This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Experimental self-generation of axisymmetric internal wave super-harmonics

S. Boury Courant Institute of Mathematical Sciences, New York University, New York, NY 10012, USA Univ Lyon, ENS de Lyon, Univ Claude Bernard, CNRS, Laboratoire de Physique, F-69342 Lyon, France Department of Mechanical Engineering, Massachusetts Institute of Technology, Cambridge, MA 02139, USA    T. Peacock Department of Mechanical Engineering, Massachusetts Institute of Technology, Cambridge, MA 02139, USA    P. Odier Univ Lyon, ENS de Lyon, Univ Claude Bernard, CNRS, Laboratoire de Physique, F-69342 Lyon, France
(December 23, 2024)
Abstract

In this paper, we present an experimental study of weakly non-linear interaction of axisymmetric internal gravity waves in a resonant cavity, supported by theoretical considerations. Contrary to plane waves in Cartesian coordinates, for which self-interacting terms are null in a linear stratifiation, the non-linear self-interaction of an internal wave mode in axisymmetric geometry is found to be efficient at producing super-harmonics, i.e. waves whose frequencies are integer multiples of the excitation frequency. Due to the range of frequencies tested in our experiments, the first harmonic frequency is below the cut-off imposed by the stratification so the lowest harmonic created can always propagate. The study shows that the super-harmonic wave field is a sum of standing waves satisfying both the dispersion relation for internal waves and the boundary conditions imposed by the cavity walls, while conserving the axisymmetry.

internal waves, axisymmetric modes, non-linear interaction

I Introduction

In a recent review paper, MacKinnon et al. mackinnon2017BAMS discussed the mechanisms and implications of the dissipation of internal wave energy in the oceans, revisiting the 2±0.6TW2\pm 0.6\mathrm{\leavevmode\nobreak\ TW} estimate of turbulent dissipation caused by tidal flow over topographies, low-frequency lee waves, and near-inertial waves produced by wind forcing waterhouse2014 ; kunze2017 ; baker2020 . In these various scenarios, through non-linear interactions, internal wave breaking can transfer energy to smaller scales at which it can be dissipated more efficiently. Evidenced by the experimental work of Joubaud et al. joubaud2012 , Triadic Resonant Instability (TRI) has been proven relevant for wave breaking and for triggering non-linear sub-harmonic resonant cascades of highly energetic waves brouzet2016a . A very similar self-interaction mechanism has been predicted and tested numerically as a means to generate sub- and super-harmonic waves, at different scales, through the interaction with either topography or stratification wunsch2015 ; sutherland2016 ; varma2017 . Super-harmonics have also been proposed as a way to transfer energy to other scales and even as a preliminary step before producing small scale waves wagner2016 .

To date, non-linear interactions have been extensively investigated in two-dimensional Cartesian geometry (for instance in joubaud2012 ; wunsch2015 ; brouzet2016a ; sutherland2016 ; varma2017 ) but three-dimensional studies remain marginal (for example, see the recent work of shmakova2018 ). Arguably relevant to geophysics, as they mimic the geometry of a localised wave source in unconfined domains, axisymmetric wave fields are still challenging to study when dealing with linear and non-linear phenomena. Classical generation mechanisms, such as vertically oscillating spheres mowbray1967a ; ermanyuk2011 , are challenging for the study of 33D non-linearities as they usually generate quite spatially localized 3D wave fields. Instead, Maurer et al.’s axisymmetric wave generator maurer2017 , built upon the technology of Gostiaux et al.’s planar generator gostiaux2006 , has proven capable of exciting pure axisymmetric wave fields shaped as Bessel functions boury2018 , which can enable more practical studies of non-linear effects.

In non-linear stratifications, ubiquituous in geophysical flows, wave-wave interaction has been demonstrated as being capable of producing higher frequency harmonics, either through the forced interaction of two different waves husseini2019 , or through the self-interaction of the wave itself sutherland2016 ; baker2020 (note that, in what follows, we will use the term self-interaction for the interaction of a single monochromatic wave with itself). Interestingly, for non-rotating flows, 22D Cartesian self-interacting terms are null in linear stratifications and the non-uniform buoyancy frequency is therefore a condition for the existence of super-harmonics. Extending these works to 33D wave fields, however, remains challenging, and poorly investigated experimentally. Closed domains, such as the resonant cavity described in (boury2018, ), are capable of sustaining enhanced amplitude wave fields and non-linear interactions both in linear and non-linear stratifications boury2018 ; boury2019 , and could possibly lead to TRI or super-harmonic generation. Through the use of an oscillating sphere and the study of its radiated conical wave field, Ermanyuk et al. ermanyuk2011 have shown that the first harmonic created by non-linear interaction is likely to produce a non-trivial structure with a symmetry breaking, e.g. dipolar or quadripolar. In their study, however, the radiated wave field cannot be decomposed over a single Bessel function and the different components of the wave field, which can be extracted through a Fourier-Hankel transform maurer2017 , can interact together via the non-linear terms of the wave equation. In the present work, we consider a purely axisymmetric wave field shaped as a radial standing wave, called a radial mode, and discuss what is, to our knowledge, the first experimental observation of super-harmonic generation through self-interaction of axisymmetric gravity waves.

The paper is organised as follows. First, we derive the theoretical framework for linear and weakly non-linear internal wave propagation. This section also discusses the wave resonator and modal selection induced by the particular closed geometry, a cylindrical cavity, giving relevant insights on the observed phenomenon. After presenting the experimental apparatus in section 33, the results are described in section 44 and the performed analysis allows for a complete description of the created harmonics. Finally, our conclusions and discussion are presented in section 55.

II Theory

II.1 Governing Equations

In a cylindrical framework (𝐞𝐫\mathbf{e_{r}}, 𝐞θ\mathbf{e_{\theta}}, 𝐞𝐳\mathbf{e_{z}}), 𝐞𝐳\mathbf{e_{z}} being the ascendent vertical, small amplitude inertia gravity waves in an inviscid fluid with a constant background stratification satisfy the following equations under the Boussinesq approximation

ρ0(𝐯t+(𝐯)𝐯)\displaystyle\rho_{0}\left(\dfrac{\partial\mathbf{v}}{\partial t}+\left(\mathbf{v}\cdot\mathbf{\nabla}\right)\mathbf{v}\right) =\displaystyle= p(ρρ¯)g𝐞𝐳,\displaystyle-\mathbf{\nabla}p-(\rho-\bar{\rho})g\mathbf{e_{z}}, (1)
ρt+(𝐯)ρ\displaystyle\dfrac{\partial\rho}{\partial t}+\left(\mathbf{v}\cdot\mathbf{\nabla}\right)\rho =\displaystyle= 0,\displaystyle 0, (2)
𝐯\displaystyle\mathbf{\nabla}\cdot\mathbf{v} =\displaystyle= 0,\displaystyle 0, (3)

where 𝐯=(vr,vθ,vz)\mathbf{v}=(v_{r},\leavevmode\nobreak\ v_{\theta},\leavevmode\nobreak\ v_{z}), pp, ρ\rho, and ρ¯\bar{\rho} are the velocity, pressure, density, and background density fields, respectively. Setting ρ0\rho_{0} a reference density, the buoyancy field bb and frequency NN can be introduced as

b=gρ0ρandN2=gρ0ρ¯z,b=-\frac{g}{\rho_{0}}\rho^{\prime}\mathrm{\leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ and\leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ }N^{2}=-\frac{g}{\rho_{0}}\dfrac{\partial\bar{\rho}}{\partial z}, (4)

where ρ=ρρ¯\rho^{\prime}=\rho-\bar{\rho} is the perturbation to the density field.

In an axisymmetric geometry, the velocity and the buoyancy fields are θ\theta-independent and the vertical and radial velocities are therefore given by derivatives of a stream function ψ\psi as

vr=1r(rψ)zandvz=1r(rψ)r.v_{r}=-\frac{1}{r}\dfrac{\partial(r\psi)}{\partial z}\mathrm{\leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ and\leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ }v_{z}=\frac{1}{r}\dfrac{\partial(r\psi)}{\partial r}. (5)

Using this formulation, the system of equations (1), (2), and (3) can be written as coupled equations in ψ\psi, vθv_{\theta}, and bb,

tΔhψ+𝒥(rψ,Δhψr)\displaystyle\partial_{t}\Delta_{h}\psi+\mathcal{J}^{\odot}\left(r\psi,\frac{\Delta_{h}\psi}{r}\right) =\displaystyle= 2rvθzvθ+rb,\displaystyle-\frac{2}{r}v_{\theta}\partial_{z}v_{\theta}+\partial_{r}b, (6)
tvθ+1r2𝒥(rψ,rvθ)\displaystyle\partial_{t}v_{\theta}+\frac{1}{r^{2}}\mathcal{J}^{\odot}(r\psi,rv_{\theta}) =\displaystyle= 0,\displaystyle 0, (7)
tb+1r𝒥(rψ,b)\displaystyle\partial_{t}b+\frac{1}{r}\mathcal{J}^{\odot}(r\psi,b) =\displaystyle= N21rr(rψ),\displaystyle-N^{2}\frac{1}{r}\partial_{r}(r\psi), (8)

where the truncated Laplacian Δhψ\Delta_{h}\psi is defined by

Δhψ=2ψz2+r(1r(rψ)r)=2ψz2+2ψr2+1rψrψr2,\Delta_{h}\psi=\dfrac{\partial^{2}\psi}{\partial z^{2}}+\dfrac{\partial}{\partial r}\left(\frac{1}{r}\dfrac{\partial(r\psi)}{\partial r}\right)=\dfrac{\partial^{2}\psi}{\partial z^{2}}+\dfrac{\partial^{2}\psi}{\partial r^{2}}+\frac{1}{r}\dfrac{\partial\psi}{\partial r}-\frac{\psi}{r^{2}}, (9)

and the cylindrical Jacobian 𝒥\mathcal{J}^{\odot} of two functions ff and gg is given by

𝒥(f,g)=frgzgrfz.\mathcal{J}^{\odot}(f,g)=\dfrac{\partial f}{\partial r}\dfrac{\partial g}{\partial z}-\dfrac{\partial g}{\partial r}\dfrac{\partial f}{\partial z}. (10)

II.2 Solution of the Linear Problem

The system of equations (6), (7), and (8) can be linearised by setting all Jacobians and the vθzvθv_{\theta}\partial_{z}v_{\theta} terms equal to zero, leading to the linear problem

tΔhψ\displaystyle\partial_{t}\Delta_{h}\psi =\displaystyle= rb,\displaystyle\partial_{r}b, (11)
tvθ\displaystyle\partial_{t}v_{\theta} =\displaystyle= 0,\displaystyle 0, (12)
tb\displaystyle\partial_{t}b =\displaystyle= N21rr(rψ).\displaystyle-N^{2}\frac{1}{r}\partial_{r}(r\psi). (13)

Equation (12) is self-consistent, and yields an orthoradial velocity that does not evolve in time. On the other hand, cross-derivatives of equations (11) and (13) give a linear equation in ψ\psi

[ψ]=0,\mathcal{L}[\psi]=0, (14)

where \mathcal{L} is an operator defined as

[ψ]=t2Δhψ+N2r(1rr(rψ)).\mathcal{L}[\psi]=\partial^{2}_{t}\Delta_{h}\psi+N^{2}\partial_{r}\left(\frac{1}{r}\partial_{r}(r\psi)\right). (15)

As discussed in Maurer et al. maurer2017 and in Boury et al. (boury2018, ), the linear solution of equations (12) and (14) that fulfills the boundary conditions imposed by an axisymmetric forcing at the surface and zero velocities normal to the other sides, with no orthoradial velocity, is

ψ(r,z,t)\displaystyle\psi(r,z,t) =\displaystyle= ψ0J1(lr)cos(ωtmz),\displaystyle\psi_{0}J_{1}(lr)\cos(\omega t-mz), (16)
vθ(r,z,t)\displaystyle v_{\theta}(r,z,t) =\displaystyle= 0,\displaystyle 0, (17)
b(r,z,t)\displaystyle b(r,z,t) =\displaystyle= N2lωψ0J0(lr)sin(ωtmz),\displaystyle-\frac{N^{2}l}{\omega}\psi_{0}J_{0}(lr)\sin(\omega t-mz), (18)

with ω\omega the frequency of the wave field, ll and mm the radial and the vertical wave numbers, and ψ0\psi_{0} a constant amplitude. The radial structure of the wave field is given by Bessel functions of zeroth and first orders, namely J0J_{0} and J1J_{1}. With the proposed solution (16) for ψ\psi, the linear operator (15) gives the dispersion relation for gravity waves

(l2+m2)ω2=N2l2.(l^{2}+m^{2})\omega^{2}=N^{2}l^{2}. (19)

II.3 Non-Linear Problem

Back to equations (6), (7), and (8), the same cross-derivation that led to the linear system can be used to derive the fully non-linear equation

[ψ]=𝒩[ψ,vθ,b],\mathcal{L}[\psi]=\mathcal{N}[\psi,v_{\theta},b], (20)

where \mathcal{L} is still defined by equation (15) and 𝒩\mathcal{N} is the non-linear operator

𝒩[ψ,vθ,b]=t𝒥(rψ,Δhψr)r(1r𝒥(rψ,b))1rtzvθ2.\mathcal{N}[\psi,v_{\theta},b]=-\partial_{t}\mathcal{J}^{\odot}\left(r\psi,\frac{\Delta_{h}\psi}{r}\right)-\partial_{r}\left(\frac{1}{r}\mathcal{J}^{\odot}(r\psi,b)\right)-\frac{1}{r}\partial_{t}\partial_{z}v_{\theta}^{2}. (21)

As discussed in the appendix of boury2019 , when the velocity amplitudes become too large, these non-linear terms cannot be neglected. If we consider the linear solution, we note that vθ=0v_{\theta}=0 so 𝒩[ψ,vθ,b]=𝒩[ψ,b]\mathcal{N}[\psi,v_{\theta},b]=\mathcal{N}[\psi,b]. The non-linear right-hand side of equation (20) therefore becomes only dependent on two Jacobians that can be evaluated. Using that nist2010

(J0(lr))r=lJ1(lr),(rJ1(lr))r=lrJ0(lr),and(r1J1(lr))r=lr1J2(lr),\dfrac{\partial(J_{0}(lr))}{\partial r}=-lJ_{1}(lr),\mathrm{\leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ }\dfrac{\partial(rJ_{1}(lr))}{\partial r}=lrJ_{0}(lr),\mathrm{\leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ and\leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ }\dfrac{\partial(r^{-1}J_{1}(lr))}{\partial r}=-lr^{-1}J_{2}(lr), (22)

and ψ\psi and bb as derived in equations (16) and (18), the non-linear terms involved in 𝒩[ψ,b]\mathcal{N}[\psi,b] become

t𝒥(rψ,Δhψr)\displaystyle\partial_{t}\mathcal{J}^{\odot}\left(r\psi,\frac{\Delta_{h}\psi}{r}\right) =\displaystyle= C[J1(lr)]2rcos(2ωt2mz),\displaystyle C\frac{\left[J_{1}(lr)\right]^{2}}{r}\cos(2\omega t-2mz), (23)
r(1r𝒥(rψ,b))\displaystyle\partial_{r}\left(\frac{1}{r}\mathcal{J}^{\odot}(r\psi,b)\right) =\displaystyle= C[cos2(ωtmz)J0(lr)rJ0(lr)+sin2(ωtmz)J1(lr)rJ1(lr)],\displaystyle C\left[\cos^{2}(\omega t-mz)J_{0}(lr)\partial_{r}J_{0}(lr)+\sin^{2}(\omega t-mz)J_{1}(lr)\partial_{r}J_{1}(lr)\right], (24)

with

C=2ω(l2+m2)mψ02.C=2\omega(l^{2}+m^{2})m\psi_{0}^{2}. (25)

As a result, the non-linear right-hand side of equation (20) is

𝒩[ψ,b]=CJ1(lr)[J2(lr)sin2(ωtmz)J1(lr)cos2(ωtmz)]0\mathcal{N}[\psi,b]=CJ_{1}(lr)\left[J_{2}(lr)\sin^{2}(\omega t-mz)-J_{1}(lr)\cos^{2}(\omega t-mz)\right]\neq 0 (26)

We can go a step further if, in the general case, inspired by the work of Thorpe thorpe1966 , we write ψ\psi as a sum over possible solutions of various frequencies ωj\omega_{j} as

ψ(r,z,t)=jχj(r,z)cos(ωjt),\psi(r,z,t)=\sum_{j}\chi_{j}(r,z)\cos(\omega_{j}t), (27)

very much like a discrete Fourier Transform, with χj\chi_{j}, jj\in\mathbb{Z}, a spatial function. The linear part of equation (20) is therefore

[ψ]=j[(N2ωj2)r(1r(rχj)r)2χjz2]cos(ωjt).\mathcal{L}[\psi]=\sum_{j}\left[(N^{2}-\omega_{j}^{2})\dfrac{\partial}{\partial_{r}}\left(\frac{1}{r}\dfrac{\partial(r\chi_{j})}{\partial_{r}}\right)-\dfrac{\partial^{2}\chi_{j}}{\partial z^{2}}\right]\cos(\omega_{j}t). (28)

As the non-linear terms are only second order products of the stream function, we assume the following development

𝒩[ψ]=g,hCg,h(r,z)cos((ωg+ωh)t),\mathcal{N}[\psi]=\sum_{g,h}C_{g,h}(r,z)\cos((\omega_{g}+\omega_{h})t), (29)

where ωg\omega_{g} and ωh\omega_{h} are frequencies, and Cg,hC_{g,h} is a spatial fonction. Projection of equation (21) over frequencies leads to

j,(N2ωj2)r(1r(rχj)r)2χjz2=Cj,\forall j\in\mathbb{Z},\leavevmode\nobreak\ (N^{2}-\omega_{j}^{2})\dfrac{\partial}{\partial_{r}}\left(\frac{1}{r}\dfrac{\partial(r\chi_{j})}{\partial_{r}}\right)-\dfrac{\partial^{2}\chi_{j}}{\partial z^{2}}=C_{j}, (30)

where CjC_{j} is defined as the function Cg,hC_{g,h} where gg and hh verify ωj=ωg+ωh\omega_{j}=\omega_{g}+\omega_{h}. The CjC_{j} functions act as forcing terms produced by non-linear wave-wave interactions. The impact of this forcing term can be further explored theoretically by using Green functions (see appendix) and a Taylor expansion of the stream function, or with Direct Numerical Simulations, which is beyond the scope of this study.

II.4 Wave Resonator and Mode Selection

Cylindrical cavities have the ability to produce enhanced modal wave fields, likely to trigger instabilities at high frequencies boury2018 . In such a confined configuration, radial and vertical modes can be excited if they satisfy the boundary condition, namely a zero orthogonal velocity at the boundaries. Hence, a mode can be described by the stream function

ψ(r,z,t)=ψ0J1(lr)sin(mz)cos(ωt).\psi(r,z,t)=\psi^{0}J_{1}(lr)\sin(mz)\cos(\omega t). (31)

As we will see next, the radial confinement sets the possible values of ll, and so does the vertical confinement for the values of mm.

II.4.1 Radial Confinement

The cylinder imposes a radial confinement within a radius RR. As such, modes of radial wave number lp,pl_{p},\leavevmode\nobreak\ p\in\mathbb{N}^{*}, are selected with the boundary condition stating that the radial velocity is zero at the boundary

vr(r=R,z,t)=(1r(rψ)z)r=R=0,v_{r}(r=R,z,t)=\left(-\frac{1}{r}\dfrac{\partial(r\psi)}{\partial z}\right)_{r=R}=0, (32)

equivalent to

J1(lpR)=0.J_{1}(l_{p}R)=0. (33)

If j1,pj_{1,p} is the ppth zero of the 11st order Bessel function J1J_{1}, then the values of lpl_{p} are given by

p,lp=j1,pR.\forall p\in\mathbb{N}^{*},\leavevmode\nobreak\ l_{p}=\frac{j_{1,p}}{R}. (34)

Table 1 presents the first zeros of the J1J_{1} Bessel function, extracted from beattie1958 , and the corresponding radial wave numbers for a cylindrical confinement of radius R=20cmR=20\mathrm{\leavevmode\nobreak\ cm}. An experimental visualisation of these radial modes for p=1p=1, 22, and 33, can be found in Boury et al. boury2018 .

pp 11 22 33 44 55
j1,pj_{1,p} 3.833.83 7.027.02 10.1710.17 13.3213.32 16.4716.47
lpl_{p} 19.1519.15 35.1035.10 50.8550.85 66.6066.60 82.3582.35
Table 1: First zeros of the J1J_{1} Bessel function and associated radial wave number for R=20cmR=20\mathrm{\leavevmode\nobreak\ cm}.

II.4.2 Vertical Confinement

The upper and lower boundaries, set by the bottom of the tank and the wave generator, respectively, impose a vertical confinement (cavity) over a given height LL. The vertical modes are found by stating that, at z=0z=0 (when neglecting the low amplitude generator motion) and at z=Lz=-L, the vertical velocity is zero, i.e.

vz(z=0)=(1r(rψ)r)z=0=0andvz(z=L)=(1r(rψ)r)z=L=0,v_{z}(z=0)=\left(\frac{1}{r}\dfrac{\partial(r\psi)}{\partial r}\right)_{z=0}=0\mathrm{\leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ and\leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ }v_{z}(z=-L)=\left(\frac{1}{r}\dfrac{\partial(r\psi)}{\partial r}\right)_{z=-L}=0, (35)

which justifies that the vertical dependence of ψ\psi is described by a sine function, and also means that

sin(mL)=0.\sin(mL)=0. (36)

As a consequence, vertical modes should have a half-integer number of wave lengths in the cavity. It follows that the vertical wave number mq,qm_{q},\leavevmode\nobreak\ q\in\mathbb{N}^{*}, is given by

q,mq=πqL.\forall q\in\mathbb{N}^{*},\leavevmode\nobreak\ m_{q}=\frac{\pi q}{L}. (37)

Table 2 gives the smallest values of mqm_{q} that can be found in a resonant cavity of height L=60cmL=60\mathrm{\leavevmode\nobreak\ cm}.

qq 11 22 33 44 55 66 77 88 99
mqm_{q} 5.235.23 10.4710.47 15.7015.70 20.9320.93 26.1726.17 31.4031.40 36.6336.63 41.8741.87 47.147.1
Table 2: Lowest vertical wave numbers mqm_{q} for L=60cmL=60\mathrm{\leavevmode\nobreak\ cm}.

II.4.3 Cavity Modes

Each mode in the cavity is designated by a couple (p,q)2(p,q)\in\mathbb{N}^{*2}, so that its radial wave number is lpl_{p} and its vertical wave number is mqm_{q}. Consequently, a mode (p,q)(p,q) has a given frequency ωp,q\omega_{p,q} fixed by the linear dispersion relation for internal waves as

ωp,qN=(lp2lp2+mq2)1/2.\frac{\omega_{p,q}}{N}=\left(\frac{l_{p}^{2}}{l_{p}^{2}+m_{q}^{2}}\right)^{1/2}. (38)

We present in table 3 the values of ωp,q/N\omega_{p,q}/N corresponding to the lowest modes (p,q)(p,q) in a cavity of radius R=20cmR=20\mathrm{\leavevmode\nobreak\ cm} and height L=60cmL=60\mathrm{\leavevmode\nobreak\ cm}.

ωp,q/N\omega_{p,q}/N q=1\leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ q=1\leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ q=2\leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ q=2\leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ q=3\leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ q=3\leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ q=4\leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ q=4\leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ q=5\leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ q=5\leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ q=6\leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ q=6\leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ q=7\leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ q=7\leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ q=8\leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ q=8\leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ q=9\leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ q=9\leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\
p=1p=1 0.96470.9647 0.87740.8774 0.77330.7733 0.67500.6750 0.59050.5905 0.52070.5207 0.46330.4633 0.41590.4159 0.37660.3766
p=2p=2 0.98910.9891 0.95830.9583 0.91280.9128 0.85890.8589 0.80170.8017 0.74530.7453 0.69190.6919 0.64240.6424 0.59750.5975
p=3p=3 0.99480.9948 0.97950.9795 0.95550.9555 0.92470.9247 0.88920.8892 0.85090.8509 0.81140.8114 0.77200.7720 0.73360.7336
p=4p=4 0.99690.9969 0.98790.9879 0.97330.9733 0.95400.9540 0.93070.9307 0.90450.9045 0.87620.8762 0.84660.8466 0.81650.8165
p=5p=5 0.99800.9980 0.99200.9920 0.98230.9823 0.96920.9692 0.95300.9530 0.93440.9344 0.91370.9137 0.89140.8914 0.86800.8680
Table 3: Values of ωp,q\omega_{p,q} corresponding to modes (p,q)(p,q) of radial wave number lpl_{p} and vertical wave number mqm_{q} for R=20cmR=20\mathrm{\leavevmode\nobreak\ cm} and L=60cmL=60\mathrm{\leavevmode\nobreak\ cm}.

As shown in equation (29), a non-linear wave-wave interaction produces a wide range of waves that can be projected over the appropriate basis of linear solutions (31). With non-zero non-linear terms, as shown in equation (26), harmonic modes can be fed by an excitation mode. A mode at frequency ω\omega is therefore in resonance with its nnth harmonic, nn\in\mathbb{N}^{*}, if there exists a couple of non-zero integers (p,q)(p,q) such that

nω=ωp,q.n\omega=\omega_{p,q}. (39)

In particular, a mode excited at frequency ω\omega is in non-linear resonance with its first harmonic if twice the forcing frequency 2ω2\omega is close to a frequency ωp,q\omega_{p,q} that corresponds to a mode (p,q)(p,q). We will verify that in the following sections.

III Experimental Apparatus

Our experiments were conducted in the experimental apparatus described in Boury et al. boury2018 and Boury et al. boury2019 , adapted from the setup of Maurer et al maurer2017 . A general schematic of the experimental device is presented in figure 1. The system is described using natural cylindrical coordinates with the origin taken at the surface of the water at the center of the tank.

Refer to caption
Figure 1: Schematic of the experimental apparatus. Left: a cylindrical tank, inside a square tank, confines the waves produced by the generator located at the surface, leading to a radial Bessel mode propagating downwards. Right: linear stratification measured in the experiments. Vertical dimension of the generator is not to scale.

The generator comprises sixteen, 12mm12\mathrm{\leavevmode\nobreak\ mm} thick, concentric PVC cylinders periodically oscillating, each of them being forced by two eccentric cams. The eccentricities can be configured to introduce a phase shift between the different cylinders, and the oscillating amplitude can be set for each individual cylinder. As a result, the vertical displacement of the nthn^{th} cylinder can be described by

an(t)=Ancos(ωt+αn),a_{n}(t)=A_{n}\cos(\omega t+\alpha_{n}), (40)

with AnA_{n} its amplitude, ω\omega the forcing frequency, and αn\alpha_{n} a phase shift. For a smooth motion of the PVC cylinders, a 1mm1\mathrm{\leavevmode\nobreak\ mm} gap is kept between each cylinder and the total diameter of the wave generator is then 402mm402\mathrm{\leavevmode\nobreak\ mm}. The generator is mounted at the surface of the water to force downwards internal waves. The wave field is forced using a mode 11 profile of radial wave number l=19m1l=19\mathrm{\leavevmode\nobreak\ m^{-1}}, for which amplitudes of each cylinder are presented in table 4. This profile has been proven efficient to generate axisymmetric Bessel-shaped wave fields boury2018 .

Cams 11 22 33 44 55 66 77 88 99 1010 1111 1212 1313 1414 1515 1616
Mode 11 amplitudes (mm)\mathrm{(mm)} 2.52.5 2.42.4 2.32.3 2.12.1 1.91.9 1.61.6 1.31.3 0.90.9 0.60.6 0.20.2 0.1-0.1 0.3-0.3 0.6-0.6 0.8-0.8 0.9-0.9 1-1
Table 4: Amplitudes of the different cams of the generator for a mode 11 radial profile. The first cam is located at r=0r=0.

Experiments were conducted in a transparent cylindrical acrylic tank of the same diameter as the generator, set into a square acrylic tank to prevent the experiment visualisation suffering from optical deformations that would occur due to curved interfaces. Both tanks were filled with salt-stratified water with the same density profile. We used the double-bucket method to fill the tanks with a linear stratification fortuin1960 ; oster1963 . Density and buoyancy were measured as a function of depth using a calibrated PME conductivity and temperature probe mounted on a motorised vertical axis. Buoyancy frequency was estimated from the mean value of the NN profile obtained from the density function ρ(z)\rho(z) measured from the free surface to within a couple of centimeters of the bottom of the tank, due to the construction of the probe. The wave generator was immersed at a depth of 2cm2\mathrm{\leavevmode\nobreak\ cm} into the stratification. The error on the buoyancy frequency was estimated using the standard deviation of this NN profile, which was about 4%4\% of the estimated NN value (see boury2018 for more details). We obtained a buoyancy frequency of N=0.89±0.06rads1N=0.89\pm 0.06\mathrm{\leavevmode\nobreak\ rad\cdot s^{-1}}.

Velocity fields were obtained via Particle Image Velocimetry (PIV). A laser sheet was created by a laser beam (Ti:Sapphire, 2watts2\mathrm{\leavevmode\nobreak\ watts}, wavelength 532nm532\mathrm{\leavevmode\nobreak\ nm}) going through a cylindrical lens. It could be oriented either horizontally (to measure the radial and orthoradial velocity) or vertically (to measure the vertical and the radial velocity). For the purpose of visualisation, 10μm10\mathrm{\leavevmode\nobreak\ \mu m} diameter hollow glass spheres of volumetric mass 1.1kgL11.1\mathrm{\leavevmode\nobreak\ kg\cdot L^{-1}} were added to the fluid while filling the tank. To obtain good quality velocity fields near the bottom of the tank and while imaging in a horizontal plane, 10μm10\mathrm{\leavevmode\nobreak\ \mu m} silver-covered spheres of volumetric mass 1.4kgL11.4\mathrm{\leavevmode\nobreak\ kg\cdot L^{-1}} were added when needed in some experiments. Images were recorded at 4Hz4\mathrm{\leavevmode\nobreak\ Hz} and data processing of the PIV raw images was done using the CIVx algorithm fincham2000 .

IV Results

IV.1 Axisymmetric Non-Linear Wave Generation

We consider a set of nine experiments indexed from 0 to 99, with forcing frequency from ω/N=0.305\omega/N=0.305 to 0.4490.449, using a mode 11 configuration at the generator boury2018 . For each frequency, a 1010 minute forcing leads to non-linear wave-wave interactions, where higher frequency waves are created. This phenomenon is illustrated in figure 2, showing the frequency spectrum from experiment number 99, forced at ω/N=0.449\omega/N=0.449. The spectrum contains not only the forcing frequency at ω/N\omega/N, but also a mean flow at ω=0\omega=0 and several harmonics at 2ω/N2\omega/N, 3ω/N3\omega/N, and so on. Here, the first harmonic would be propagating as ω/N<0.5\omega/N<0.5 and therefore 2ω/N<12\omega/N<1. This behaviour is observed for all frequencies, regardless of any resonant cavity aspect identified in Boury et al. boury2018 . In the performed experiments, due to the range of frequencies choosen, the first harmonic at 2ω/N2\omega/N is always propagating and, except for experiments 11 and 22, the second harmonic at 3ω/N3\omega/N is evanescent.

Refer to caption
Figure 2: Example of Fourier transform performed over the last two minutes of experiment 99 with a forcing at ω/N=0.449\omega/N=0.449. The solid line shows the cut-off frequency at NN. Dashed, dashed-dotted, and dotted lines show ω\omega, 2ω2\omega, and 3ω3\omega frequencies.

Figure 3 gives snapshot examples of such an experiment, for a forcing at ω/N=0.449\omega/N=0.449, corresponding to the spectrum presented in figure 2. The first row shows the wave field filtered at the forcing frequency ω\omega with, from left to right, vzv_{z} and vrv_{r} in the vertical plane, and vrv_{r} and vθv_{\theta} in the horizontal plane. The excited wave field corresponds to a mode 11 and presents all the features of an axisymmetric wave field boury2018 : right-left symmetry of vzv_{z} and right-left antisymmetry of vrv_{r} in the vertical plane, invariance by rotation of center (0,0)(0,0) for vrv_{r} in the horizontal plane, and a negligible orthoradial velocity vθv_{\theta} when observed in the horizontal plane (vrv_{r} dominates vθv_{\theta} by a factor 10\sim 10 in magnitude).

The first harmonic created at 2ω/N2\omega/N is filtered and shown in the second row of figure 3. Though no spatial wave lengths can be directly infered from the wave field, we notice that the created harmonic field shares the same axisymmetric properties as the excitation wave field. The right-left symmetry/antisymmetry of vzv_{z} and vrv_{r} in the vertical plane is also consistent with an axisymmetric description of the wave field. More specifically, a zero of radial velocity is observed at the center of the experimental domain whereas filtered wave fields in Triadic Resonant Instability (TRI) maurerPhD have shown, in some cases, non-zero velocity, breaking the axisymmetry of the system. The orthoradial velocity vθv_{\theta} is, as compared to the excitation wave field, a noisy signal and does not show any sign of symmetry breaking.

Refer to caption
Figure 3: Velocity fields in experiment 99, at forcing frequency ω/N=0.449\omega/N=0.449. The first row shows the filtered wave field at ω\omega and the second row shows the filtered wave field at 2ω2\omega. From left to right: vzv_{z}, vrv_{r} in a vertical cross-section, and vrv_{r}, vθv_{\theta} in a horizontal cross section. For the purpose of visualisation, negative values of rr are used in the vertical plane views.

IV.2 Proper Orthogonal Decomposition (POD)

In order to describe quantitatively the wave field at a given frequency and, more particularly, the wave field created in the harmonic at twice the forcing frequency, we used a Proper Orthogonal Decomposition (POD) method. As we know the relevant stream functions (modes) in the cavity, the POD process simply consists of projecting the experimental wave field over an orthogonal basis built accordingly.

The series of {rJ0(lpr)|p}\left\{r\mapsto J_{0}(l_{p}r)\leavevmode\nobreak\ |\leavevmode\nobreak\ p\in\mathbb{N}^{*}\right\} and {zsin(mqz)|q}\left\{z\mapsto\sin(m_{q}z)\leavevmode\nobreak\ |\leavevmode\nobreak\ q\in\mathbb{N}^{*}\right\} forms an orthogonal basis in an axisymmetric geometry, which is appropriate for this study. A direct consequence is that the stream function can be written in terms of normalised modes (p,q)(p,q), denoted ψp,q\psi_{p,q}, in the axisymmetric domain (r,θ,z)𝒞=[0;R]×[0; 2π]×[0;L](r,\theta,z)\in\mathcal{C}=[0;\leavevmode\nobreak\ R]\times[0;\leavevmode\nobreak\ 2\pi]\times[0;\leavevmode\nobreak\ -L], defined as

ψp,q(r,z)=ψp,q0J1(lpr)sin(mqz),withψp,q0=1J1(lpR)2Lπ,\psi_{p,q}(r,z)=\psi_{p,q}^{0}J_{1}(l_{p}r)\sin(m_{q}z),\mathrm{\leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ with\leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ }\psi_{p,q}^{0}=\frac{1}{J_{1}^{\prime}(l_{p}R)}\sqrt{\frac{2}{L\pi}}, (41)

satsifying the condition

𝒞[ψp,q(r,z)]2rdrdθdz=1,\int_{\mathcal{C}}\left[\psi_{p,q}(r,z)\right]^{2}r\text{d}r\,\text{d}\theta\,\text{d}z=1, (42)

from which we can derive the vertical and radial velocities associated to a mode (p,q)(p,q), using equations (5), as follows

vzp,q(r,z)\displaystyle v_{z}^{p,q}(r,z) =\displaystyle= 1r(rψp,q)r=lpψp,q0J0(lpr)sin(mqz),\displaystyle\frac{1}{r}\dfrac{\partial(r\psi_{p,q})}{\partial r}=-l_{p}\psi_{p,q}^{0}J_{0}(l_{p}r)\sin(m_{q}z), (43)
vrp,q(r,z)\displaystyle v_{r}^{p,q}(r,z) =\displaystyle= 1r(rψp,q)z=mqψp,q0J1(lpr)cos(mqz).\displaystyle-\frac{1}{r}\dfrac{\partial(r\psi_{p,q})}{\partial z}=m_{q}\psi_{p,q}^{0}J_{1}(l_{p}r)\cos(m_{q}z). (44)

Hence, the kinetic energy contained in one mode (p,q)(p,q) is given by

Kp,q=[𝒞vz(r,z)vzp,q(r,z)rdrdθdz]2+[𝒞vr(r,z)vrp,q(r,z)rdrdθdz]2,K_{p,q}=\left[\int_{\mathcal{C}}v_{z}(r,z)\cdot v_{z}^{p,q}(r,z)r\text{d}r\,\text{d}\theta\,\text{d}z\right]^{2}+\left[\int_{\mathcal{C}}v_{r}(r,z)\cdot v_{r}^{p,q}(r,z)r\text{d}r\text{d}\,\theta\,\text{d}z\right]^{2}, (45)

which, accounting for the discretisation of our domain, can be written as follows

Kp,q=[r,z2πrvz(r,z)vzp,q(r,z)]2+[r,z2πrvr(r,z)vrp,q(r,z)]2,K_{p,q}=\left[\sum_{r,z}2\pi rv_{z}(r,z)\cdot v_{z}^{p,q}(r,z)\right]^{2}+\left[\sum_{r,z}2\pi rv_{r}(r,z)\cdot v_{r}^{p,q}(r,z)\right]^{2}, (46)

with sums over all spatial grid points. As a result, if we denote K0K_{0} the kinetic energy of the total wave field, the fraction of energy in a mode (p,q)(p,q) is given by the scalar product (46) normalised by K0K_{0}. The higher the quantity, the more dominant the mode is in the observed field. Note that the prefactors for vzp,qv_{z}^{p,q} and vrp,qv_{r}^{p,q} can differ if the normalisation process is directly applied to the stream function or to the velocities, though it will not affect the conclusions.

The left panel of figure 4 shows an example of such a POD decomposition using the excitation wave field previously presented in the top part of figure 3, and the right panel of figure 4 shows the decomposition of the first harmonic wave field, respectively at ω/N=0.449\omega/N=0.449 and 2ω/N=0.8982\omega/N=0.898. Figure 4 (top) depicts the colormap of the energetic distribution onto different modes and figure 4 (bottom) plots the kinetic energy contained in every tested mode as a function of the quantity p.qp.q, defined as p.q=p+0.1qp.q=p+0.1q, used to identify the different modes with a single number or, in other words, to transform the 22D plot of the top row of figure 4 into the more easily quantifiable 11D plot of the bottom of figure 4. As expected, figure 4 (left) shows that nearly all the energy of the wave field produced by the generator lies in a radial mode 11 wave, more exactly a (1,7)(1,7) mode, that can be clearly identified in figure 3 with 77 zeros along the vertical direction and one along the horizontal direction. In addition, while the filtered wave field at 2ω/N2\omega/N in figure 3 does not explicitly display a mode, we see from figure 4 (right) that the energy is mostly split into mode (1,2)(1,2) (47%47\%), mode (3,5)(3,5) (32%32\%), and mode (5,8)(5,8) (8%8\%), the other contributions being negligible. The first harmonic generated in experiment 99 can therefore be described as a sum of modes (1,2)(1,2), (3,5)(3,5), and (5,8)(5,8), with given prefactors to account for the distribution of energy between the three modes.

Refer to caption
Figure 4: POD performed on experiment number 99. The top row is the colormap of the energy distribution in the different modes in the (p,q)(p,q) plane, and the bottom row is the transposition of this 22D plot into a 11D plot of axis p.q=p+0.1qp.q=p+0.1q. Left: POD over the filtered wave field at the excitation frequency ω/N=0.449\omega/N=0.449. Right: POD over the filtered wave field at the first harmonic frequency 2ω/N=0.8982\omega/N=0.898.

IV.3 Harmonics and Mode Selection

The POD discrimination process is applied to experiments 11 to 99 and its results are summarised in table 5. In some cases, the first harmonic generated by the non-linear wave-wave interaction can be clearly identified by eye as a single mode (p,q)(p,q), as shown for experiments 55 (with 2ω/N=0.7552\omega/N=0.755, see figure 5) and 66 (with 2ω/N=0.7912\omega/N=0.791, see figure 6), with the generation of a mode (1,3)(1,3) and a mode (2,6)(2,6), respectively. In such cases, this clearly observed mode is the one identified as the most energetic in the POD decomposition. Figure 7 illustrates this point with experiment 55 and 66, showing a single mode containing about 8080 to 90%90\% of the total kinetic energy of the first harmonic. The remaining fraction of energy, from 1010 to 20%20\%, is sparsely distributed into lower modes that contain less than 5%5\% of the total energy each and do not contribute significantly to the general form and behaviour of the wave field. In other cases, as described in the previous section for experiment 99, the energy is more evenly distributed between several modes.

Refer to caption
Figure 5: Velocity fields in experiment 66, with 2ω/N=0.8002\omega/N=0.800, with generation of a mode (1,3)(1,3).
Refer to caption
Figure 6: Velocity fields in experiment 55, with 2ω/N=0.7652\omega/N=0.765, with generation of a mode (2,6)(2,6).
Refer to caption
Figure 7: POD projection for experiments 55 (left) and 66 (right). The top row is the colormap of the energy distribution in the different modes in the (p,q)(p,q) plane, and the bottom row is the transposition of this 22D plot into a 11D plot of axis p.q=p+0.1qp.q=p+0.1q.

The structure of the generated harmonic modes does not always match the one of the excitation mode. For example, experiments 33 and 66 lead to generation of a dominant mode with a radial structure similar to the excitation wave field (l1=19m1l_{1}=19\mathrm{\leavevmode\nobreak\ m^{-1}}), whereas experiments 22, 55, and 88, generate a dominant higher order radial structure (l2=35m1l_{2}=35\mathrm{\leavevmode\nobreak\ m^{-1}}). For a given radial structure, we note that the vertical wave number increases with the frequency 2ω/N2\omega/N, consistently with the dispersion relation (19). Interestingly, the mode frequencies are not ordered by lower pp or lower qq but seem randomly distributed: for example, by increasing ω\omega, the harmonics can be dominated by a radial mode 11, then a radial mode 22, and then a radial mode 11 again.

Experiment ω/N\leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \omega/N\leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ 2ω/N\leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ 2\omega/N\leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ (p,q)[%](p,q)\leavevmode\nobreak\ [\%]\leavevmode\nobreak\
11 0.3050.305 0.6110.611 (1,5)[61%](1,5)\leavevmode\nobreak\ [61\%]\leavevmode\nobreak\ (2,9)[18%](2,9)\leavevmode\nobreak\ [18\%]\leavevmode\nobreak\
22 0.3230.323 0.6470.647 (2,8)[61%](2,8)\leavevmode\nobreak\ [61\%]\leavevmode\nobreak\ (1,4)[10%](1,4)\leavevmode\nobreak\ [10\%]\leavevmode\nobreak\
33 0.3410.341 0.6830.683 (1,4)[89%](1,4)\leavevmode\nobreak\ [89\%]\leavevmode\nobreak\
44 0.3590.359 0.7190.719 (1,4)[22%](1,4)\leavevmode\nobreak\ [22\%]\leavevmode\nobreak\ (2,7)[21%](2,7)\leavevmode\nobreak\ [21\%]\leavevmode\nobreak\
55 0.3770.377 0.7550.755 (2,6)[82%](2,6)\leavevmode\nobreak\ [82\%]\leavevmode\nobreak\
66 0.3950.395 0.7910.791 (1,3)[78%](1,3)\leavevmode\nobreak\ [78\%]\leavevmode\nobreak\
77 0.4130.413 0.8270.827 (2,5)[35%](2,5)\leavevmode\nobreak\ [35\%]\leavevmode\nobreak\ (1,3)[14%](1,3)\leavevmode\nobreak\ [14\%]\leavevmode\nobreak\ (3,7)[8%](3,7)\leavevmode\nobreak\ [8\%]\leavevmode\nobreak\
88 0.4310.431 0.8630.863 (2,4)[69%](2,4)\leavevmode\nobreak\ [69\%]\leavevmode\nobreak\ (3,6)[9%](3,6)\leavevmode\nobreak\ [9\%]\leavevmode\nobreak\
99 0.4490.449 0.8990.899 (1,2)[47%](1,2)\leavevmode\nobreak\ [47\%]\leavevmode\nobreak\ (3,5)[32%](3,5)\leavevmode\nobreak\ [32\%]\leavevmode\nobreak\ (5,8)[8%](5,8)\leavevmode\nobreak\ [8\%]\leavevmode\nobreak\
Table 5: Forcing frequencies and their first harmonics, with identified dominating modes (in %\% of kinetic energy) using POD.
ωp,q/N\omega_{p,q}/N q=1\leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ q=1\leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ q=2\leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ q=2\leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ q=3\leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ q=3\leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ q=4\leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ q=4\leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ q=5\leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ q=5\leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ q=6\leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ q=6\leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ q=7\leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ q=7\leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ q=8\leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ q=8\leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ q=9\leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ q=9\leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\
p=1p=1 0.96470.9647 0.8774\mathbf{0.8774} [9] 0.7733\mathbf{0.7733} [6,7] 0.6750\mathbf{0.6750} [2,3,4] 0.5905\mathbf{0.5905} [1] 0.52070.5207 0.46330.4633 0.41590.4159 0.37660.3766
p=2p=2 0.98910.9891 0.95830.9583 0.91280.9128 0.8589\mathbf{0.8589} [8] 0.8017\mathbf{0.8017} [7] 0.7453\mathbf{0.7453} [5] 0.6919\mathbf{0.6919} [4] 0.6424\mathbf{0.6424} [2] 0.5975\mathbf{0.5975} [1]
p=3p=3 0.99480.9948 0.97950.9795 0.95550.9555 0.92470.9247 0.8892\mathbf{0.8892} [9] 0.8509\mathbf{0.8509} [8] 0.8114\mathbf{0.8114} [7] 0.77200.7720 0.73360.7336
p=4p=4 0.99690.9969 0.98790.9879 0.97330.9733 0.95400.9540 0.93070.9307 0.90450.9045 0.87620.8762 0.84660.8466 0.81650.8165
p=5p=5 0.99800.9980 0.99200.9920 0.98230.9823 0.96920.9692 0.95300.9530 0.93440.9344 0.91370.9137 0.8914\mathbf{0.8914} [9] 0.86800.8680
Table 6: Reproduction of table 3, showing the modes (p,q)(p,q) identified in the POD. Bold red values of ωp,q/N\omega_{p,q}/N are the theoretical frequencies corresponding to the modes observed in the experiments. The associated experiments are indicated between brackets.

Table 6 reproduces the theoretical results of table 3, highlighting the frequencies corresponding to the observed modes of experiments 11 to 99, which can be compared to the values in table 5. This comparison shows that, when a dominant mode (p,q)(p,q) is observed, the frequency of the harmonic (third column in table 5) is close to the frequency of the same mode (p,q)(p,q) stated by the dispersion relation (table 6). For instance, in experiment 66, 78%78\% of the energy of the super-harmonic is in a mode (1,3)(1,3), which is the mode observed in figure 5, and the frequency predicted for such a mode is ω1,3/N=0.773\omega_{1,3}/N=0.773 according to table 6, close to the harmonic frequency observed in the experiment 2ω/N=0.7912\omega/N=0.791 (table 5). Similarly, in experiment 55, 82%82\% of the energy of the super-harmonic is in a mode (2,6)(2,6), which is the mode observed in figure 6, and the frequency predicted for such a mode is ω2,6/N=0.745\omega_{2,6}/N=0.745, also close to the harmonic frequency observed in the experiment 2ω/N=0.7552\omega/N=0.755.

To generalize this observation, we summarised our results in figure 8. The blue circles show, in the phase space (x=ω/N,y=p.q)(x=\omega/N,y=p.q), all the theoretical cavity modes that can be created according to table 3 (for p=1p=1 through 55 and q=1q=1 through 99). In parallel, each first harmonic frequency for experiment 11 to 99 is represented by a vertical dashed line at constant ω/N\omega/N. At each of these frequencies, modes extracted from the POD projection of the filtered harmonic wave fields are represented by red dots in the phase space, where the color intensity shows the energetic contribution of each dominant mode. To guide the eye, black rectangles link together an experimentally observed mode with its theoretical counterpart whose frequency is within 11 to 7%7\% of the experimental frequency in all cases. This margin is of the order of the error on the buoyancy frequency NN measured with the probe (about 4%4\%). On a vertical dashed line, when a single red dot is plotted (see for example experiment 5), it means that the harmonic wave field looks like a single mode which, as shown by the linking rectangle, corresponds to the mode selected by the non-linear interaction. On the contrary, when several dots appear for a given harmonic frequency (see, for example, experiment 1), resonant conditions stated by the dispersion relation are not exactly fulfilled and the harmonic wave field is a combination of two or more modes, the system being unable to select a solitary one.

Refer to caption
Figure 8: Selected modes vs. frequency. The blue circles show the frequencies and mode numbers of the cavity modes (p,q)(p,q) (from table 3). The red dashed lines show the first harmonics (2ω/N)(2\omega/N) seen in the 99 conducted experiments, which numbers are indicated above each line. On each line, red dots indicate the modes identified with POD, darker red meaning a more energetic mode. For each experimentally observed mode (dots), the corresponding theoretical mode (circles) is indicated by a linking rectangle. The error on NN is illustrated by an error bar in experiment 11.

V Conclusions and Discussion

This study presents, to our knowledge, the first experimental evidence of internal wave harmonics created by weakly non-linear internal wave self-interaction in a linear stratification. We have presented a configuration in which, as the excited wave field has a frequency ω\omega below N/2N/2, the first harmonic created can always propagate and we have observed that the experimental PIV velocity fields display coherent structures when filtered at the harmonic frequency 2ω/N2\omega/N. With a simple analytical argument, we have shown that the confined domain provides restrictive boundary conditions so that the non-linear wave field has to satisfy a given set of constraints, leading to a discrete collection of radial and vertical wave numbers that can be theoretically predicted. The modes, or standing waves, defined by these relations constitute an excellent description of the first harmonic field in terms of a sum of several contributions, sometimes with a clear dominant one.

Interestingly, self-interacting non-linear terms (the Jacobians) are null when computed using plane waves in Cartesian geometry. Hence, such a non-linear interaction cannot exist in linearly stratified fluids and, to obtain super-harmonics as in baker2020 , a non-linear buoyancy frequency profile is required so that non-linearities can excite waves at higher frequencies through a second order forcing term. This study shows that for pure axisymmetric wave fields, however, self-interaction and super-harmonic generation is a common process even within linear stratifications.

Through a Proper Orthogonal Decomposition (POD) technique, we have shown that a mode selection occurs while generating the harmonic wave field. For some frequencies, the first harmonic reproduces almost perfectly a mode, both in the vertical and in the radial directions and, in other cases, the first harmonic is a sum of modal wave fields. Moreover, in all cases, the newly generated wave field remains axisysmmetric. Modes of lowest radial wave numbers seem likely to be selected through this weakly non-linear interaction and, indeed, the most energetic modes are found to be for p=1p=1 and p=2p=2, whereas the few contributions observed at p=3p=3 and beyond are usually small. This behaviour can be explained by the resistance of the stratification against vertical motions, which increases the efficiency of energy transfers to lower radial modes and, especially, to a mode p=1p=1 whose structure is already forced by the wave generator. Note that higher modes, for p>5p>5 and q>9q>9, could also exist in the harmonic wave field but, for clarity, they are not displayed in figure 8. When tested using the POD decomposition, all of these modes showed very little contribution to the overall kinetic energy (about a few percent maximum).

From a theoretical point of view, the exact generation process is not fully understood yet. A more detailed study, using Green functions could be undertaken to derive the exact forcing terms that generate super-harmonics, but falls beyond the scope of this work. Moreover, in such an experiment, other non-linear phenomena such as wave breaking or Triadic Resonant Instability (TRI), might occur. The underlying mechanisms, however, are likely to be different as, for example, they yield to the generation of sub-harmonics in the case of TRI, or they may involve a symmetry breaking maurerPhD and a fully developped cylindrical wave field with potential orthoradial velocities, which are still to be explained.

Acknowledgements:

This work has been partially supported by the ANR through grant ANR-17-CE30-0003 (DisET) and by ONR Physical Oceanography Grant N000141612450. S.B. wants to thank Labex iMust for supporting his research. This work has been achieved thanks to the resources of PSMN from ENS de Lyon.

Appendix A Axisymmetric Green Function and Weakly Non-Linear Asymptotics

Inspired by the approach of voisin2003 ; ermanyuk2011 ; voisin2011 , we compute the Green function of the linear operator \mathcal{L} in cylindrical coordinates (for more details on the calculus, see alastuey2008 ; nist2010 ). In the axisymmetric case, without θ\theta-dependence, the Green function of the linear wave operator (15) can be written as follows

G(𝐫,t;𝐫,t)=G04π2+eiω(tt)+eim(zz)[J1(lr<)Y1(lr>)]dmdω,G(\mathbf{r},t;\mathbf{r^{\prime}},t^{\prime})=\frac{G_{0}}{4\pi^{2}}\int_{-\infty}^{+\infty}e^{i\omega(t-t^{\prime})}\int_{-\infty}^{+\infty}e^{im(z-z^{\prime})}[J_{1}(lr_{<})Y_{1}(lr_{>})]\text{d}m\text{d}\omega, (47)

where G0G_{0} is a normalisation coefficient, and ω\omega, ll, and mm satisfy the dispersion relation (19). The radial dependence is expressed through the first order Bessel functions J1J_{1} and Y1Y_{1}, two analytical solutions of the radial part of the wave equation boury2018 . The variables are noted tt, 𝐫=(r,θ,z)\mathbf{r}=(r,\theta,z), tt^{\prime}, and 𝐫=(r,θ,z)\mathbf{r^{\prime}}=(r^{\prime},\theta^{\prime},z^{\prime}), and we use the notation r<=min(r,r)r_{<}=\min(r,r^{\prime}) and r>=max(r,r)r_{>}=\max(r,r^{\prime}).

From the wave equation and the top boundary forcing, the linear wave field can be obtained as a solution of the problem

{[ψ,b]=0for𝐫𝒞ψ=δ(z)J0(lr)cos(ωt)for𝐫𝒞,\left\{\begin{array}[]{ll}\mathcal{L}[\psi,b]=0&\mathrm{\leavevmode\nobreak\ for\leavevmode\nobreak\ }\mathbf{r}\in\mathcal{C}\\ \psi=\delta(z)J_{0}(lr)\cos(\omega t)&\mathrm{\leavevmode\nobreak\ for\leavevmode\nobreak\ }\mathbf{r}\in\partial\mathcal{C},\end{array}\right. (48)

where 𝒞={(r,θ,z)[0;R]×[0; 2π]×[0;L]}\mathcal{C}=\left\{(r,\theta,z)\in[0;\leavevmode\nobreak\ R]\times[0;\leavevmode\nobreak\ 2\pi]\times[0;\leavevmode\nobreak\ -L]\right\} is the cavity domain and 𝒞\partial\mathcal{C} stands for its boundaries. The forcing at the top boundary is expressed by a Dirac distribution δ(z)\delta(z). Equation  48 admits for solution

ψ(z,r,t)=𝒞×δ(z)J0(lr)cos(ωt)G𝐧(𝐫,t;𝐫,t)dSdt,\psi(z,r,t)=\int_{\partial\mathcal{C}\times\mathbb{R}}\delta(z^{\prime})J_{0}(lr^{\prime})\cos(\omega t^{\prime})\dfrac{\partial G}{\partial\mathbf{n}}(\mathbf{r},t;\mathbf{r^{\prime}},t^{\prime})\text{d}S^{\prime}\text{d}t^{\prime}, (49)

while integrating over the boundary 𝒞\partial\mathcal{C}.

A small parameter ε=ψ0T/L2\varepsilon=\psi_{0}T/L^{2} characterising the contribution of non-linear terms can be defined from the derivation proposed in (boury2019, ) with ψ0\psi_{0} the stream function amplitude, TT and LL characteristic temporal and spatial periods, allowing to write the stream function with a similar development as in Husseini et al. husseini2019 in the following expansion

ψ=εψ(1)+ε2ψ(2)+𝒪(ε3),\psi=\varepsilon\psi^{(1)}+\varepsilon^{2}\psi^{(2)}+\mathcal{O}\left(\varepsilon^{3}\right), (50)

where ψ(1)\psi^{(1)} is the solution to the linear problem previously discussed. Considering that the second order terms come from the non-linear term

𝒩[ψ(1),b(1)]=ε2CJ1(lr)[J2(lr)sin2(ωtmz)J1(lr)cos2(ωtmz)],\mathcal{N}[\psi^{(1)},b^{(1)}]=\varepsilon^{2}CJ_{1}(lr)\left[J_{2}(lr)\sin^{2}(\omega t-mz)-J_{1}(lr)\cos^{2}(\omega t-mz)\right], (51)

at lowest order, the weakly non-linear correction ψ(2)\psi^{(2)} can be found by defining a new Green problem where the non-linear self-interaction of the linear term acts as a forcing field over the whole domain

{[ψ(2),b(2)]=ε2𝒩[ψ(1),b(1)]for𝐫𝒞ψ=0for𝐫𝒞,\left\{\begin{array}[]{ll}\mathcal{L}[\psi^{(2)},b^{(2)}]=\varepsilon^{2}\mathcal{N}[\psi^{(1)},b^{(1)}]&\mathrm{\leavevmode\nobreak\ for\leavevmode\nobreak\ }\mathbf{r}\in\mathcal{C}\\ \psi=0&\mathrm{\leavevmode\nobreak\ for\leavevmode\nobreak\ }\mathbf{r}\in\partial\mathcal{C},\end{array}\right. (52)

whose solution is found analytically as

ε2ψ(2)(r,z,t)=ε2𝒞×𝒩[ψ(1),b(1)]G(𝐫,t;𝐫,t)drdzdt.\varepsilon^{2}\psi^{(2)}(r,z,t)=\varepsilon^{2}\int_{\mathcal{C}\times\mathbb{R}}\mathcal{N}[\psi^{(1)},b^{(1)}]G(\mathbf{r},t;\mathbf{r^{\prime}},t^{\prime})\text{d}r^{\prime}\text{d}z^{\prime}\text{d}t^{\prime}. (53)

This expression predicts that the wave created through self-interaction is a first harmonic at 2ω2\omega, or a mean flow at zero frequency. The spatial dependence, however, being coupled between ll and mm, cannot be directly infered.

References

  • [1] A. Alastuey, M. Magro, and P. Pujol. Physique et outils mathématiques, Méthodes et exemples. EDP Sciences, 2008.
  • [2] L.E. Baker and B.R. Sutherland. The evolution of superharmonics excited by internal tides in non-uniform stratification. Journal of Fluid Mechanics, 2020.
  • [3] C.L. Beattie. Table of first 700 zeros of bessel functions – Jl(x)J_{l}(x) and Jl(x)J_{l}^{\prime}(x). The Bell system technical journal, pages 689 – 697, 1958.
  • [4] S. Boury, P. Odier, and T. Peacock. Axisymmetric internal wave transmission and resonance in non-linear stratifications. Journal of Fluid Mechanics, 886:A8, 2020.
  • [5] S. Boury, T. Peacock, and P. Odier. Excitation and resonant enhancement of axisymmetric internal wave modes. Physical Review Fluids, 4:034802, 2019.
  • [6] C. Brouzet, E.V. Ermanyuk, S. Joubaud, I.N. Sibgatullin, and T. Dauxois. Energy cascade in internal wave attractors. Europhysics Letters, 113:44001, 2016.
  • [7] E.V. Ermanyuk, J.-B. Flor, and B. Voisin. Jan-bert flór, bruno voisin. spatial structure of first and higher harmonic internal waves from a horizontally oscillating sphere. Journal of Fluid Mechanics, 671:364 – 383, 2011.
  • [8] A. Fincham and G. Delerce. Advanced optimization of correlation imaging velocimetry algorithms. Experiments in Fluids, 29:13 – 22, 2000.
  • [9] J.M.H. Fortuin. Theory and application of two supplementary methods of constructing density gradient columns. Journal of Polymer Science, 44:505 – 515, 1960.
  • [10] L. Gostiaux, H. Didelle, S. Mercier, and T. Dauxois. A novel internal waves generator. Experiments in Fluids, 42:123 – 130, 2006.
  • [11] P. Husseini, D. Varma, T. Dauxois, S. Joubaud, P. Odier, and Manikandan Mathur. Experimental study on superharmonic wave generation by resonant interaction between internal wave modes. Physical Review Fluids (submission), 2019.
  • [12] S. Joubaud, J. Munroe, P. Odier, and T. Dauxois. Experimental parametric subharmonic instability in stratified fluids. Physics of Fluids, 24, 2012.
  • [13] E. Kunze. Internal-wave-driven mixing: Geography and budgets. Journal of Physical Oceanography, 47:1325 – 1345, 2017.
  • [14] J.A. MacKinnon, Z. Zhao, C.B. Whalen, A.F. Waterhouse, O.M. Trossman, D.S. ans Sun, L.C. St.Laurent, H.L. Sommons, K. Polzin, R. Pinkel, A. Pickering, N.J. Norton, J.D. Nash, R. Musgrave, L.M. Merchant, A.V. Melet, B. Mater, S. Legg, W.G. Large, Z. Kunze, J.M. Klymak, M. Jochum, S.R. Jayne, R.W. Hallberg, S.M. Griffies, S. Diggs, G. Danabasoglu, E.P. Chassignet, M.C. Buijsman, F.O. Bryan, B.P. Briegleb, A. Barna, B.K. Arbic, J.K. Ansong, and M.H. Alford. Climate process team on internal wave-driven ocean mixing. Bulletin of the American Meteorological Society, 98:2429 – 2454, 2017.
  • [15] P. Maurer. Approche expérimentale de la dynamique non-linéaire d’ondes internes en rotation. PhD thesis, Université de Lyon, 2017.
  • [16] P. Maurer, S.J. Ghaemsaidi, S. Joubaud, T. Peacock, and P. Odier. An axisymmetric inertia-gravity wave generator. Experiments in Fluids, 58:143, 2017.
  • [17] D.E. Mowbray and B.S.H. Rarity. The internal wave pattern produced by a sphere moving vertically in a density stratified liquid. Journal of Fluid Mechanics, 30:489 – 495, 1967.
  • [18] F.W.J. Olver, D.W. Lozier, R.F. Boisvert, and C.W. Clark. NIST Handbook of Mathematical Functions. Cambridge University Press, 2010.
  • [19] G. Oster and M. Yamamoto. Density gradient techniques. Chemical Review, 63:257 – 268, 1963.
  • [20] N.D. Shmakova and J.-B. Flor. Nonlinear aspects of focusing internal waves. Journal of Fluid Mechanics, 2018.
  • [21] B.R. Sutherland. Excitation of superharmonics by internal modes in non-uniformly stratified fluid. Journal of Fluid Mechanics, 793:335 – 352, 016.
  • [22] S.A. Thorpe. On wave interactions in a stratified fluid. Journal of Fluid Mechanics, 24:737 – 751, 1966.
  • [23] D. Varma and M. Mathur. Internal wave resonant triads in finite-depth non-uniform stratifications. Journal of Fluid Mechanics, 824:286 – 311, 2017.
  • [24] B. Voisin. Limit states of internal waves beams. Journal of Fluid Mechanics, 496:243 – 293, 2003.
  • [25] B. Voisin, E.V. Ermanyuk, and J.-B. Flór. Internal wave generation by oscillation of a sphere, with application to internal tides. Journal of Fluid Mechanics, 666:308 – 357, 2011.
  • [26] G.L. Wagner and W.R. Young. A three-component model for the coupled evolution of near-inertial waves, quasi-geostrophic flow and the near-inertial second harmonic. Journal of Fluid Mechanics, 802:806 – 837, 2016.
  • [27] A.F. Waterhouse, J.A. MacKinnon, J.D. Nash, M.H. Alford, E. Kunze, H.L. Simmons, K.L. Polzin, L.C. St Laurent, O.M. Sun, R. Pinkel, L.D. Talley, C.B. Whalen, T.N. Huussen, G.S. Carter, I. Fer, S. Waterman, A.C. Naveira Garabato, T.B. Sanford, and C.M. Lee. Global patterns of diapycnal mixing from measurements of the turbulent dissipation rate. Journal of Physical Oceanography, 44:1854 – 1872, 2014.
  • [28] S. Wunsch, I. Delwiche, G. Frederick, and A. Brandt. Experimental study of nonlinear harmonic generation by internal waves incident on a pycnocline. Experiments in Fluids, 56:87, 2015.