5
RAPID COMMUNICATIONS PHYSICAL REVIEW E 88, 021201(R) (2013) Modified reduced Ostrovsky equation: Integrability and breaking E. R. Johnson Department of Mathematics, University College London, UK R. H. J. Grimshaw Department of Mathematical Sciences, Loughborough University, UK (Received 2 May 2013; published 27 August 2013) The modified reduced Ostrovsky equation is a reduction of the modified Korteweg–de Vries equation, in which the usual linear dispersive term with a third-order derivative is replaced by a linear nonlocal integral term, representing the effect of background rotation. Here we study the case when the cubic nonlinear term has the same polarity as the rotation term. This equation is integrable provided certain slope constraints are satisfied. We demonstrate, through theoretical analysis and numerical simulations, that when this constraint is not satisfied at the initial time, wave breaking inevitably occurs. DOI: 10.1103/PhysRevE.88.021201 PACS number(s): 05.45.Yv, 92.10.Ei, 92.10.Hm, 92.10.Iv I. INTRODUCTION It is well known that the extended Korteweg–de Vries, or Gardner, equation can be used to model internal solitary waves in the atmosphere and ocean, see for instance the reviews by Grimshaw [1,2] and Helfrich and Melville [3], u t + μuu x + νu 2 u x + λu xxx = 0. (1) Here u(x,t ) is the amplitude of an appropriate linear long wave mode, with linear long wave speed c 0 , and (1) is expressed in a frame moving with that speed. The coefficients μ,ν,λ are found from certain internal expressions involving the modal function and the background density stratification. When the coefficient μ = 0, (1) becomes the modified KdV equation. This case can be realized in practice, for instance, when the background stratification is that for a two-layer fluid, with equal layer depths. However, when the effects of background rotation through the Coriolis parameter f need to be taken into account, an extra term is needed, and (1) is replaced by (u t + μuu x + νu 2 u x + λu xxx ) x = γ u, (2) where γ = f 2 /2c 0 = 0. When (1) is just the KdV equation, that is the coefficient ν = 0, then (2) becomes the Ostrovsky equation; see Ostrovsky [4], Grimshaw [5], or the review by Grimshaw et al. [6]. Our concern here is with the modified reduced Ostrovsky equation which is obtained by setting λ = 0 and μ = 0 in (2), (u t + νu 2 u x ) x = γ u. (3) Importantly, we note that when γ = 0, Eq. (3) reduces to a modified Hopf equation. It is then easily demonstrated that all localized solutions, or all periodic solutions, will break. That is, the solution will develop an infinite slope in finite time. The issue then is how this breaking is affected when γ = 0. This is to be contrasted with the regularization by the mKdV equation when breaking is replaced by the emergence of modulated periodic waves; see Kamchatnov et al. [7]. This present study is motivated by the recent article by Grimshaw et al. [8] who examined the reduced Ostrovsky equation, that is (2) with λ,ν = 0. They showed that then the equation is either integrable and solutions exist for all time, or wave breaking occurs, depending on a certain criterion on the curvature of the initial condition. Since (3) can be regarded as the reduced Ostrovsky equation with cubic nonlinearity rather than quadratic nonlinearity, we expect that a similar type of analysis can be applied here, and that is indeed the case, with the curvature criterion there being replaced here with slope criteria. A canonical form of the modified reduced Ostrovsky equation is obtained by setting u(γ/2|ν |) 1/2 ˜ u,t = ˜ t/γ,x = ˜ x , so that u t ± u 2 u x 2 x = u. (4) Here the tilde (˜) has been removed. There are two equations, called respectively MROI, MROII according to the sign ± in the coefficient in the cubic term, corresponding to the sign of γν in (2). MROII is also known as the short pulse equation arising in nonlinear optics; see Sakovich and Sakovich [9,10], Liu et al. [11], Pelinovsky and Sakovich [12] and the references therein. Note that Eq. (4) has the symmetry that if u is a solution, so also is u. Also, the equation is invariant under the transformation u(x,t ) = D ˜ ux, ˜ t ) where ˜ x = x/D, ˜ t = Dt for any D> 0, and in particular, the slope u x = ˜ u ˜ x is invariant. The equation has a zero mass constraint for periodic or localized solutions, D udx = 0, (5) where D is the periodic, or infinite, domain. There are also conservation laws for “momentum” and “energy,” similar to those for the Ostrovsky equation; see Grimshaw and Helfrich [13]. Further, multiplying (4) by u x yields ∂t ± u 2 2 ∂x u 2 x + 2uu x ( ± u 2 x 1 ) = 0, and hence there is another conservation law, F t ± u 2 F 2 x = 0, F = 1 u 2 x 1/2 . (6) This conservation law is crucial for the issue of integrability and breaking. 021201-1 1539-3755/2013/88(2)/021201(5) ©2013 American Physical Society

Modified reduced Ostrovsky equation: Integrability and breaking

  • Upload
    r-h-j

  • View
    214

  • Download
    0

Embed Size (px)

Citation preview

RAPID COMMUNICATIONS

PHYSICAL REVIEW E 88, 021201(R) (2013)

Modified reduced Ostrovsky equation: Integrability and breaking

E. R. JohnsonDepartment of Mathematics, University College London, UK

R. H. J. GrimshawDepartment of Mathematical Sciences, Loughborough University, UK

(Received 2 May 2013; published 27 August 2013)

The modified reduced Ostrovsky equation is a reduction of the modified Korteweg–de Vries equation, inwhich the usual linear dispersive term with a third-order derivative is replaced by a linear nonlocal integral term,representing the effect of background rotation. Here we study the case when the cubic nonlinear term has thesame polarity as the rotation term. This equation is integrable provided certain slope constraints are satisfied. Wedemonstrate, through theoretical analysis and numerical simulations, that when this constraint is not satisfied atthe initial time, wave breaking inevitably occurs.

DOI: 10.1103/PhysRevE.88.021201 PACS number(s): 05.45.Yv, 92.10.Ei, 92.10.Hm, 92.10.Iv

I. INTRODUCTION

It is well known that the extended Korteweg–de Vries, orGardner, equation can be used to model internal solitary wavesin the atmosphere and ocean, see for instance the reviews byGrimshaw [1,2] and Helfrich and Melville [3],

ut + μuux + νu2ux + λuxxx = 0. (1)

Here u(x,t) is the amplitude of an appropriate linear long wavemode, with linear long wave speed c0, and (1) is expressed ina frame moving with that speed. The coefficients μ,ν,λ arefound from certain internal expressions involving the modalfunction and the background density stratification. When thecoefficient μ = 0, (1) becomes the modified KdV equation.This case can be realized in practice, for instance, when thebackground stratification is that for a two-layer fluid, withequal layer depths. However, when the effects of backgroundrotation through the Coriolis parameter f need to be taken intoaccount, an extra term is needed, and (1) is replaced by

(ut + μuux + νu2ux + λuxxx)x = γ u, (2)

where γ = f 2/2c0 �= 0. When (1) is just the KdV equation,that is the coefficient ν = 0, then (2) becomes the Ostrovskyequation; see Ostrovsky [4], Grimshaw [5], or the review byGrimshaw et al. [6].

Our concern here is with the modified reduced Ostrovskyequation which is obtained by setting λ = 0 and μ = 0 in (2),

(ut + νu2ux)x = γ u. (3)

Importantly, we note that when γ = 0, Eq. (3) reduces to amodified Hopf equation. It is then easily demonstrated thatall localized solutions, or all periodic solutions, will break.That is, the solution will develop an infinite slope in finitetime. The issue then is how this breaking is affected whenγ �= 0. This is to be contrasted with the regularization by themKdV equation when breaking is replaced by the emergenceof modulated periodic waves; see Kamchatnov et al. [7]. Thispresent study is motivated by the recent article by Grimshawet al. [8] who examined the reduced Ostrovsky equation, that is(2) with λ,ν = 0. They showed that then the equation is eitherintegrable and solutions exist for all time, or wave breakingoccurs, depending on a certain criterion on the curvature

of the initial condition. Since (3) can be regarded as thereduced Ostrovsky equation with cubic nonlinearity ratherthan quadratic nonlinearity, we expect that a similar type ofanalysis can be applied here, and that is indeed the case, withthe curvature criterion there being replaced here with slopecriteria.

A canonical form of the modified reduced Ostrovskyequation is obtained by setting u(γ /2|ν|)1/2u,t = t/γ,x = x,so that (

ut ± u2ux

2

)x

= u. (4)

Here the tilde (˜) has been removed. There are two equations,called respectively MROI, MROII according to the sign ± inthe coefficient in the cubic term, corresponding to the sign ofγ ν in (2). MROII is also known as the short pulse equationarising in nonlinear optics; see Sakovich and Sakovich [9,10],Liu et al. [11], Pelinovsky and Sakovich [12] and the referencestherein. Note that Eq. (4) has the symmetry that if u is asolution, so also is −u. Also, the equation is invariant underthe transformation u(x,t) = Du(x,t) where x = x/D,t = Dt

for any D > 0, and in particular, the slope ux = ux is invariant.The equation has a zero mass constraint for periodic orlocalized solutions, ∫

Du dx = 0, (5)

where D is the periodic, or infinite, domain. There are alsoconservation laws for “momentum” and “energy,” similar tothose for the Ostrovsky equation; see Grimshaw and Helfrich[13]. Further, multiplying (4) by ux yields

{∂

∂t± u2

2

∂x

}u2

x + 2uux

( ± u2x − 1

) = 0,

and hence there is another conservation law,

Ft ±(

u2F

2

)x

= 0, F = ∣∣1 ∓ u2x

∣∣1/2. (6)

This conservation law is crucial for the issue of integrabilityand breaking.

021201-11539-3755/2013/88(2)/021201(5) ©2013 American Physical Society

RAPID COMMUNICATIONS

E. R. JOHNSON AND R. H. J. GRIMSHAW PHYSICAL REVIEW E 88, 021201(R) (2013)

II. INTEGRABILITY

To establish integrability, we follow the same procedureused for the reduced Ostrovsky equation; see Grimshaw et al.[8]. Thus we transform to characteristic coordinates,

x = θ (X,T ) = X ±∫ T

0

U 2(X,T ′)2

dT ′,(7)

t = T , u(x,t) = U (X,T ).

Equivalently these are defined by

dx

dt= ±u2

2, where x = X at t = 0. (8)

It then follows that

UT = ut ± u2ux

2, UX = φux,

(9)

φ = θX = 1 ±∫ T

0U (X,T ′)UX(X,T ′) dT ′, φT = ±UUX.

The Jacobian of the transformation is

J = ∂(x,t)

∂(X,T )= φ.

Thus Eq. (4) becomes

UXT = φU, and so UUXT T − UT UXT = ±U 3UX. (10)

This is more conveniently written as the system

φT = ±WU, WT = φU, W = UX. (11)

Eliminating U yields the conservation law

(φ2 ∓ W 2)T = 0, (12)

showing that φ2 ∓ W 2 is conserved along the characteristicsX = constant. Importantly, this is just the conservation law (6)since

φ2 ∓ W 2 = φ2(1 ∓ u2

x

) = φ2F 2. (13)

Next consider the evolution from an initial profile u(x,0) =u0(x) for some smooth u0(x). Since X(x,0) = x,φ = 1,T = 0when t = 0,

φ2 ∓ W 2 = 1 ∓ W 20 , φF = F0, T � 0, (14)

where W0 = W (X,0) = UX(X,0) = u0x(x), F0(X) = |1 ∓u2

0x |1/2.We next introduce the transformations associating a unique

Z(X,T ) with each pair (φ,W ),

φ = F0(X) cosh Z, W = F0(X) sinh Z,

for MROI, u20x < 1, (15)

φ = F0(X) sinh Z, W = F0(X) cosh Z,

for MROI, u20x > 1, (16)

φ = F0(X) cos Z, W = F0(X) sin Z, for MROII.

(17)

These imply in turn

ZT = U, ZXT = F0(X) sinh Z, for MROI, u20x < 1,

(18)

ZT = U, ZXT = F0(X) cosh Z, for MROI, u20x > 1,

(19)

ZT = U, ZXT = F0(X) sin Z, for MROII. (20)

Thus Z satisfies the sinh-Gordon, cosh-Gordon, or sine-Gordon equation, respectively. All are integrable equations,but only (18) and (20) have soliton solutions. For furtherprogress we examine each of MROI, MROII separately.

III. ANALYSIS AND NUMERICAL RESULTS

A. The modified reduced Ostrovsky equation MROI

1. |u0x| < 1 everywhere

First we examine the case of a + sign in (4), which is themain case of interest here. If the initial slope |u0x | < 1 for allx, then 0 < F0(X) � 1 for all X. It follows from (14) that thenW 2

0 < 1 for all X, and so φ > |W | � 0 for all X,T . Breakingcannot occur and since the slope ux = UX/φ = W/φ, |ux | < 1for all x,t . The equation remains integrable for all t .

2. |u0x| > 1 somewhere

Next suppose that there is a set of intervals X1 < x < X2

where |u0x | > 1 with equality only at the end points, sothat F0(X1,2) = 0. Since F is conserved, see (14), F = 0 isconserved on characteristics; that is, F (X1,2,T ) = 0 for allT � 0. When the initial value of 1 − u2

0x takes both positiveand negative values, then as long as the solution exists, that is0 < φ < ∞, the arguments above show that the X,T domainis divided into regions where |W/φ| = |ux | > 1, namely theregion between the characteristic boundaries X = X1,2 and theremaining regions where |ux | < 1 with |ux | = 1 on X = X1,2.Moreover, the wave cannot break in the regions of the (X,T )domain where |ux | < 1.

In the region X1 < X < X2,

F0(X) = [u2

0x − 1]1/2 = [W 2

0 − 1]1/2 > 0, (21)

and it follows that |ux | > 1, throughout the region. In this case,we can use (16) to obtain the cosh-Gordon equation (19). Thiscan be written in terms of φ or F alone, as follows:{

ψX

[1 + ψ2]1/2

}T

={

ψT

[1 + ψ2]1/2

}X

= F0(1 + ψ2)1/2,

ψ = φ

F0= 1

F. (22)

Integrating (22) yields

ψX = (1 + ψ2)1/2

{− F0X

F0(1 + F 20 )1/2

+∫ T

0[1 + ψ2(X,T ′)]1/2 dT ′

}. (23)

Hence, in any region where F0X � 0,ψX > 0 and hence ψ

cannot take its minimum value in any such region. Instead a

021201-2

RAPID COMMUNICATIONS

MODIFIED REDUCED OSTROVSKY EQUATION: . . . PHYSICAL REVIEW E 88, 021201(R) (2013)

0 0.5 1 1.5 2

−1

−0.8

−0.6

−0.4

−0.2

0

0.2

0.4

0.6

0.8

1

x/

u(x

)

0 0.5 1 1.5 20

0.2

0.4

0.6

0.8

1

1.2

1.4

1.6

1.8

2

X/ππφ(X

)(a) (b)

FIG. 1. (a) The wave profile u(x,tb) at the instant of breaking, T = t = tb = 108.25, for the initial profile (24) with b = 1.00005 sothat max[u2

0x − 1)1/2] = 10−2 computed with N = 1024 nodes. (b) The Jacobian φ(X,tb) at the instant of breaking. The dashed curve showsF0(X) = |1 − u2

0x(X)|1/2. The initial profile slope satisfies |u0x | > 1 in small intervals surrounding X = π/2 and X = 3π/2. As expected from(14), the Jacobian exceeds F0(X) outside these intervals and so breaking first occurs inside these intervals.

minimum can only be reached where F0X > 0, that is, where|u0x | > 1 and is increasing.

The system (11) subject to the initial conditions that φ =1,W = u0x at T = 0 is solved to spectral accuracy followingEsler et al. [14] and Grimshaw et al. [8]. As noted by Esleret al. [14] solving in characteristics space is particularly usefulfor investigating wave breaking as the wave breaks when anorder unity quantity passes smoothly through zero. Esler et al.[14] show that the characteristic integrations can be carriedsmoothly past breaking where they agree closely with finitevolume integrations which fit “equal area” shocks to the wavesafter breaking.

The system (11) was integrated numerically with spectralaccuracy by performing the integration in Fourier space andthen normalizing in real space using the result that thetrapezium rule is spectrally accurate for periodic functions.Integrations up to T = 100 with 2048 nodes showed F wasconserved with an accuracy of 10−6. Figure 1 shows thewave profile u(x,tb) in the original space coordinate x andthe Jacobian φ(X,tb) in the characteristic coordinate X at theinstant of breaking, T = t = tb = 108.25, for the initial profile

u0x(x) = b cos x, (24)

with b = 1.00005 so that max[u20x − 1]1/2 = 10−2. For this

initial condition F0 has period π and thus so too does φ,while u0x has period 2π and hence so too do W and U . Thewave first breaks in small intervals surrounding X = π/2 andX = 3π/2 where |u0x | > 1. In particular, the wave first breaksjust below the points of maximum |u0x | at X = π/2 and X =3π/2. Let the minimum of the Jacobian φ in (0,π ) at time

T be φm(T ) and lie at Xm(T ). Figure 2(a) shows φm(T ) forthe same evolution as Fig. 1, with φm(T ) first vanishing attb = 108.25 and Fig. 2(b) shows that Xm(T ) lies below, butclose to X = π/2 as expected. Figure 3 shows the variation inthe time to breaking, tb with Fm, the value of F0 at the pointof initial maximum gradient for the initial profile (24) withvarying b. The behavior with decreasing Fm is consistent withthe relation

tb = AF−4/3m (25)

for some constant A.

B. The modified reduced Ostrovsky equation MROII

Although this case is not our main interest here, we recordhere some known results for completeness and comparisonwith the case MROI; see Liu et al. [11] and Pelinovsky andSakovich [12] for instance. Now the initial condition is suchthat F0(X) � 1 is defined for all X as a bounded smoothfunction of X. Hence MROI is integrable provided that φ �= 0,which is the case unless ux → ∞, that is, unless breakingoccurs. Also W = UX is bounded for all X,T , and so the caseφ → ∞ cannot occur. At T = 0,φ = 1,|Z| < π/2, see (17),and so φ > 0 until breaking occurs, when φ = 0,Z = ±π/2.Hence the issue of breaking/not breaking in MROI (4) froman initial condition u(x,0) = u0(x) can be restated in terms ofthe sine-Gordon (SG) equation (20), which can be written as

ZYT = sin Z, Y =∫ X

F0(X′) dX′. (26)

021201-3

RAPID COMMUNICATIONS

E. R. JOHNSON AND R. H. J. GRIMSHAW PHYSICAL REVIEW E 88, 021201(R) (2013)

0 20 40 60 80 1000

0.05

0.1

0.15

0.2

0.25

0.3

0.35

0.4

φm(T

)

T0 20 40 60 80 100

0

0.02

0.04

0.06

0.08

0.1

0.12

0.14

0.16

0.18

0.2

Tπ/2−

Xm(T

)(a) (b)

FIG. 2. (a) φm(T ), the minimum of the Jacobian in (0,π ) at time T for the evolution of Fig. 1. φm(T ) first vanishes at tb = 108.25. (b) Thedistance that Xm(T ), the position of the minimum of the Jacobian, lies below π/2.

That is, u0(x) generates a Z(X,0) = Z0(X) initial condition forthe SG equation (20) such that −π/2 < Z0(X) < π/2. Then ifthe evolving solution is such that −π/2 < Z(Y,T ) < π/2 forall T � 0, then there is no breaking. In particular, note that thekink solution of the SG equation violates this condition, andhence any initial condition for the SG equation which generatesa kink corresponds to a breaking solution for the MROII

0 50 100 150 200 250 300 3500.08

0.1

0.12

0.14

0.16

0.18

0.2

0.22

0.24

0.26

0.28

1/Fm

t bF

4/3

m

FIG. 3. The scaled time to breaking, F 4/3m tb, as a function of 1/Fm,

where Fm is the value of F0 at the point of initial maximum gradientfor the initial profile (24) with varying b.

equation. Equation (26) has an infinite set of conservationlaws inherited from the SG equation, and especially we note

(Z2

Y

2

)T

− (1 − cos Z)Y = 0,

(1 − cos Z)T −(

Z2T

2

)Y

= 0,

(− Z2

YY + Z4Y

4

)T

+ (Z2

Y cos Z)Y

= 0, (27)

(Z2

T cos Z)T

+(

− Z2T T + Z4

T

4

)Y

= 0.

Using these and other conservation laws, it can be shown that,on the one hand, there is now breaking and solutions exist forall time if the initial condition has sufficiently small amplitudeand slope, while on the other hand, an initial condition witha large amplitude and slope will break; see Liu et al. [11]and Pelinovsky and Sakovich [12]. However, a more precisebreaking condition analogous to that for MROI in Sec. III A iselusive.

IV. CONCLUSION

For the MROI equation (4), we have shown that whenthe initial condition u(x,0) = u0(x) is such that |u0x | > 1somewhere, then wave breaking inevitably occurs, and doesso in a location close to the left-hand end point of the regionwhere |u0x | > 1. On the other hand, if u0x | < 1 everywhere,then the equation is integrable, since it can be transformed tothe sinh-Gordon equation (18). The precision of this result isanalogous to that found for the reduced Ostrovsky equation,see Grimshaw et al. [8], where the slope condition is replaced

021201-4

RAPID COMMUNICATIONS

MODIFIED REDUCED OSTROVSKY EQUATION: . . . PHYSICAL REVIEW E 88, 021201(R) (2013)

by a curvature condition. On the other hand, although theMROII equation (4) can be transformed to the integrable sine-Gordon equation (20), we have not found any similar precisewave-breaking conditions, although the available numericaland analytical evidence suggests that initial conditions withsufficiently steep slopes lead to breaking, but solutions existfor all time when the initial slopes are sufficiently small, seeLiu et al. [11] and Pelinovsky and Sakovich [12].

Finally, we recall that the MROI, MROII equations areobtained from the extended Ostrovsky equation (2) by omit-

ting the quadratic nonlinear term and the third-order lineardispersive term. When wave breaking occurs, then this latterterm needs to be restored, with the outcome that the potentialdiscontinuity with an infinite slope is, in practice, resolved bythe formation of an undular bore; see Kamchatnov et al. [7].Hence, we infer that the breaking condition, now expressedin terms of the original variables, |u0x | > |γ /2ν|1/2, impliesthe formation of an undular bore in the full equation (2) (withμ = 0). Thus, as expected strong rotation (large γ ) inhibits theformation of undular bores.

[1] R. Grimshaw, in Environmental Stratified Flows, edited byR. Grimshaw (Kluwer, Dordrecht, 2001), pp. 1–27.

[2] R. H. J. Grimshaw, in Nonlinear Waves in Fluids: Recent Ad-vances and Modern Applications, CISM Courses and Lectures,Vol. 483, edited by R. H. J. Grimshaw (Springer, Berlin, 2005),pp. 1–28.

[3] K. R. Helfrich and W. K. Melville, Annu. Rev. Fluid Mech. 38,395 (2006).

[4] L. Ostrovsky, Oceanology 18(2), 119 (1978).[5] R. Grimshaw, Stud. Appl. Math. 73, 1 (1985).[6] R. H. J. Grimshaw, L. A. Ostrovsky, V. I. Shrira, and Y. A.

Stepanyants, Surv. Geophys. 19, 289 (1998).[7] A. M. Kamchatnov, Y. H. Kuo, T. C. Lin, T. L. Horng, S. C.

Gou, R. Clift, G. A. El, and R. H. J. Grimshaw, Phys. Rev. E 86,036605 (2012).

[8] R. H. J. Grimshaw, K. Helfrich, and E. R.Johnson, Stud. Appl. Math. 129, 414(2012).

[9] A. Sakovich and S. Sakovich, J. Phys. Soc. Jpn. 74, 239(2004).

[10] A. Sakovich and S. Sakovich, J. Phys. A, Math. Gen. 39, L361(2006).

[11] Y. Liu, D. Pelinovsky, and A. Sakovich, Dynamics of PDE 6,291 (2009).

[12] D. Pelinovsky and A. Sakovich, Commun. Partial Diff. Eq. 35,613 (2010).

[13] R. Grimshaw and K. R. Helfrich, IMA J. Appl. Math. 77, 326(2012).

[14] J. G. Esler, O. J. Rump, and E. R. Johnson, Phys. Fluids 21,066601 (2009).

021201-5