Internal waves in marginally stable abyssal stratified flows...660 N. Makarenko et al.: Internal...

11
Nonlin. Processes Geophys., 25, 659–669, 2018 https://doi.org/10.5194/npg-25-659-2018 © Author(s) 2018. This work is distributed under the Creative Commons Attribution 4.0 License. Internal waves in marginally stable abyssal stratified flows Nikolay Makarenko 1,2 , Janna Maltseva 1,2 , Eugene Morozov 3 , Roman Tarakanov 3 , and Kseniya Ivanova 4 1 Lavrentyev Institute of Hydrodynamics, 6300090, Novosibirsk, Russia 2 Novosibirsk State University, Department of Mechanics and Mathematics, 630090, Novosibirsk, Russia 3 Shirshov Institute of Oceanology, 117997, Moscow, Russia 4 Aix-Marseille University, Department of Mechanics, Marseille 06, France Correspondence: Nikolay Makarenko ([email protected]) Received: 11 January 2018 – Discussion started: 26 January 2018 Revised: 13 June 2018 – Accepted: 10 July 2018 – Published: 5 September 2018 Abstract. The problem on internal waves in a weakly strat- ified two-layer fluid is studied semi-analytically. We dis- cuss the 2.5-layer fluid flows with exponential stratification of both layers. The long-wave model describing travelling waves is constructed by means of a scaling procedure with a small Boussinesq parameter. It is demonstrated that solitary- wave regimes can be affected by the Kelvin–Helmholtz in- stability arising due to interfacial velocity shear in upstream flow. 1 Introduction In this paper, we consider an analytical model of internal soli- tary waves in a two-layer fluid with the density continuously increasing with depth in both layers. This model is a develop- ment of non-linear two-layer models previously suggested by Ovsyannikov (1985), Miyata (1985) and Choi and Camassa (1999), as well as the latest 2.5-layer models considered by Voronovich (2003) and Makarenko and Maltseva (2008, 2009a, b). Two-layer approximation is a standard model of a sharp pycnocline in a stratified fluid with constant densities in each layer, but which is discontinuous at the interface. Cor- respondingly, the 2.5-layer model takes into account a slight density gradient in stratified layers which is comparable with the density jump at the interface. In all these cases, internal solitary waves can be described in closed form by the solu- tions resulting from the quadrature dη dx 2 = f (η) (1) for stationary wave elevation η(x). The simplest version of non-linearity f appears in a two-layer system; hence, it is the rational function f (η) = P (η)/Q(η) where P is a fourth- degree polynomial and Q depends linearly on η. Equation (1) also appears as a travelling wave equation for non-linear evo- lution systems similar to the single-layer dispersive Green– Naghdi model (see Choi and Camassa, 1999). These non- linear dispersive equations can be obtained by means of a long-wave perturbation technique as well as by Whitham’s variational method. Several authors noted that solitary-wave solutions of such approximate models are in good agreement with the numerical solutions of fully non-linear Euler equa- tions for a perfect two-layer fluid. In this context, Camassa and Tiron (2011) also compared numerical travelling wave solutions, supported by smooth stratification, with the known explicit solitary-wave solutions in order to optimize a two- layer model of the Euler system with smooth stratification. We apply the method of derivation involving asymptotic analysis of the non-linear Dubreil-Jacotin–Long equation that results from fully non-linear Euler equations of strat- ified fluid. The long-wave scaling procedure uses a small Boussinesq parameter which characterizes slightly increas- ing density in the layers and a small density jump at their interface. This method combines the approaches applied for- merly to a pure two-fluid system with the perturbation tech- nique discussed for the first time by Long (1965) and de- veloped by Benney and Ko (1978) for a continuous strat- ification. The parametric range of a solitary wave is con- sidered in the framework of the constructed mathematical model. It is demonstrated that these wave regimes can ap- proach the parametric domain of the Kelvin–Helmholtz in- stability. The stability of solitary travelling-wave solutions of Published by Copernicus Publications on behalf of the European Geosciences Union & the American Geophysical Union.

Transcript of Internal waves in marginally stable abyssal stratified flows...660 N. Makarenko et al.: Internal...

Page 1: Internal waves in marginally stable abyssal stratified flows...660 N. Makarenko et al.: Internal waves in marginally stable abyssal stratified flows 0 2 2 1 0 -12 2 1 Figure 1.

Nonlin. Processes Geophys., 25, 659–669, 2018https://doi.org/10.5194/npg-25-659-2018© Author(s) 2018. This work is distributed underthe Creative Commons Attribution 4.0 License.

Internal waves in marginally stable abyssal stratified flowsNikolay Makarenko1,2, Janna Maltseva1,2, Eugene Morozov3, Roman Tarakanov3, and Kseniya Ivanova4

1Lavrentyev Institute of Hydrodynamics, 6300090, Novosibirsk, Russia2Novosibirsk State University, Department of Mechanics and Mathematics, 630090, Novosibirsk, Russia3Shirshov Institute of Oceanology, 117997, Moscow, Russia4Aix-Marseille University, Department of Mechanics, Marseille 06, France

Correspondence: Nikolay Makarenko ([email protected])

Received: 11 January 2018 – Discussion started: 26 January 2018Revised: 13 June 2018 – Accepted: 10 July 2018 – Published: 5 September 2018

Abstract. The problem on internal waves in a weakly strat-ified two-layer fluid is studied semi-analytically. We dis-cuss the 2.5-layer fluid flows with exponential stratificationof both layers. The long-wave model describing travellingwaves is constructed by means of a scaling procedure with asmall Boussinesq parameter. It is demonstrated that solitary-wave regimes can be affected by the Kelvin–Helmholtz in-stability arising due to interfacial velocity shear in upstreamflow.

1 Introduction

In this paper, we consider an analytical model of internal soli-tary waves in a two-layer fluid with the density continuouslyincreasing with depth in both layers. This model is a develop-ment of non-linear two-layer models previously suggested byOvsyannikov (1985), Miyata (1985) and Choi and Camassa(1999), as well as the latest 2.5-layer models consideredby Voronovich (2003) and Makarenko and Maltseva (2008,2009a, b). Two-layer approximation is a standard model of asharp pycnocline in a stratified fluid with constant densitiesin each layer, but which is discontinuous at the interface. Cor-respondingly, the 2.5-layer model takes into account a slightdensity gradient in stratified layers which is comparable withthe density jump at the interface. In all these cases, internalsolitary waves can be described in closed form by the solu-tions resulting from the quadrature

(dηdx

)2

= f (η) (1)

for stationary wave elevation η(x). The simplest version ofnon-linearity f appears in a two-layer system; hence, it isthe rational function f (η)= P(η)/Q(η)where P is a fourth-degree polynomial andQ depends linearly on η. Equation (1)also appears as a travelling wave equation for non-linear evo-lution systems similar to the single-layer dispersive Green–Naghdi model (see Choi and Camassa, 1999). These non-linear dispersive equations can be obtained by means of along-wave perturbation technique as well as by Whitham’svariational method. Several authors noted that solitary-wavesolutions of such approximate models are in good agreementwith the numerical solutions of fully non-linear Euler equa-tions for a perfect two-layer fluid. In this context, Camassaand Tiron (2011) also compared numerical travelling wavesolutions, supported by smooth stratification, with the knownexplicit solitary-wave solutions in order to optimize a two-layer model of the Euler system with smooth stratification.

We apply the method of derivation involving asymptoticanalysis of the non-linear Dubreil-Jacotin–Long equationthat results from fully non-linear Euler equations of strat-ified fluid. The long-wave scaling procedure uses a smallBoussinesq parameter which characterizes slightly increas-ing density in the layers and a small density jump at theirinterface. This method combines the approaches applied for-merly to a pure two-fluid system with the perturbation tech-nique discussed for the first time by Long (1965) and de-veloped by Benney and Ko (1978) for a continuous strat-ification. The parametric range of a solitary wave is con-sidered in the framework of the constructed mathematicalmodel. It is demonstrated that these wave regimes can ap-proach the parametric domain of the Kelvin–Helmholtz in-stability. The stability of solitary travelling-wave solutions of

Published by Copernicus Publications on behalf of the European Geosciences Union & the American Geophysical Union.

Page 2: Internal waves in marginally stable abyssal stratified flows...660 N. Makarenko et al.: Internal waves in marginally stable abyssal stratified flows 0 2 2 1 0 -12 2 1 Figure 1.

660 N. Makarenko et al.: Internal waves in marginally stable abyssal stratified flows

������0

�2

�2

�1

��

�0

-�1

�2

�1

Figure 1. Scheme of the flow.

the Euler equations for continuously stratified, near two-layerfluids was studied numerically and analytically by Almgrenet al. (2012). They demonstrated that the wave-induced shearcan locally reach unstable configurations and give rise to lo-cal convective instability. This is in good qualitative agree-ment with the laboratory experiments performed by Grue etal. (2000). It seems that such a marginal stability of long in-ternal waves could explain the formation mechanism of verylong billow trains in abyssal flows observed by Van Haren etal. (2014).

2 Basic equations

We consider a 2-D motion of inviscid two-layer fluid whichis weakly stratified due to gravity in both layers. The fullynon-linear Euler equations describing the flow are

ρ(ut + uux + vuy)+px = 0, (2)ρ(vt + uvx + vvy)+py =−ρg, (3)ρt + uρx + vρy = 0, (4)ux + vy = 0, (5)

where ρ is the fluid density, (u,v) is the fluid velocity, p isthe pressure and g is the gravity acceleration. We assume thatthe flow domain is bounded by the flat bottom y =−h1 andthe rigid lid y = h2 (see Fig. 1), with the boundary condition

v = 0∣∣y=−h1, y=h2

. (6)

The layers are separated by the interface y = η(x, t) withthe equilibrium level at y = 0. Non-linear kinematic and dy-namic boundary conditions at this interface are

ηt + uηx = v∣∣y=η

, [p ] = 0∣∣y=η

, (7)

where the square brackets denote the discontinuity jump atthe interface between the layers. Non-disturbed parallel flowhas no vertical velocity and elevation (i.e. v = 0, η = 0), butthe horizontal velocity u= u0(y) may be piece-wise con-stant:

u0(y)=

{u1 (−h1 < y < 0),u2 (0< y < h2).

In this stationary case, the fluid density ρ = ρ0(y) and pres-sure p = p0(y) should be coupled by the hydrostatic equa-tion dp0/dy = gρ0. We consider the density profile depend-ing exponentially on height,

ρ0(y)=

{ρ1 exp (−N2

1y/g) (−h1 < y < 0),

ρ2 exp (−N22y/g) (0< y < h2),

(8)

where Nj = const is the Brunt–Väisälä frequency in the j thlayer, and constant densities ρ1 and ρ2 are related as ρ2 < ρ1.The special case Nj = 0 (j = 1,2) gives a familiar two-fluidsystem with piece-wise constant density ρ = ρj in the j thlayer.

Further we consider a steady non-uniform flow; hence,we have ηt = 0 and ut = vt = ρt = 0 in Eqs. (2)–(4). Weintroduce the stream function ψ by standard formulae u=ψy, v =−ψx ; hence, the mass conservation implies the de-pendence ρ = ρ(ψ), and pressure p can be found from theBernoulli equation

12|∇ψ |2+

1ρ(ψ)

p+ gy = b(ψ). (9)

Seeking solitary-wave solutions, we require that the up-stream velocity of the fluid (u,v) tend to (uj ,0) as x→−∞.In this case, boundary conditions (6) transform to the condi-tions for the stream function as

ψ =−u1h1∣∣y=−h1

, ψ = 0∣∣y=η

, ψ = u2h2∣∣y=h2

. (10)

It is known (Yih, 1980) that systems (2)–(5) can be reducedin a stationary case to the non-linear Dubreil-Jacotin–Long(DJL) equation for the stream function

ρ(ψ)∇2ψ + ρ′(ψ)

(gy+

12|∇ψ |2

)=H ′(ψ). (11)

Here, the functionH(ψ)= ρ(ψ)b(ψ) involves the Bernoullifunction b(ψ) and the density function ρ(ψ), so that H isspecified by the upstream condition. More exactly, the den-sity function is determined by the relation ρ(ψ)= ρ0(ψ/uj )

in the j th layer, and the Bernoulli function b(ψ) is definedby the formula

b =

12u2

1+ gψ

u1+g2

N21

(1− e

N21ψgu1

)(−h1 < y < η(x)),

12u2

2+ gψ

u2+g2

N22

(1− e

N22ψgu2

)(η(x) < y < h2).

As a consequence, we can rewrite the DJL equation (11) asfollows:

∇2ψ =

N2j

guj

{g

(y−

ψ

uj

)+

12

(|∇ψ |2− u2

j

)}, (12)

where j = 1 is related to the lower layer, and j = 2 to theupper layer. Further, in accordance with relations (7) and (9),

Nonlin. Processes Geophys., 25, 659–669, 2018 www.nonlin-processes-geophys.net/25/659/2018/

Page 3: Internal waves in marginally stable abyssal stratified flows...660 N. Makarenko et al.: Internal waves in marginally stable abyssal stratified flows 0 2 2 1 0 -12 2 1 Figure 1.

N. Makarenko et al.: Internal waves in marginally stable abyssal stratified flows 661

the continuity of pressure p provides a non-linear boundarycondition for stream function ψ :

[ρ(ψ)(|∇ψ |2+ 2gy− 2b(ψ)] = 0∣∣y=η

. (13)

Using the explicit form of functions ρ(ψ) and b(ψ), condi-tion (13) can also be rewritten in detail as follows:

2g(ρ1− ρ2)η =

= ρ2(|∇ψ |2− u2

2)∣∣y=η(x)+0− ρ1(|∇ψ |

2− u2

1)∣∣y=η(x)−0.

We reformulate this boundary condition in view of conserva-tion of the total horizontal momentum in a steady two-layerflow, which has the integral formulation

h2∫−h1

(p+ ρ u2) dy = C,

where constant C is determined by the upstream condition.Excluding pressure p from this relation using the Bernoulliequation (9) leads to the integral relation

ρ1

η(x)∫−h1

e−N2

1ψgu1 91 dy+ ρ2

h2∫η(x)

e−N2

2ψgu2 92 dy = C, (14)

where the integrand functions 9j are

9j = ψ2y −ψ

2x + u

2j + 2g

uj− y

)−

2g2

N2j

e N2jψ

guj − 1

,and constant C depends on the parameters of the upstreamflow as follows:

C = 2ρ1g

[(eN2

1 h1g − 1

)(u2

1

N21+g2

N41

)−gh1

N21

]+

+ 2ρ2g

[(1− e−

N22 h2g

)(u2

2

N22+g2

N42

)−gh2

N22

].

It is important here that the integral relation (14) is equiva-lent to the boundary condition (13), which is rather simple.This equivalence can be checked immediately by differentia-tion of the relation (14) with respect to the variable x, so theintegrals can be evaluated explicitly due to Eq. (12). Equa-tion (14) will be used later instead of Eq. (13) by the con-struction model differential equation for the function η(x)describing strongly non-linear waves.

3 Non-dimensional formulation

Now we introduce scaled independent variables x and y andscaled unknown functions η and ψ in order to reformulatethe basic equations in the dimensionless form. That is, the

fixed ratio h1/π is used as an appropriate length scale for x,y, and η, and normalized volume discharges ujhj/π serveas the units for the stream function; thus, we have

(x,y,η)=h1

π(x,y,η), ψ =

ujhj

πψ

separately in the lower layer (j = 1) or in the upper layer(j = 2). The number π is only introduced here due to the spe-cific form of trigonometric modal functions which are typicalfor the exponential density (Eq. 8). The scaling procedurewith this density profile uses the Boussinesq parameters σ1and σ2 and the Atwood number µ defined by the formulae

σj =N2j hj

πg(j = 1,2), µ=

ρ1− ρ2

ρ2. (15)

Here, constants σj characterize the slope of the density pro-file in continuously stratified layers, and parameter µ deter-mines the density jump at the interface.

Following Turner (1973), we introduce the densimetric (orinternal) Froude number

Fj =uj√gjhj

(j = 1,2),

which presents scaled fluid velocity uj in the j th layer, de-fined with reduced gravity acceleration gj = (ρ1− ρ2)g/ρj .In addition to the Froude numbers Fj , it is also convenient touse the pair of Long’s numbers λj given by the formula

λj =Njhj

πuj(j = 1,2).

Long’s numbers λj are coupled with the Boussinesq param-eters σ1 and σ2, the Atwood number µ and the Froude num-bers Fj by the relations

λ21 =

πσ1(1+µ)µF 2

1, λ2

2 =πσ2

µF 22. (16)

Finally, we introduce the ratio of undisturbed thicknesses ofthe layers r = h1/h2. By that notation, we locate the bottomas y =−π , and relation y = π/r defines the rigid lid. Thus,we obtain the equations for scaled stream function ψ andnon-dimensional wave elevation η as follows (bar is omittedthroughout what follows):

∇2ψ + λ2

1 (ψ − y)=12σ1

(|∇ψ |2− 1

)(17)

in the lower layer −π < y < η(x), and

∇2ψ + λ2

2 r2(ψ − ry)=

12σ2

(|∇ψ |2− r2

)(18)

in the upper layer η(x) < y < π/r . Kinematic boundary con-ditions (10) can be rewritten now as follows:

ψ(x,−π)=−π, ψ(x,η(x))= 0, ψ(x,π/r)= π. (19)

www.nonlin-processes-geophys.net/25/659/2018/ Nonlin. Processes Geophys., 25, 659–669, 2018

Page 4: Internal waves in marginally stable abyssal stratified flows...660 N. Makarenko et al.: Internal waves in marginally stable abyssal stratified flows 0 2 2 1 0 -12 2 1 Figure 1.

662 N. Makarenko et al.: Internal waves in marginally stable abyssal stratified flows

Correspondingly, Eq. (13), providing continuity of pressureat interface y = η(x), leads to the non-linear boundary con-dition

2η = F 22 (|∇ψ |

2− r2)

∣∣y=η+0−F

21 (|∇ψ |

2− 1)

∣∣y=η−0, (20)

and the dimensionless version of integral relation (14) takesthe form

η∫−π

e−σ1ψ 91 dy+

π/r∫η

e−σ2ψ 92 dy = C, (21)

where the following is denoted:

91 =µF 2

12

(ψ2y −ψ

2x + 1

)+

1+µπ

(ψ − y−

eσ1ψ − 1σ1

),

92 =µF 2

22r3

(ψ2y −ψ

2x + r

2)+

1πr

(ψ − ry−

eσ2ψ − 1σ2

).

Constant

C = πµ

(F 2

1 +F 2

2r2

)+

+ (1+µ)eσ1π − 1− σ1ππ(λ2

1+ σ21 )+

1− σ2π − e−σ2π

πr2(λ22+ σ

22 )

is chosen here so that the horizontal upstream flow given bythe solution

η = 0, ψ0(y)=

{y (−π < y < 0),ry (0< y < π/r),

(22)

satisfies momentum relation (21).The model of fully non-linear travelling waves in a two-

layer irrotational flow, with the interface y = η(x) betweenthe fluids with constant densities ρ2 in the upper layer andρ1 > ρ2 in the lower layer, can be specified as follows. Inthis limit case, at least formally, the Boussinesq parametersσj and Long’s numbers λj vanish: σ1 = σ2 = λ1 = λ2 = 0.Therefore, we obtain the Laplace equation

∇2ψ = 0 (23)

instead of Eqs. (17)–(18), but all the boundary conditions(19) and (20) still remain unchanged.

4 Spectrum of harmonic waves

In many cases, a parametric range of solitary waves can bedetermined a priori as the domain being supercritical withrespect to the spectrum of small-amplitude sinusoidal waves.It is helpful while the critical phase speed can be simply de-fined from the dispersion relation of infinitesimal waves. In

our case, linearizing of Eqs. (17)–(20) for the upstream solu-tion (22) leads to the dispersion relation

1(k;F1,F2)= 0 (24)

for stationary harmonic wave packets

η(x)= a eikx, ψ = ψ0(y)+W(y)eikx .

Here, k is the non-dimensional wave number, a is the ampli-tude of the interfacial wave, and W(y) is the modal eigen-function which describes deformation of streamlines withinthe fluid layers. For the given Long numbers λ1 and λ2 andthe Boussinesq parameters σ1 and σ2, we also introduce non-dimensional values

~j =

√|λ2j − k

2j −

14π2σ 2

j | (j = 1,2), (25)

where k1 = rk and k2 = k are dimensionless wave numbersspecified for each layer. According to these notations, disper-sion function 1(k;F1,F2) in (24) has the form

1= F 21

(~1Cot1~1+

πσ1

2

)+F 2

2

(~2Cot2~2−

πσ2

2

)− 1,

where functions Cotj (j = 1,2) are denoted as follows:

Cotj ~j =

Cot~j

(λ2j > k

2j +

14π2σ 2

j

)coth~j

(λ2j < k

2j +

14π2σ 2

j

).

In fact, function 1 takes such a combined form since modalfunction W(y) depends on y trigonometrically or hyperboli-cally, if the radicand term λ2

j−k2j−

14π

2σ 2j in (25) is positive

or negative. Explicit formulae for these modal eigenfunctionsW(y) are given in Appendix A.

A spectrum of stationary harmonic waves, defined on the(F1,F2) plane, is formed by the Froude points (F1,F2) sothat dispersion function 1(k;F1,F2), which is even in k,has at least one pair of real roots ±k. Wave modes differ bythe number of these pairs, and this number can change onlyby passing of the root across the value k = 0. Therefore, themodal bounds should satisfy the equation 1(0;F1,F2)= 0;these bounds are defined by separate branches of the curve

F 21

{√λ2

1−(πσ1

2

)2Cot

√λ2

1−(πσ1

2

)2−πσ1

2

}+ (26)

+F 22

{√λ2

2−(πσ2

2

)2Cot

√λ2

2−(πσ2

2

)2+πσ2

2

}= 1,

where parameters λj should be coupled with the Froudenumbers Fj using the formula (16).

We emphasize that parameters σj characterize the slopeof the density profile in continuously stratified layers, and µdefines the density jump at the interface. As usual, all these

Nonlin. Processes Geophys., 25, 659–669, 2018 www.nonlin-processes-geophys.net/25/659/2018/

Page 5: Internal waves in marginally stable abyssal stratified flows...660 N. Makarenko et al.: Internal waves in marginally stable abyssal stratified flows 0 2 2 1 0 -12 2 1 Figure 1.

N. Makarenko et al.: Internal waves in marginally stable abyssal stratified flows 663

0

1.0

2.0

F2

1.0 2.0

1F

-1.0-2.0

-2.0

-1.0

A

B

O

Kelvin - HelmholtzinstabilityKelvin - Helmholtzinstability

SolitarywavesSolitarywaves

Linear waves

Internal fronts

1F

F2

1.6 2.01.8

1.0

0.8

(a) (b)

Figure 2. Spectrum of linear waves (coloured modes 1–3) (a) and fragment of the parametric domain of solitary waves (b).

parameters are small in the case of low stratification. How-ever, the interfacial mode dominates over the modes of inter-nal waves in stratified layers when σj � µ is valid. In thislimit case, linearized boundary conditions (19)–(20), consid-ered together with the linear Laplace equation (23), lead tothe standard dispersion relation of two-layer fluid:

F 21 rk coth rk+F 2

2 k cothk = 1. (27)

This relation determines only a single pair of real wave num-bers of the interfacial mode, so the spectral domain of a per-fect two-layer system occupies the unit disk:

F 21 +F

22 61. (28)

The 2.5-layer model starts with the hypotheses that theBoussinesq parameters σ1 and σ2 and the Atwood number µare of the same order, so we can use a single small parameterσ by setting

σ = σ1 = σ2 = µ. (29)

The limit passage σ → 0 is singular because Long’s numbersλj involve the ratios σj/µ in formula (16). However, condi-tion (29) allows us to simplify the spectral portrait; hence,modal curve (26) defining the critical wave speeds takes theform

√πF1 Cot

√π

F1+√πF2 Cot

√π

F2= 1. (30)

Figure 2 demonstrates the parts of the spectrum defined bycurve (30) for the dominating modes. The domain coveredonly by the first mode is marked with the blue colour. Cor-respondingly, the embedded domain of the second mode ishighlighted with the yellow colour, and the third mode ismarked with the pink colour. It is important that this spec-trum differs essentially from the ordinary two-layer spectrum(28), even when the flow is characterized by a pair of the

Froude numbers F1 and F2 defined in the same manner. Weespecially note that the 2.5-layer spectrum extends infinitelyon the spectral plane by involving unbounded Froude num-bers Fj .

5 The non-linear long-wave model

The derivation procedure of the non-linear long-wave 2.5-layer model should involve, in accordance with hypothesis(29), the slow horizontal variable ξ =

√σ x, as was demon-

strated by Benney and Ko (1978) in the case of slight linearstratification. Scaling with the parameter σ gives the equa-tion

σψξξ +ψyy + λ21 (ψ − y)=

12σ(σψ2

ξ +ψ2y − 1

)(31)

in the lower layer −π < y < η(ξ), and

σψξξ +ψyy + λ22r

2 (ψ − ry)=12σ(σψ2

ξ +ψ2y − r

2)

(32)

in the upper layer η(ξ) < y < π/r . Kinematic boundary con-ditions (10) can be rewritten now as follows:

ψ(ξ,−π)=−π, ψ(ξ,η(ξ))= 0, ψ(ξ,π/r)= π. (33)

We find that stream function ψ is expanded in a power serieswith respect to σ as

ψ = ψ (0)(ξ,y)+ σ ψ (1)(ξ,y)+ . . ., (34)

where the leading-order term ψ (0) defines the hydrostaticmode, and the coefficient ψ (1) provides the correction dueto non-linear dispersion. All these coefficients ψ (k) can beuniquely determined from Eqs. (31) and (32) (with fixedLong numbers λ1 and λ2) under kinematic boundary con-dition (33). Thus, we obtain

ψ (0) = y− ηsinα1(y)

sinα1(η)(−π < y < η)

www.nonlin-processes-geophys.net/25/659/2018/ Nonlin. Processes Geophys., 25, 659–669, 2018

Page 6: Internal waves in marginally stable abyssal stratified flows...660 N. Makarenko et al.: Internal waves in marginally stable abyssal stratified flows 0 2 2 1 0 -12 2 1 Figure 1.

664 N. Makarenko et al.: Internal waves in marginally stable abyssal stratified flows

and

ψ (0) = ry− rηsinα2(y)

sinα2(η)(η < y < π/r),

where

α1(y)= λ1(π + y), α2(y)= λ2(π − ry). (35)

The final form of the dispersive termψ (1) is much more com-plicated; therefore, this coefficient is given in Appendix A.

Now we substitute power expansion (34) for function ψinto the scaled version of integral relation (21) and truncatethe terms with the powers higher than the first power of σ . Bythat, Eq. (21) reduces to the first-order ordinary differentialequation for the wave elevation η(x) and is written as(

dηdx

)2

= η2 D(η;F1,F2)

Q(η;F1,F2). (36)

Here function D is given by the formula

D(η;F1,F2)=

=√πF1 cotα1(η)+

√πF2 cotα2(η)+

13(1− r)η− 1,

where α1 and α2 should be taken as

α1(η)=π + η√πF1

, α2(η)=π − rη√πF2

, (37)

since we have at the leading order in σ the relations λj =1/√πFj (j = 1,2) obtained under condition (29). Denomi-

nator Q in Eq. (36) has a complicated form; therefore, thisfunction is given in Appendix C. Solitary-wave solutions ofEq. (36) are given in implicit form by the formula

x =±

η∫a

√Q(s;F1,F2)

D(s;F1,F2)

dss, (38)

where parameter a determines the non-dimensional ampli-tude of the wave.

Small-amplitude waves can be modelled by a simplifiedweakly non-linear version of Eq. (36) with the form(

dηdx

)2

= η2 D0+D1 η+D2 η2

Q(0;F1,F2), (39)

where the coefficients D0 =D(0;F1,F2) andD1 =D

′η(0;F1,F2) are

D0 =√πF1 cot

√π

F1+√πF2 cot

√π

F2− 1,

D1 =−cot2√π

F1+ rcot2

√π

F2+

23(r − 1),

and the explicit form of coefficient D2 is not important here.This model takes into account the balance of quadratic and

cubic non-linearities in the weakly non-linear KdV–mKdV–Gardner model (Kakutani and Yamasaki, 1978; Gear andGrimshaw, 1983; Helfrich and Melville, 2006; Grimshaw etal., 2002).

Solitary-wave regimes are obtained depending on the mul-tiplicity of the roots aj (F1,F2, r) (j = 1,2) of the numeratoron the right-hand side of Eq. (39). The profile of a solitarywave is given by the formulae

η(x)= a1− tanh2kx

1− θ2tanh2kx, k =

a√

3/q∗2θ

,

with q∗ =Q(0), a = a1 and θ2= a1/a2 < 1, and the bore

(internal front) corresponds to the double root a = a1 = a2;it has the following profile:

η(x)=a

2

(1+ tanhkx

), k =

a√

3/q∗2

.

The parametric range of strongly non-linear solitary wavesdescribed by Eq. (36) is formed by the domain in the (F1,F2)

plane where the radical function Q/D in (38) is ensured tobe non-negative. It is easy to check that Q(0;F1,F2) > 0;hence, function Q(s;F1,F2) is positive in the vicinity ofpoint s = 0. Therefore, function D plays the determiningrole here. Depending on F1 and F2, this function can changesign even by small s, where the leading-order coefficient D0from formula (39) dominates. As a consequence, the mapof solitary-wave regimes is formed by the Froude numbers(F1,F2) such that inequality D0(F1,F2) > 0 holds. Indeed,this inequality defines the range of non-linear waves, whichare supercritical with respect to the phase speed of linear har-monic wave packets (see Fig. 2).

6 Marginally stable layered flows

Large-amplitude internal waves are generated in deep oceanlayers due to the interaction of internal tides with irregularbottom topography near underwater ridges (Morozov, 1995,2018; Morozov et al., 2010). These waves play a significantrole in the energy transformation and mass transport in theoceanic stratified flows, while they intensify mixing of theabyssal waters. Note that internal Froude numbers F1 and F2characterize the magnitude of the velocity jump at the inter-face in upstream flow. The shear u1 6= u2 between the lay-ers can initiate the development of the Kelvin–Helmholtz in-stability which provides non-stationary formation of billowtrains (Thorpe, 1985; Drazin, 2002). In this context, long-wave perturbations give the greatest contribution to this in-stability due to their increased power intensity. Constant two-layer flow is linearly stable in the long-wave limit if the in-equality

|u1− u2|<

√g(ρ1− ρ2)(ρ1h2+ ρ2h1)

ρ1ρ2(40)

Nonlin. Processes Geophys., 25, 659–669, 2018 www.nonlin-processes-geophys.net/25/659/2018/

Page 7: Internal waves in marginally stable abyssal stratified flows...660 N. Makarenko et al.: Internal waves in marginally stable abyssal stratified flows 0 2 2 1 0 -12 2 1 Figure 1.

N. Makarenko et al.: Internal waves in marginally stable abyssal stratified flows 665

holds, and this flow is unstable in the opposite case (Lamb,1932). Exactly the same bound (40) follows from the non-linear stability criteria predicted by the shallow water theory(Ovsyannikov, 1979; see also Baines, 1995; Gavrilyuk et al.,2017) for a variable difference |u1− u2| and variable layerthicknesses

h1(x, t)= h1+ η(x, t), h2(x, t)= h2− η(x, t).

In this case, velocity uj is the horizontal fluid velocity aver-aged over the depth of the j th layer (j = 1,2),

u1(x, t)=1

h1(x, t)

η(x,t)∫−h1

u(x,y, t)dy,

u2(x, t)=1

h2(x, t)

h2∫η(x,t)

u(x,y, t)dy.

Condition (40) considered for non-constant velocities u1 andu2 and non-constant layer thicknesses h1 and h2 (h1+h2 =

h1+h2 = const) provides hyperbolicity of hydrostatic two-layer shallow water equations

h1t + (u1h1)x = 0, h2t + (u2h2)x = 0,

ρ1(u1t + u1u1x + gh1x)= ρ2(u2t + u2u2x + gh1x).

In accordance with this non-linear formulation, layered flowcan be locally unstable due to increased wave-induced sheareven when the constant upstream flow satisfies condition(40). This marginal instability of internal solitary waves wasobserved in laboratory experiments (Gavrilov, 1994; Grueet al., 1999, 2000; Carr et al., 2008; Fructus et al., 2009)and studied theoretically (Choi et al., 2009; Almgren etal., 2012). Mathematical models of layered stable–unstableflows, which account also for the non-linear dispersion ofthe waves and non-hydrostatic effects, are currently under in-tense discussion in Barros and Choi (2013), Lannes and Ming(2015), Duchene et al. (2016), Liapidevskii et al. (2018), andMakarenko et al. (2018). However, the hydrostatic condition(40) seems to be appropriate by estimation of the stabilitylimit with non-dimensional parameters from Sect. 3. Hence,we obtain the stability domain

|√rF1−F2|<

√1+ r,

which gives an admissible set of the Froude points (F1,F2)

defined for the instant ratio of local layer depths r = h1/h2and local Froude numbers in the solitary-wave-type flow de-scribed by stationary solution (38). This domain is shown forr = 3 in Fig. 2 (right panel) as the unshadowed domain in theselected part of quarter-plane (F1,F2).

7 Waves in abyssal shear flows

We present in this section a comparison of solutions of sug-gested mathematical models with the field data measured

4450

4500

Dep

th, m

09:35 10:35Time, h (4 Dec 2013)

Figure 3. Internal front in abyssal stratified flow.

4500

4550

4450

15:50 18:15Time, h (4 Dec 2013)

Dep

th, m

Figure 4. Train of interfacial solitary waves affected by the Kelvin–Helmholtz instability.

for internal solitary waves in weakly stratified abyssal cur-rents. Figures 3 and 4 demonstrate fragments of tempera-ture distribution in a quasi-steady shear bottom flow recordedfrom a mooring station with a 350 m line of thermistors lo-cated over a depth of 4720 m at the entrance to the Ro-manche Fracture Zone in the equatorial Atlantic (Van Harenet al., 2014). Trains of short-period (20/30 min) internalwaves modulated by tides propagate here along a sharp inter-face corresponding to the 0.85 ◦C isotherm which separatesthe lower layer of cold Antarctic Bottom Water (AABW)from the overlying warmer AABW layer with temperatureθ < 2.0 ◦C. It is seen from the figures that moored temper-ature data show permanently marginal stability of the flowwith the Richardson numbers 0.25<Ri < 1. Tidal amplifi-cation of the shear triggers the formation of small-scale over-turns which create long trains of the Kelvin–Helmholtz bil-lows. Bold curves in Figs. 3 and 4 show overlapped profilesof internal waves calculated from Eq. (38). Figure 3 demon-strates a front-like wave (internal bore) affected by moder-ate small-scale shear instability. The non-dimensional am-plitude of this wave is a = ηmax/h1 = 0.16, and the calcu-lated solution corresponds to Froude point A with coordi-nates (F1,F2)= (1.719,0.891) shown in Fig. 2 (panel (b)).Point A belongs to the bore diagram which is tangential tothe spectrum boundary at point O corresponding to the van-ishing amplitude limit a = 0. Figure 4 demonstrates the trainof three subsequent waves with intense overturns distributedalong with gently sloping wave tops. The amplitude of theleading wave shown in Fig. 4 is clearly lower than the am-plitudes of the following two waves, which occur randomlyand are almost identical. The solitary-wave type of thesetwo waves corresponds to Froude point B with coordinates(F1,F2)= (1.634,0.959).

The upstream parameters used in the calculation were cho-sen from conductivity, temperature and pressure (CTD) andlowered acoustic Doppler current profiler (LADCP) data of

www.nonlin-processes-geophys.net/25/659/2018/ Nonlin. Processes Geophys., 25, 659–669, 2018

Page 8: Internal waves in marginally stable abyssal stratified flows...660 N. Makarenko et al.: Internal waves in marginally stable abyssal stratified flows 0 2 2 1 0 -12 2 1 Figure 1.

666 N. Makarenko et al.: Internal waves in marginally stable abyssal stratified flows

4100

4200

4300

4400

4500

4600

1047 1048 1049

34.75 34.80 34.8534.70 34.90Salinity, PSU

Density, kg m-3

Temperature, Co

Depth

, m

1046

0.5 1.0 1.50.0 2.0

Figure 5. Profiles of the density, salinity and temperature.

density and currents measured immediately at the fronts ofselected waves. Undisturbed depth of lower-layer h1 wasfixed by the equilibrium level of isotherm 0.85 ◦C recorded atthe buoy station shortly before the arrival of the large wave.In fact, this level changed gradually due to the tide, so thatthe value of h1 varied within a day in the range from 170to 240 m. The depth of the upper layer h2 was evaluated byfixing the appropriate overlying isotherm, which remainedalmost horizontal; hence, this depth can be considered the“rigid lid” in the model of the 2.5-layer flow. This kind of“rigid” streamline appears frequently in abyssal shear flowsdue to the formation of critical layers which separate clearlythe currents with different thermohaline parameters (Moro-zov et al., 2012). The estimated depth h2 also depends on thetidal activity, and this value varied in a wider range from 70to 350 m. Nevertheless, the resulting total depth h= h1+h2did not exceed the thickness of the entire AABW layer con-fined below a depth of 4000 m.

Upstream velocities u1 and u2 were assumed as u1 =

35/45 cm s−1 and u2 = 10/20 cm s−1. These values corre-spond to the mean velocities measured in the western partof the Romanche Fracture Zone by a velocity LADCP pro-filer in the cruises of the R/V Akademik Sergey Vavilov

(Tarakanov et al., 2013) and measured on the mooring sta-tion (Van Haren et al., 2014). Maximal velocity of the flowin the lower layer, recorded by the mooring station within6 months, is as high as 65 cms−1. We consider that intenseshear is induced here by the permanent inflow of bottom wa-ters to the fracture zone through a narrow gap in its southernwall. Figure 5 demonstrates vertical profiles of temperature,salinity, and density measured on the shipborne mooring overa depth of 4760 m downstream from the location of the bot-tom station. These thermohaline data highlight the layeringof near-bottom flow with very smooth pycnoclines thickeneddue to intense mixing. In addition, the slope of an almostlinear density profile shown in Fig. 5 permits us to evaluatethe Boussinesq parameter σ by the value σ = 0.0027. TheFroude points (F1,F2) related to these parameters match ad-equately with the range of internal solitary waves shown inthe right panel of Fig. 2 (narrow band marked by the orangecolour). The narrowness of this range reveals anomalouslylow-amplitude dispersion of solitary waves, which agreeswith the field observations.

8 Conclusions

In this paper we have considered the problem of internal sta-tionary waves at the interface between exponentially strati-fied fluid layers. We demonstrated that the non-linear DJLmodel of weakly stratified 2.5-layer fluid flow can be reducedexplicitly to an approximate non-linear ordinary differentialequation describing large-amplitude internal solitary waves.The parametric range of solitary waves is described, includ-ing regimes of broad plateau-shaped solitary waves and in-ternal fronts. These wave regimes can be affected by theKelvin–Helmholtz instability induced by the velocity shear atthe interface; hence, the marginal stability of internal wavescould explain the formation mechanism of very long billowtrains observed in the Romanche Fracture Zone.

Data availability. The data can be requested from author Eu-gene Morozov by the e-mail [email protected].

Nonlin. Processes Geophys., 25, 659–669, 2018 www.nonlin-processes-geophys.net/25/659/2018/

Page 9: Internal waves in marginally stable abyssal stratified flows...660 N. Makarenko et al.: Internal waves in marginally stable abyssal stratified flows 0 2 2 1 0 -12 2 1 Figure 1.

N. Makarenko et al.: Internal waves in marginally stable abyssal stratified flows 667

Appendix A: Modal functions of the linearized problem

Eigenfunction W(y) considered in the strip −π < y < 0 hasthe form

W = a1e12 σ1y

e~1πsinh~1(π + y)

~1

(λ2

1 < k21 +

π2

4σ 2

1

)π + y

(λ2

1 = k21 +

π2

4σ 2

1

)sin~1(π + y)

~1

(λ2

1 > k21 +

π2

4σ 2

1

).

Similarly, eigenfunction W(y) defined in the upper layerwhich corresponds to the strip 0< y < π/r has the form

W = a2e12 σ2y

e~2πsinh~2(π − ry)

~2

(λ2

2 < k22 +

π2

4σ 2

2

)π − ry

(λ2

2 = k22 +

π2

4σ 2

2

)sins1(π − ry)

~2

(λ2

2 > k22 +

π2

4σ 2

2

).

The dimensionless wave numbers ~j (j = 1,2) are intro-duced in formula (25), and the factors aj are the amplitudeparameters.

Appendix B: Dispersive term of the long-waveexpansion

The coefficient ψ (1)(ξ,y) which gives the correction due tothe dispersion in the power expansion (34) has the form

ψ (1) =η(η−y)

2sinα1(y)

sinα1(η)+

sinα1(y)

2λ1

sinα1(η)

)ξξ

×

×

{(π + η)cotα1(η)− (π + y)cotα1(y)

}+η2

×

{sinλ1(y−η)−sinα1(y)

sin3α1(η)+

1+ sin2α1(y)

sin2α1(η)−

sinα1(y)

sinα1(η)

}

in the lower layer −π < y < η. Similarly, we have

ψ (1) =r2η(η−y)

2sinα2(y)

sinα2(η)+

sinα2(y)

2λ2

sinα2(η)

)ξξ

×

×

{(y−π/r)cotα2(y)− (η−π/r)cotα2(η)

}+r2η2

×

{sinλ2r(η−y)−sinα2(y)

sin3α2(η)+

1+ sin2α2(y)

sin2α2(η)−

sinα2(y)

sinα2(η)

}

in the upper layer η < y < π/r . Here the functions αj aregiven by the formula (35).

Appendix C: Denominator of the non-linear long-waveequation

Denominator Q in Eq. (36) has the form

2Q(η;F1,F2)=(πF 2

1 − 2√πF1ηcotα1(η)+ η

2cot2α1(η)

×

(η+π

sin2α1(η)−√πF1 cotα1(η)

)+

+

(πF 2

2r2 − 2

√πF2

rηcotα2(η)+ η

2cot2α2(η)

×

(π − rη

sin2α2(η)−√πF2 cotα2(η)

),

where functions αj (η) are given by formula (37), which is anapproximate version of Eq. (35).

www.nonlin-processes-geophys.net/25/659/2018/ Nonlin. Processes Geophys., 25, 659–669, 2018

Page 10: Internal waves in marginally stable abyssal stratified flows...660 N. Makarenko et al.: Internal waves in marginally stable abyssal stratified flows 0 2 2 1 0 -12 2 1 Figure 1.

668 N. Makarenko et al.: Internal waves in marginally stable abyssal stratified flows

Author contributions. All authors made the same contribution tothis work.

Competing interests. The authors declare that they have no conflictof interest.

Special issue statement. This article is part of the special issue “Ex-treme internal wave events”. It is a result of the EGU, Vienna, Aus-tria, 23–28 April 2017.

Acknowledgements. This work was supported by the RussianFoundation for Basic Research (grant nos. 15-01-03942, 17-08-00085 and 18-01-00648) and Interdisciplinary Program II.1 ofSB RAS (project no. 2). Roman Tarakanov was supported by theRussian Science Foundation (grant no. 16-17-10149).

Edited by: Kateryna TerletskaReviewed by: two anonymous referees

References

Almgren, A., Camassa, R., and Tiron, R.: Shear instability of inter-nal solitary waves in Euler fluids with thin pycnoclines, J. FluidMech., 710, 324–361, 2012.

Baines, P. G.: Topographic effects in stratified flows, CambridgeUniversity Press, Cambridge, UK, 1995.

Barros, R. and Choi, W.: On regularizing the strongly nonlinearmodel for two-dimensional internal waves, Physica D, 264, 27–34, 2013.

Benney, D. J. and Ko, D. R. S.: The propagation of long large am-plitude internal waves, Stud. Appl. Math., 59, 187–199, 1978.

Camassa, R. and Tiron, R.: Optimal two-layer approximation forcontinuous density stratification, J. Fluid Mech., 669, 32–54,2011.

Carr, M., Fructus, D., Grue, J., Jensen, A., and Davies, P.A.: Convectively induced shear instability in large am-plitude internal solitary waves, Phys. Fluids, 20, 12660,https://doi.org/10.1063/1.3030947, 2008.

Choi, W. and Camassa, R.: Fully nonlinear internal waves in a two-fluid system, J. Fluid Mech., 396, 1–36, 1999.

Choi, W., Barros, R., and Jo, T.-C.: A regularized model for stronglynonlinear internal solitary waves, J. Fluid Mech., 629, 73–85,2009.

Drazin, P.: Introduction to hydrodynamic stability, Cambridge Univ.Press, Cambridge UK, 2002.

Duchene, V., Israwi, S., and Talhouk, R.: A new class of two-layerGreen – Naghdi systems with improved frequency dispersion,Stud. Appl. Math., 137, 356–415, 2016.

Fructus, D., Carr, M., Grue, J., Jensen, A., and Davies, P.: Shear-induced breaking of large internal solitary waves, J. Fluid Mech.,620, 1–29, 2009.

Gavrilov, N. V.: Internal solitary waves and smooth bores whichare stationary in a laboratory coordinate system, J. Appl. Mech.Techn. Phys., 35, 29–33, 1994.

Gavrilyuk, S. L., Makarenko, N. I., and Sukhinin, S. V.: Waves incontinuous media, Lecture notes in geosystem mathematics andcomputing, Birkhäuser/Springer, Cham, Switzerland, 2017.

Gear, J. A. and Grimshaw, R.: A second-order theory for solitarywaves in shallow fluids, Phys. Fluids, 26, 14–29, 1983.

Grimshaw, R., Pelinovsky, E., and Poloukhina, O.: Higher-orderKorteweg-de Vries models for internal solitary waves in a strati-fied shear flow with a free surface, Nonlin. Processes Geophys.,9, 221–235, https://doi.org/10.5194/npg-9-221-2002, 2002.

Grue, J., Jensen, A., Rusås, P.-O., and Sveen, J. K.: Properties oflarge-amplitude internal waves, J. Fluid Mech., 380, 257–278,1999.

Grue, J., Jensen, A., Rusås, P.-O., and Sveen, J. K.: Breaking andbroadening of internal solitary waves, J. Fluid Mech., 413, 181–217, 2000.

Helfrich, K. R. and Melville, W. K.: Long nonlinear internal waves,Annu. Rev. Fluid Mech., 38, 395–425, 2006.

Kakutani, T. and Yamasaki, N.: Solitary waves on a two-layer fluid,J. Phys. Soc. Jpn., 45, 674–679, 1978.

Lamb, H.: Hydrodynamics, 6th edn., Cambridge Univ. Press, Cam-bridge, UK, 1932.

Lannes, D. and Ming, M.: The Kelvin – Helmholtz instabilitiesin two-fluids shallow water models, Comm. Fields Inst., hal-01101993, https://hal.archives-ouvertes.fr/hal-01101993 (lastaccess: 11 January 2018), 2015.

Liapidevskii, V. Yu., Dutykh, D., and Gisclon, M.: On the modellingof shallow turbidity flows, Adv. Water Resources, 113, 310–327,2018.

Long, R. R.: On the Boussinesq approximation and its role in thetheory of internal waves, Tellus, 17, 46–52, 1965.

Makarenko, N., Maltseva, J., Tarakanov, R., and Ivanova, K.: Inter-nal solitary waves in a layered weakly stratified flow, in: TheOcean in Motion, edited by: Velarde, M., Tarakanov, R., andMarchenko, A., Springer, Dordrecht, the Netherlands, 55–66,2018.

Makarenko, N. I. and Maltseva, J. L.: An analytical model of largeamplitude internal solitary waves, in: Extreme Ocean Waves,edited by: Pelinovsky, E. and Kharif, C., Springer, Dordrecht,the Netherlands, 179–189, 2008.

Makarenko, N. I. and Maltseva, J. L.: Phase velocity spectrum ofinternal waves in a weakly-stratified two-layer fluid, Fluid Dyn.,44, 278–294, 2009a.

Makarenko, N. I. and Maltseva, J. L.: Solitary waves in a weaklystratified two-layer fluid, J. Appl. Mech. Techn. Phys., 50, 229–234, 2009b.

Miyata, M.: An internal solitary wave of large amplitude, La Mer,23, 43–48, 1985.

Morozov, E., Demidov, A., Tarakanov, R., and Zenk, W.:Abyssal Channels in the Atlantic Ocean: Water Structure andFlows, Springer, Dordrecht, the Netherlands, 2010, 266 pp.,https://doi.org/10.1007/978-90-481-9358-5, 2010.

Morozov, E. G.: Semidiurnal internal wave global field, Deep-SeaRes., 42, 135–148, 1995.

Morozov, E. G.: Oceanic internal tides, Observations, Analysis, andModelling, Springer, Dordrecht, the Netherlands, 2018.

Morozov, E. G., Tarakanov, R. Yu., Lyapidevskii, V. Yu., andMakarenko, N. I.: Abyssal cataracts in the Romanche and Chainfracture zones, Dokl. Earth Sci., 446, 1211–1214, 2012.

Nonlin. Processes Geophys., 25, 659–669, 2018 www.nonlin-processes-geophys.net/25/659/2018/

Page 11: Internal waves in marginally stable abyssal stratified flows...660 N. Makarenko et al.: Internal waves in marginally stable abyssal stratified flows 0 2 2 1 0 -12 2 1 Figure 1.

N. Makarenko et al.: Internal waves in marginally stable abyssal stratified flows 669

Ovsyannikov, L. V.: Two-layer shallow water model, J. Appl. Mech.Tech. Phys., 20, 127–135, 1979.

Ovsyannikov, L. V., Makarenko, N. I., Nalimov, V. I., Liapide-vskii, V. V., Plotnikov, P. I., Sturova, I. V., Bukreev, V. I., andVladimirov, V. A.: Nonlinear Problems of the Theory of Surfaceand Internal Waves, Nauka, Novosibirsk, Russia, 1985 (in Rus-sian).

Tarakanov, R. Yu., Makarenko, N. I., and Morozov, E. G.: Antarcticbottom water flow in the western part of the Romanche FractureZone based on the measurements in October of 2011, Oceanol-ogy, 53, 655–667, https://doi.org/10.1134/S0001437013050147,2013.

Thorpe, S. A.: Laboratory observations of secondary structures inKelvin – Helmholtz billows and consequences for ocean mixing,Geophys. Astrophys. Fluid Dyn., 34, 175–190, 1985.

Turner, J. S.: Buoyancy effects in fluid, Cambridge Univ. Press,Cambridge, UK, 1973.

Van Haren, H., Gostiaux, L., Morozov, E., and Tarakanov, R.:Extremely long Kelvin – Helmholtz billow trains in the Ro-manche Fracture Zone, Geophys. Res. Lett., 44, 8445–8451,https://doi.org/10.1002/2014GL062421, 2014.

Voronovich, A. G.: Strong solitary internal waves in a 2.5-layermodel, J. Fluid Mech., 474, 85–94, 2003.

Yih, C. S.: Stratified flows, Academic Press, New York, USA, 1980.

www.nonlin-processes-geophys.net/25/659/2018/ Nonlin. Processes Geophys., 25, 659–669, 2018