. As a result of the calculations, cases with “optimal” and “non-optimal” interaction of breathers were analyzed. It is shown that, as a result of the interaction, the amplitude of the resulting pulse can reach double the amplification of the initial amplitudes of the breathers. These results are important for further study of multi-soliton-breather dynamics.

Author Contributions

Conceptualization, E.P.; investigation, E.D.; writing—original draft, E.D.; writing—review and editing, E.P. All authors have read and agreed to the published version of the manuscript.

Funding

The work was funded by RFBR grant No. 19-35-60022. Studies were partially supported by the Laboratory of Dynamical Systems and Applications NRU HSE, of the Ministry of science and higher education of the RF grant ag. No. 075-15-2019-1931, and the Foundation for the Advancement of Theoretical Physics and Mathematics “BASIS” No. 20-1-3-3-1.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Zakharov, V.E.; Manakov, S.V.; Novikov, S.P.; Pitaevskii, L.P. Theory of Solitons. In The Method of the Inverse Problem; Nauka: Moscow, Russia, 1980; p. 339. [Google Scholar]
  2. Slyunyaev, A.V. Dynamics of localized waves with large amplitude in a weakly dispersive medium with a quadratic and positive cubic nonlinearity. JETP 2001, 92, 529–534. [Google Scholar] [CrossRef]
  3. Slyunyaev, A.V.; Pelinovskii, E.N. Dynamics of large-amplitude solitons. JETP 1999, 89, 173–181. [Google Scholar] [CrossRef]
  4. Anco, S.C.; Ngatat, N.T.; Willoughby, M. Interaction properties of complex modified kortewegde Vries (mKdV) solitons. Physica D 2011, 240, 1378–1394. [Google Scholar] [CrossRef] [Green Version]
  5. Ali, R.; Chatterjee, P. Three-Soliton Interaction and Soliton Turbulence in Superthermal Dusty Plasmas. Zeitschrift für Naturforschung A 2019, 74, 757–766. [Google Scholar] [CrossRef]
  6. Kachulin, D.; Gelash, A. On the phase dependence of the soliton collisions in the Dyachenko–Zakharov envelope equation. Nonlin. Processes Geophys. 2018, 25, 553–563. [Google Scholar] [CrossRef] [Green Version]
  7. Kurkina, O.E.; Kurkin, A.A.; Ruvinskaya, E.A.; Pelinovsky, E.N.; Soomere, T. Dynamics of solitons in a nonintegrable version of the modified Korteweg-de Vries equation. JETP Lett. 2012, 95, 91–95. [Google Scholar] [CrossRef]
  8. Agafontsev, D.S.; Zakharov, V.E. Integrable turbulence generated from modulational instability of cnoidal waves. Nonlinearity 2016, 29, 3551–3578. [Google Scholar] [CrossRef] [Green Version]
  9. Kachulin, D.; Dyachenko, A.; Dremov, S. Multiple Soliton Interactions on the Surface of Deep Water. Fluids 2020, 5, 65. [Google Scholar] [CrossRef]
  10. Kachulin, D.; Dyachenko, A.; Zakharov, V. Soliton Turbulence in Approximate and Exact Models for Deep Water Waves. Fluids 2020, 5, 67. [Google Scholar] [CrossRef]
  11. Pelinovsky, E.; Shurgalina, E. KDV soliton gas: Interactions and turbulence. In Challenges in Complexity: Dynamics, Patterns, Cognition; Aronson, I., Pikovsky, A., Rulkov, N., Tsimring, L., Eds.; Series: Nonlinear Systems and Complexity; Springer: Berlin, Germany, 2017; Volume 20, pp. 295–306. [Google Scholar]
  12. Didenkulova, E.G.; Pelinovsky, E.N. The Role of a Thick Soliton in the Dynamics of the Soliton Gas within the Framework of the Gardner Equation. Radiophys. Quantum Electron. 2019, 61, 623–632. [Google Scholar] [CrossRef]
  13. Dutykh, D.; Pelinovsky, E. Numerical simulation of a solitonic gas in KdV and KdV–BBM equations. Phys. Lett. A 2014, 378, 3102–3110. [Google Scholar] [CrossRef] [Green Version]
  14. Shurgalina, E.G.; Pelinovsky, E.N. Nonlinear dynamics of a soliton gas: Modified Korteweg-de Vries equation framework. Phys. Lett. A 2016, 380, 2049–2053. [Google Scholar] [CrossRef]
  15. Gelash, A.A.; Agafontsev, D.S. Strongly interacting soliton gas and formation of rogue waves. Phys. Rev. E. 2018, 98, 1–11. [Google Scholar] [CrossRef] [Green Version]
  16. Pelinovsky, E.N.; Shurgalina, E.G. Formation of freak waves in a soliton gas described by the modified Korteweg–de Vries equation. Dokl. Phys. 2016, 61, 423–426. [Google Scholar] [CrossRef]
  17. Didenkulova (Shurgalina), E.G. Numerical modeling of soliton turbulence within the focusing Gardner equation: Rogue wave emergence. Physica D 2019, 399, 35–41. [Google Scholar] [CrossRef]
  18. Soto-Crespo, J.M.; Devine, N.; Akhmediev, N. Integrable Turbulence and Rogue Waves: Breathers or Solitons? Phys. Rev. Lett. 2016, 116, 103901. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  19. Pelinovsky, E.N.; Shurgalina, E.G.; Sergeeva, A.V.; Talipova, T.G.; El, G.A.; Grimshaw, R.H.J. Two-soliton interaction as an elementary act of soliton turbulence in integrable systems. Phys. Lett. A 2013, 377, 272–275. [Google Scholar] [CrossRef] [Green Version]
  20. Pelinovsky, E.N.; Shurgalina, E.G. Two-soliton interaction in the frameworks of modified Korteweg – de Vries equation. Radiophys. Quantum Electron. 2015, 57, 737–744. [Google Scholar] [CrossRef] [Green Version]
  21. Shurgalina, E.G. The features of the paired soliton interactions within the framework of the Gardner equation. Radiophys. Quantum Electron. 2018, 60, 703–708. [Google Scholar] [CrossRef]
  22. Shurgalina, E.G. The mechanism of the formation of freak waves in the result of interaction of internal waves in stratified basin. Fluid Dyn. 2018, 53, 59–64. [Google Scholar] [CrossRef]
  23. Slunyaev, A.V.; Pelinovsky, E.N. Role of multiple soliton interactions in the generation of rogue waves: The modified korteweg-de vries framework. Phys. Rev. Lett. 2016, 117, 214501. [Google Scholar] [CrossRef] [Green Version]
  24. Slunyaev, A. On the optimal focusing of solitons and breathers in long-wave models. Stud. Appl. Math. 2019, 142, 385–413. [Google Scholar] [CrossRef]
  25. Fedele, F. On the persistence of breathers at deep water. JETP Lett. 2014, 98, 523–527. [Google Scholar] [CrossRef]
  26. Kachulin, D.; Dyachenko, A.; Gelash, A. Interactions of coherent structures on the surface of deep water. Fluids 2019, 4, 83. [Google Scholar] [CrossRef] [Green Version]
  27. Wang, J.; Ma, Q.W.; Yan, S.; Chabchoub, A. Breather Rogue Waves in Random Seas. Phys. Rev. Appl. 2018, 9, 014016. [Google Scholar] [CrossRef]
  28. Lamb, K.G.; Polukhina, O.; Talipova, T.; Pelinovsky, E.; Xiao, W.; Kurkin, A. Breather generation in fully nonlinear models of a stratified fluid. Phys. Rev. E 2007, 75, 046306. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  29. Talipova, T.; Kurkina, O.; Kurkin, A.; Didenkulova, E.; Pelinovsky, E. Internal Wave Breathers in the Slightly Stratified Fluid. Microgravity Sci. Technol. 2020, 32, 69–77. [Google Scholar] [CrossRef]
  30. Rouvinskaya, E.; Talipova, T.; Kurkina, O.; Soomere, T.; Tyugin, D. Transformation of internal breathers in the idealised shelf sea conditions. Cont. Shelf Res. 2015, 110, 60–71. [Google Scholar] [CrossRef]
  31. Nakayama, K.; Lamb, K. Breathers in a three-layer fluid. J. Fluid Mech. 2020, 903, A40. [Google Scholar] [CrossRef]
  32. Grimshaw, R.; Pelinovsky, E.; Talipova, T. The modified Korteweg–de Vries equation in the theory of large-amplitude internal waves. Nonlinear Process. Geophys. 1997, 4, 237–250. [Google Scholar] [CrossRef]
  33. Clarke, S.; Grimshaw, R.; Miller, P.; Pelinovsky, E.; Talipova, T. On the generation of solitons and breathers in the modified Korteweg—de Vries equation. Chaos 2000, 10, 383–392. [Google Scholar] [CrossRef] [PubMed]
  34. Lee, J.H.; Lozovatsky, I.; Jang, S.-T.; Jang, C.J.; Hong, C.S.; Fernando, H.J.S. Episodes of nonlinear internal waves in the northern East China Sea. Geophys. Res. Lett. 2006, 33, L18601. [Google Scholar] [CrossRef]
  35. Shroyer, E.L.; Moum, J.N.; Nash, J.D. Mode 2 waves on the continental shelf: Ephemeral components of the nonlinear internal wavefield. JGR 2010, 115, C07001. [Google Scholar] [CrossRef] [Green Version]
  36. Chow, K.W.; Grimshaw, R.H.J.; Ding, E. Interactions of breathers and solitons in the extended Korteweg–de Vries equation. Wave Motion 2005, 43, 158–166. [Google Scholar] [CrossRef] [Green Version]
  37. Didenkulova, E.; Pelinovsky, E. Soliton–Breather Interaction: The Modified Korteweg–de Vries Equation Framework. Symmetry 2020, 12, 1445. [Google Scholar] [CrossRef]
  38. Kevrekidis, P.G.; Saxena, A.; Bishop, A.R. Interaction between sine-Gordon breathers. Phys. Rev. E 2001, 64, 026613. [Google Scholar] [CrossRef] [PubMed]
  39. Kevrekidis, P.G.; Khare, A.; Saxena, A. Solitary wave interactions in dispersive equations using Manton’s approach. Phys. Rev. E 2004, 70, 057603. [Google Scholar] [CrossRef] [Green Version]
  40. Nishida, M.; Nishida, M.; Furukawa, Y.; Fujii, T.; Hatakenaka, N. Breather-breather interactions in sine-Gordon systems using collective coordinate approach. Phys. Rev. E 2009, 80, 036603. [Google Scholar] [CrossRef]
  41. Rouvinskaya, E.A.; Kurkina, O.E.; Kurkin, A.A.; Korol, A.A. Dynamics of breathers in the framework of the Gardner equation. In Proceedings of the International Scientific and Technical Conference “Information Systems and Technologies”, Nizhny Novgorod, Russia, 21 April 2017. [Google Scholar]
  42. Talipova, T.; Pelinovsky, E.; Lamb, K.; Grimshaw, R.; Holloway, P. Cubic nonlinearity effects in the propagation of intense internal waves. Doklady Earth Sci. 1999, 365, 241–244. [Google Scholar]
  43. Didenkulova, E.; Pelinovsky, E. Breather’s Properties within the Framework of the Modified Korteweg–de Vries Equation. Symmetry 2020, 12, 638. [Google Scholar] [CrossRef]
  44. Fronberg, B. A Practical Guide to Pseudospectral Methods; Cambridge University Press: Cambridge, UK, 1998; p. 231. [Google Scholar]
  45. Monin, A.S.; Yaglom, A.M. Statistical Fluid Mechanics; MIT Press: Cambridge, MA, USA, 1971; p. 769. [Google Scholar]
Figure 1. Examples of the initial conditions.
Figure 1. Examples of the initial conditions.
Fluids 05 00205 g001
Figure 2. Pairwise breather interaction with the following parameters: f01 = 0; R01 = 37; p01 = 0.05; q01 = 0.25; and f02 = 0; R02 = 0; p02 = 2; q02 = 0.25.
Figure 2. Pairwise breather interaction with the following parameters: f01 = 0; R01 = 37; p01 = 0.05; q01 = 0.25; and f02 = 0; R02 = 0; p02 = 2; q02 = 0.25.
Fluids 05 00205 g002
Figure 3. Spatiotemporal diagram of the breather-breather interaction.
Figure 3. Spatiotemporal diagram of the breather-breather interaction.
Fluids 05 00205 g003
Figure 4. Maximum and minimum values of umax and umin in each experiment.
Figure 4. Maximum and minimum values of umax and umin in each experiment.
Fluids 05 00205 g004
Figure 5. Shapes of resulting pulses at the moment of optimal focusing: (a) f01 = 0; R01 = 47; p01 = 0.05; q01 = 0.25; and f02 = 0; R02 = 0; p02 = 2; q02 = 0.25, (b) f01 = 0; R01 = 31.3; p01 = 0.05; q01 = 0.25; and f02 = 0; R02 = 0; p02 = 2; q02 = 0.25.
Figure 5. Shapes of resulting pulses at the moment of optimal focusing: (a) f01 = 0; R01 = 47; p01 = 0.05; q01 = 0.25; and f02 = 0; R02 = 0; p02 = 2; q02 = 0.25, (b) f01 = 0; R01 = 31.3; p01 = 0.05; q01 = 0.25; and f02 = 0; R02 = 0; p02 = 2; q02 = 0.25.
Fluids 05 00205 g005
Figure 6. Evolution of M3 and M4 in time for two experiments with maximum (dotted curve) and minimum (solid curve) amplification.
Figure 6. Evolution of M3 and M4 in time for two experiments with maximum (dotted curve) and minimum (solid curve) amplification.
Fluids 05 00205 g006
Figure 7. (a) Fourier spectra of the wave field corresponded to the positive “optimal” interaction (R01 = 47) at t = 0, t = t* and t = tend, (b) Fourier spectra of the wave fields corresponded to positive “optimal” interaction (R01 = 47) and negative “optimal” interaction (R01 = 31.3) at the interaction moment.
Figure 7. (a) Fourier spectra of the wave field corresponded to the positive “optimal” interaction (R01 = 47) at t = 0, t = t* and t = tend, (b) Fourier spectra of the wave fields corresponded to positive “optimal” interaction (R01 = 47) and negative “optimal” interaction (R01 = 31.3) at the interaction moment.
Fluids 05 00205 g007
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Didenkulova, E.; Pelinovsky, E. Interaction Features of Internal Wave Breathers in a Stratified Ocean. Fluids 2020, 5, 205. https://doi.org/10.3390/fluids5040205

AMA Style

Didenkulova E, Pelinovsky E. Interaction Features of Internal Wave Breathers in a Stratified Ocean. Fluids. 2020; 5(4):205. https://doi.org/10.3390/fluids5040205

Chicago/Turabian Style

Didenkulova, Ekaterina, and Efim Pelinovsky. 2020. "Interaction Features of Internal Wave Breathers in a Stratified Ocean" Fluids 5, no. 4: 205. https://doi.org/10.3390/fluids5040205

Article Metrics

Back to TopTop