Three-wave equation

From HandWiki

In nonlinear systems, the three-wave equations, sometimes called the three-wave resonant interaction equations or triad resonances, describe small-amplitude waves in a variety of non-linear media, including electrical circuits and non-linear optics. They are a set of completely integrable nonlinear partial differential equations. Because they provide the simplest, most direct example of a resonant interaction, have broad applicability in the sciences, and are completely integrable, they have been intensively studied since the 1970s.[1]

Informal introduction

The three-wave equation arises by consideration of some of the simplest imaginable non-linear systems. Linear differential systems have the generic form

[math]\displaystyle{ D\psi=\lambda\psi }[/math]

for some differential operator D. The simplest non-linear extension of this is to write

[math]\displaystyle{ D\psi-\lambda\psi=\varepsilon\psi^2. }[/math]

How can one solve this? Several approaches are available. In a few exceptional cases, there might be known exact solutions to equations of this form. In general, these are found in some ad hoc fashion after applying some ansatz. A second approach is to assume that [math]\displaystyle{ \varepsilon\ll 1 }[/math] and use perturbation theory to find "corrections" to the linearized theory. A third approach is to apply techniques from scattering matrix (S-matrix) theory.

In the S-matrix approach, one considers particles or plane waves coming in from infinity, interacting, and then moving out to infinity. Counting from zero, the zero-particle case corresponds to the vacuum, consisting entirely of the background. The one-particle case is a wave that comes in from the distant past and then disappears into thin air; this can happen when the background is absorbing, deadening or dissipative. Alternately, a wave appears out of thin air and moves away. This occurs when the background is unstable and generates waves: one says that the system "radiates". The two-particle case consists of a particle coming in, and then going out. This is appropriate when the background is non-uniform: for example, an acoustic plane wave comes in, scatters from an enemy submarine, and then moves out to infinity; by careful analysis of the outgoing wave, characteristics of the spatial inhomogeneity can be deduced. There are two more possibilities: pair creation and pair annihilation. In this case, a pair of waves is created "out of thin air" (by interacting with some background), or disappear into thin air.

Next on this count is the three-particle interaction. It is unique, in that it does not require any interacting background or vacuum, nor is it "boring" in the sense of a non-interacting plane-wave in a homogeneous background. Writing [math]\displaystyle{ \psi_1, \psi_2, \psi_3 }[/math] for these three waves moving from/to infinity, this simplest quadratic interaction takes the form of

[math]\displaystyle{ (D-\lambda)\psi_1=\varepsilon\psi_2\psi_3 }[/math]

and cyclic permutations thereof. This generic form can be called the three-wave equation; a specific form is presented below. A key point is that all quadratic resonant interactions can be written in this form (given appropriate assumptions). For time-varying systems where [math]\displaystyle{ \lambda }[/math] can be interpreted as energy, one may write

[math]\displaystyle{ (D-i\partial/\partial t)\psi_1=\varepsilon\psi_2\psi_3 }[/math]

for a time-dependent version.

Review

Formally, the three-wave equation is

[math]\displaystyle{ \frac{\partial B_j}{\partial t} + v_j \cdot \nabla B_j=\eta_j B^*_\ell B^*_m }[/math]

where [math]\displaystyle{ j,\ell,m=1,2,3 }[/math] cyclic, [math]\displaystyle{ v_j }[/math] is the group velocity for the wave having [math]\displaystyle{ \vec k_j, \omega_j }[/math] as the wave-vector and angular frequency, and [math]\displaystyle{ \nabla }[/math] the gradient, taken in flat Euclidean space in n dimensions. The [math]\displaystyle{ \eta_j }[/math] are the interaction coefficients; by rescaling the wave, they can be taken [math]\displaystyle{ \eta_j=\pm 1 }[/math]. By cyclic permutation, there are four classes of solutions. Writing [math]\displaystyle{ \eta=\eta_1\eta_2\eta_3 }[/math] one has [math]\displaystyle{ \eta=\pm 1 }[/math]. The [math]\displaystyle{ \eta=-1 }[/math] are all equivalent under permutation. In 1+1 dimensions, there are three distinct [math]\displaystyle{ \eta=+1 }[/math] solutions: the [math]\displaystyle{ +++ }[/math] solutions, termed explosive; the [math]\displaystyle{ --+ }[/math] cases, termed stimulated backscatter, and the [math]\displaystyle{ -+- }[/math] case, termed soliton exchange. These correspond to very distinct physical processes.[2][3] One interesting solution is termed the simulton, it consists of three comoving solitons, moving at a velocity v that differs from any of the three group velocities [math]\displaystyle{ v_1, v_2, v_3 }[/math]. This solution has a possible relationship to the "three sisters" observed in rogue waves, even though deep water does not have a three-wave resonant interaction.

The lecture notes by Harvey Segur provide an introduction.[4]

The equations have a Lax pair, and are thus completely integrable.[1][5] The Lax pair is a 3x3 matrix pair, to which the inverse scattering method can be applied, using techniques by Fokas.[6][7] The class of spatially uniform solutions are known, these are given by Weierstrass elliptic ℘-function.[8] The resonant interaction relations are in this case called the Manley–Rowe relations; the invariants that they describe are easily related to the modular invariants [math]\displaystyle{ g_2 }[/math] and [math]\displaystyle{ g_3. }[/math][9] That these appear is perhaps not entirely surprising, as there is a simple intuitive argument. Subtracting one wave-vector from the other two, one is left with two vectors that generate a period lattice. All possible relative positions of two vectors are given by Klein's j-invariant, thus one should expect solutions to be characterized by this.

A variety of exact solutions for various boundary conditions are known.[10] A "nearly general solution" to the full non-linear PDE for the three-wave equation has recently been given. It is expressed in terms of five functions that can be freely chosen, and a Laurent series for the sixth parameter.[8][9]

Applications

Some selected applications of the three-wave equations include:

  • In non-linear optics, tunable lasers covering a broad frequency spectrum can be created by parametric three-wave mixing in quadratic ([math]\displaystyle{ \chi^{(2)} }[/math]) nonlinear crystals.[citation needed]
  • Surface acoustic waves and in electronic parametric amplifiers.
  • Deep water waves do not in themselves have a three-wave interaction; however, this is evaded in multiple scenarios:
    • Deep-water capillary waves are described by the three-wave equation.[4]
    • Acoustic waves couple to deep-water waves in a three-wave interaction,[11]
    • Vorticity waves couple in a triad.
    • A uniform current (necessarily spatially inhomogenous by depth) has triad interactions.
These cases are all naturally described by the three-wave equation.

References

  1. 1.0 1.1 Zakharov, V. E.; Manakov, S. V. (1975). "On the theory of resonant interaction of wave packets in nonlinear media". Soviet Physics JETP 42 (5): 842–850. http://jetp.ac.ru/cgi-bin/dn/e_042_05_0842.pdf. 
  2. Degasperis, A.; Conforti, M.; Baronio, F.; Wabnitz, S.; Lombardo, S. (2011). "The Three-Wave Resonant Interaction Equations: Spectral and Numerical Methods". Letters in Mathematical Physics 96 (1–3): 367–403. doi:10.1007/s11005-010-0430-4. Bibcode2011LMaPh..96..367D. https://hal.archives-ouvertes.fr/hal-02395151/file/3WNUMERFINALE.pdf. 
  3. Kaup, D. J.; Reiman, A.; Bers, A. (1979). "Space-time evolution of nonlinear three-wave interactions. I. Interaction in a homogeneous medium". Reviews of Modern Physics 51 (2): 275–309. doi:10.1103/RevModPhys.51.275. Bibcode1979RvMP...51..275K. 
  4. 4.0 4.1 Segur, H.; Grisouard, N. (2009). "Lecture 13: Triad (or 3-wave) resonances". Woods Hole Oceanographic Institution. https://gfd.whoi.edu/wp-content/uploads/sites/18/2018/03/lecture13-harvey_136505.pdf. 
  5. Zakharov, V. E.; Manakov, S. V.; Novikov, S. P.; Pitaevskii, L. I. (1984). Theory of Solitons: The Inverse Scattering Method. New York: Plenum Press. Bibcode1984lcb..book.....N. 
  6. Fokas, A. S.; Ablowitz, M. J. (1984). "On the inverse scattering transform of multidimensional nonlinear equations related to first‐order systems in the plane". Journal of Mathematical Physics 25 (8): 2494–2505. doi:10.1063/1.526471. Bibcode1984JMP....25.2494F. 
  7. Lenells, J. (2012). "Initial-boundary value problems for integrable evolution equations with 3×3 Lax pairs". Physica D 241 (8): 857–875. doi:10.1016/j.physd.2012.01.010. Bibcode2012PhyD..241..857L. 
  8. 8.0 8.1 Martin, R. A. (2015). Toward a General Solution of the Three-Wave Resonant Interaction Equations (Thesis). University of Colorado.
  9. 9.0 9.1 Martin, R. A.; Segur, H. (2016). "Toward a General Solution of the Three-Wave Partial Differential Equations". Studies in Applied Mathematics 137: 70–92. doi:10.1111/sapm.12133. 
  10. Kaup, D. J. (1980). "A Method for Solving the Separable Initial-Value Problem of the Full Three-Dimensional Three-Wave Interaction". Studies in Applied Mathematics 62: 75–83. doi:10.1002/sapm198062175. 
  11. Kadri, U. (2015). "Triad Resonance in the Gravity–Acousic Family". AGU Fall Meeting Abstracts 2015: OS11A–2006. doi:10.13140/RG.2.1.4283.1441. Bibcode2015AGUFMOS11A2006K. 
  12. Kim, J.-H.; Terry, P. W. (2011). "A self-consistent three-wave coupling model with complex linear frequencies". Physics of Plasmas 18 (9): 092308. doi:10.1063/1.3640807. Bibcode2011PhPl...18i2308K. https://zenodo.org/record/569793.