• No results found

WELL-POSEDNESS OF HIGHER-ORDER CAMASSA-HOLM EQUATIONS

N/A
N/A
Protected

Academic year: 2022

Share "WELL-POSEDNESS OF HIGHER-ORDER CAMASSA-HOLM EQUATIONS"

Copied!
34
0
0

Laster.... (Se fulltekst nå)

Fulltekst

(1)

Pure Mathematics No. 15 ISSN 0806–2439 October 2006

WELL-POSEDNESS OF

HIGHER-ORDER CAMASSA–HOLM EQUATIONS

G. M. COCLITE, H. HOLDEN, AND K. H. KARLSEN

Abstract. We consider higher-order Camassa–Holm equations describing ex- ponential curves of the manifold of smooth orientation preserving diffeomor- phisms of the unit circle in the plane. We establish the existence of a strongly continuous semigroup of global weak solutions. We also present some invariant spaces under the action of that semigroup. Moreover, we prove a “weak equals strong” uniqueness result.

1. Introduction

Consider the unit circle S1 in the plane and the manifold D of the smooth orientation-preserving diffeomorphisms ofS1. Following [14] we study the equation for the exponential curves on D using the Riemannian structure induced by the Sobolev inner product (·,·)Hk(R), k ∈ N (where we identify H0(R) and L2(R)).

Letk∈Nand

Γ :t≥07→u(t,·)∈ D

be a curve. It is a an exponential curve if it satisfies the following equation [14, (3.7)]

(1.1) ∂tu=Bk(u, u), t >0, x∈R, where (see [14, (3.2), (3.3), and Proof of Theorem 2])

Bk(u, u) :=A−1k Ck(u)−u∂xu, Ak(u) :=

k

X

j=0

(−1)jx2ju, Ck(u) :=−uAkxu

+Ak u∂xu

−2∂xuAk(u).

In the casesk= 0 andk= 1, (1.1) becomes the inviscid Burgers equation [24]

(1.2) ∂tu+ 3u∂xu= 0,

and the Camassa–Holm equation [2, 9]

(1.3) ∂tu−∂t2xu+ 3u∂xu= 2∂xu∂x2u+u∂x3u,

respectively (see [14, Examples 1 and 2]). This infinite sequence of higher-order Camassa–Holm equations is distinct from what is normally called the Camassa–

Holm hierarchy, where the equations beyond the Camassa–Holm equation itself are non-local and all equations are completely integrable in the sense that one can find

Date: October 24, 2006.

1991Mathematics Subject Classification. 35G25, 35L05, 35A05.

Key words and phrases. Higher order Camassa–Holm equation, weak solutions, existence, uniqueness, geometric interpretation.

The research is supported in part by the Research Council of Norway. The research of KHK is supported by an Outstanding Young Investigators Award from the Research Council of Norway.

The current address of GMC is Department of Mathematics, University of Bari, Via E. Orabona 4, 70125 Bari, Italy.

1

(2)

a zero-curvature formulation for each equation in the hierarchy. Indeed, that is the main mechanism behind their construction. For details about the Camassa–Holm hierarchy, see [20, 21] and references therein.

In this paper we study the wellposedness of the equation (1.1). In particular, we show that it possesses a globally defined weak solution uin C([0,∞);Ck−1(R))∩ L [0,∞);Hk(R)

when the initial data u0 ∈ Hk(R), ∂xku0 ∈ Lp(R), for some 2< p <∞, see Definition 2.3 and Theorem 2.4. Furthermore, we show the existence of a semigroup Stassociated with the problem in the sense that u=St(u0) solves the equation (1.1) with initial datau0. The semigroup is continuous in the following sense: If u0,n → u0 in Hk(R), then S(u0,n) → S(u0) in L([0, T];Hk(R)), see Theorem 2.4.

Similar results holds for the Camassa–Holm equation (1.3) in the casek= 1 [5].

This equation models the propagation of unidirectional shallow water waves on a flat bottom, andu(t, x) represents the fluid velocity at timetin the horizontal di- rectionx[2, 23]. The Camassa–Holm equation possesses a bi-Hamiltonian structure (and thus an infinite number of conservation laws) [19, 2] and is completely inte- grable [2, 1, 12, 8]. Moreover, it has an infinite number of solitary wave solutions, called peakonsdue to the discontinuity of their first derivatives at the wave peak, interacting like solitons: u(t, x) =ce−|x−ct|, c∈R. From a mathematical point of view the Camassa–Holm equation is well studied. Local well-posedness results are proved in [10, 22, 25, 27]. It is also known that there exist global solutions for a particular class of initial data and also solutions that blow up in finite time for a large class of initial data [7, 10, 9]. Here blow up means that the slope of the solu- tion becomes unbounded while the solution itself stays bounded. More relevant for the present paper, we recall that existence and uniqueness results for global weak solutions of (1.3) are proved in [11, 13, 29, 30, 15, 16], see also [5].

On the other hand we recall that the solutions of the Burgers equation (1.2) in the case k= 0 experience shock formation and indeed it is well-posed in the space L([0,∞);BV(R)). Let us mention the Degasperis–Procesi equation [17, 18]

(1.4) ∂tu−∂t2xu+ 4u∂xu= 3∂xu∂x2u+u∂x3u.

It appears to be similar to the Camassa–Holm equation (1.3), but its solutions are in general discontinuous, see [6] and the references cited therein.

To keep the presentation short, details are presented for the case k= 2 only. In Appendices A and B we show how to extend the theory to general k >2.

The paper is organized as follows. In Section 2 we introduce the equations and state the main result. The existence result is obtained as a singular limit of a viscous regularization. The necessary a priori estimates are treated in Section 3.

In Section 4 stability with respect to the viscous regularization is proved using a homotopy argument. The necessary compactness arguments as well as regularity of the solution is obtained in Section 6. In Section 7 we prove a “weak equals strong”

uniqueness result. Appendices A and B deal with the general casek >2.

2. The governing equations and the main theorem

We construct a family of higher-order Camassa–Holm equations as follows. Let k∈N0:=N∪ {0}. Consider the equation

(2.1) ∂tu=Bk(u, u)

where u=u(t, x) : [0,∞)×R→Ris the unknown function and Bk(u, u) :=A−1k Ck(u)−u∂xu,

(3)

Ak(u) :=

k

X

j=0

(−1)jx2ju, (2.2)

Ck(u) :=−uAkxu

+Ak u∂xu

−2∂xuAk(u).

It turns out that the operator Ck(u) is a total derivative, that is, there exists a differential polynomial in udenoted by Fk such that

Ck(u) =−∂xFk(u).

One can see this as follows

−Fk(u) = Z x

−∞

Ck(u(ξ))dξ

= Z x

−∞

−uAk(∂xu) +Ak(u∂xu)−2∂xuAk(u) dξ

=

k

X

j=0

(−1)j Z x

−∞

−u∂x2jxu+∂x2j(u∂xu)−2∂xu∂x2ju dξ

=1 2

k

X

j=0

(−1)jx2j(u2)−

k

X

j=0

(−1)j Z x

−∞

u∂x2jxu+ 2∂xu∂x2ju dξ.

Lemma A.2 shows that indeed the integrand in each term is a total derivative, makingFk(u) a differential polynomial inu.

Remark 2.1. The operatorA−1k has a convolution structure, more precisely (2.3) A−1k (f)(x) =

Z

R

Gk(x−y)f(y)dy, x∈R, where Gk has Fourier transformGbk given by

Gbk(ζ) = 1

1 +ζ2+· · ·+ζ2k, ζ∈R. We also have

(2.4) Gk≥0, kGkkW2k−1,1(R),kGkkW2k−1,∞(R)≤C0, for some constant C0>0. In the special casek= 1we find

G1(x) =1 2e−|x|. We will repeatedly use that

xjAk(u) =Ak(∂xju) as well as

Z

R

vAk(w)dx= Z

R

Ak(v)w dx.

Example 2.2. (2.1)reads in the casesk= 0,1,2,3 as follows[14].

(i) For k= 0 we find the following:

tu+u∂xu=−∂x(u2) or ∂tu+ 3u∂xu= 0, which constitutes the inviscid Burgers equation, and

A0(u) =u, C0(u) =−2u∂xu, F0(u) =u2. (ii) For k= 1 we obtain the following equation:

tu+u∂xu=−∂xA−11 (u2+1

2(∂xu2))

(4)

or

tu−∂t2xu+ 3u∂xu= 2∂xu∂x2u+u∂x3u, which is the Camassa–Holm equation. Furthermore,

A1(u) =u−∂x2u, C1(u) =−2u∂xu−∂xu∂x2u, F1(u) =u2+1 2(∂xu)2. (iii) For k= 2, equation (2.1)becomes

(2.5) ∂tu+u∂xu=−A−12x

u2+1

2 ∂xu2

−1 2 ∂x2u2

−3∂xxu∂x2u

, or equivalently

(2.6) ∂tu−∂tx2u+∂t4xu+ 3u∂xu−2∂xu∂2xu−u∂x3u+ 2∂xu∂x4u+u∂x5u= 0.

In particular,

A2(u) =∂x4u−∂x2u+u, C2(u) =−uA2xu

+A2 u∂xu

−2∂xuA2(u)

=−∂xu∂x2u+ 10∂x2u∂x3u+ 3∂xu∂x4u−2u∂xu, F2(u) =−

Z x

−∞

C2(u)dx

=u2+1 2 ∂xu2

−7 2 ∂x2u2

−3∂xu∂x3u

=u2+1 2 ∂xu2

−1 2 ∂x2u2

−3∂xxu∂2xu . (iv) Fork= 3, equation (2.1)becomes

tu+u∂xu=−A−12x

u2+1 2 ∂xu2

−7 2 ∂x2u2

−3∂xu∂x3u (2.7)

+5∂xu∂x5u+ 16∂x2u∂x4u+19 2 ∂x3u2

, or equivalently

tu−∂t2xu+∂tx4u−∂t6xu+ 3u∂xu−2∂xu∂2xu−u∂x3u

+ 2∂xu∂x4u+u∂x5u−2∂xu∂x6u−u∂x7u= 0.

In particular,

A3(u) =−∂x6u+∂x4u−∂2xu+u=A2(u)−∂x6u, C3(u) =−uA3xu

+A3(u∂xu)−2∂xuA3(u)

=C2(u)−35∂x3u∂x4u−21∂x2u∂x5u−5∂xu∂x6u, F3(u) =−

Z x

−∞

C3(u)dx

=u2+1 2 ∂xu2

−7 2 ∂x2u2

−3∂xu∂x3u + 5∂xu∂5xu+ 16∂x2u∂x4u+19

2 ∂x3u2

. We are interested in the Cauchy problem

(2.8)

(∂tu=Bk(u, u), (t, x)∈(0,∞)×R, u(0, x) =u0(x), x∈R

in the casek≥2. We will assume

(2.9) u0∈Hk(R), ∂xku0∈Lp(R) for some 2< p <∞.

(5)

For the definition of weak solutions of (2.8) we reformulate the equation as a system of an hyperbolic equation and an higher order elliptic one, namely

(2.10)

(∂tu+u∂xu+∂xP = 0, Ak(P) =Fk(u).

This formulation is formally equivalent to (2.8).

Definition 2.3. We call a functionu: [0,∞)×R→Ra weak solution of (2.8)if (i) u∈C([0,∞);Ck−1(R))∩L [0,∞);Hk(R)

; (ii) usatisfies (2.10)in the sense of distributions;

(iii) u(0, x) =u0(x)for everyx∈R;

(iv) ku(t,·)kHk(R)≤ ku0kHk(R), for eacht >0.

Our main result is the following.

Theorem 2.4. Let 2 < p < ∞. There exists a strongly continuous semigroup of solutions

S: [0,∞)× Hk,p→C([0,∞);Ck−1(R))∩L [0,∞);Hk(R) associated with the Cauchy problem (2.8), where

Hk,p:=n

f ∈Hk(R)|∂xkf ∈Lp(R)o . More precisely, we have

(j) for each u0∈ Hk,p the map u(t, x) =St(u0)(x)is a weak solution of (2.8) according to Definition 2.3;

(jj) the semigroup is stable with respect to the initial condition:

(2.11) u0,n→u0 inHk(R)impliesS(u0,n)→S(u0)in L([0, T];Hk(R)), for every {u0,n}n∈N⊂ Hk,p,u0∈ Hk,p,T >0.

Moreover, the spacesHk+1(R)andHk,r,2≤r <∞are invariant under the action of S, i.e.,

S([0, T]×Hk+1(R))⊂L([0, T];Hk+1(R)), (2.12)

S([0, T]× Hk,r)⊂L([0, T];Hk,r), 2≤r <∞, (2.13)

for each T >0.

Moreover we show the uniqueness of the solution of the Cauchy problem (2.8) within the class of the maps with bounded second spatial derivative. A similar result was proved in [30], in the casek= 1, for the Camassa–Holm equation. More precisely, we prove the following “weak equals strong” uniqueness principle.

Theorem 2.5. Assume k = 2. Let ube a weak solution of the Cauchy problem (2.8) in the sense of Definition 2.3. If there exists a map b ∈ L1([0, T]), T > 0, such that

x2u(t,·) L(

R)≤b(t), t≥0,

then, uis unique within the class of the maps satisfying such a condition.

In particular, here we assumeb∈L1([0, T]), T >0,and in [30] whenk= 1, the authors assumedb∈L2([0, T]), T >0.

One should observe that the behavior of the Camassa–Holm equation (k= 1) is quite different from the behaviour of (2.8). Indeed the equation forq=∂x2u, which is a relevant quantity for (2.8), is

tq+u∂xq+Pe= 0,

(6)

where Pe is a given function that will be defined later on. On the other hand, ifu solves the Camassa–Holm equation (1.3), thenq=∂xu, which is the corresponding relevant quantity, satisfies (P is a another given function)

tq+u∂xq+1

2q2−u2+P = 0, which now contains the nonlinear termq2.

We apply the following singular perturbation approach. Letε >0, and consider the system

(2.14)





tuε+uεxuε+∂xPε=ε∂x2uε, Ak(Pε) =Fk(uε),

uε(0, x) =u0,ε(x).

We call the solution uε =uε(t, x) of (2.14) aviscous approximant to the solution u=u(t, x) of (2.8). Furthermore, we shall assume

(2.15) u0,ε∈Hk+1(R), ku0,εkHk(R)≤ ku0kHk(R), u0,ε→u0 in Hk(R).

Example 2.6. The equations (2.10) and (2.14) read in the special cases k = 0,1,2,3 as follows.

(i) For k= 0 we find the following:

tu+u∂xu+∂xP= 0, P=u2, and

tuε+uεxuε+∂xPε=ε∂x2uε, Pε=u2ε. (ii) For k= 1 we obtain the following equations

tu+u∂xu+∂xP = 0, P−∂x2P =u2+1 2(∂xu)2, and

tuε+uεxuε+∂xPε=ε∂x2uε, Pε−∂x2Pε=u2ε+1

2(∂xuε)2. (iii) For k= 2 we find

tu+u∂xu+∂xP = 0, (2.16)

x4P−∂2xP+P =u2+1 2 ∂xu2

−1 2 ∂x2u2

−3∂xxu∂x2u , and

tuε+uεxuε+∂xPε=ε∂x2uε, (2.17)

x4Pε−∂x2Pε+Pε=u2ε+1

2 ∂xuε2

−7

2 ∂2xuε2

−3∂xuεx3uε, or equivalently

(2.18) ∂tuε−∂tx2uε+∂tx4uε+ 3uεxuε−2∂xuε2xuε

−u∂x3uε+ 2∂xuεx4uε+u∂x5uε=ε∂2xuε−ε∂x4uε+ε∂x6uε. (iv) Fork= 3we find

tu+u∂xu+∂xP= 0, (2.19)

−∂x6P+∂4xP−∂x2P+P=u2+1 2 ∂xu2

−7 2 ∂x2u2

−3∂xu∂x3u + 5∂x2xu∂x3u

+ 6∂xx2u∂x3u

−3 2 ∂x3u2

.

(7)

Remark 2.7. Introducing the quantity

m:=Ak(u), mε:=Ak(uε), we have, see [14], that equations (2.10)and (2.14)equal

tm+u∂xm+ 2m∂xu= 0, (2.20)

and

tmε+uεxmε+ 2mεxuε=ε∂2xmε, (2.21)

respectively.

3. Viscous approximants: Global existence and energy estimate We begin with the existence of the viscous approximants to (2.8).

Lemma 3.1. Assume (2.9) and (2.15). Let ε > 0. Then there exists a unique global smooth solution uε = uε(t, x) of the Cauchy problem (2.14) belonging to C([0,∞);Hk+1(R)).

Proof. The proof of this statement is similar to the one of [4, Theorem 2.3], and is

therefore omitted.

Lemma 3.2 (Energy estimate). Assume (2.9)and (2.15). The identity (3.1) kuε(t,·)k2Hk(R)+ 2ε

Z t 0

xuε(τ,·)

2

Hk(R)dτ =ku0,εk2Hk(R)

holds for each t≥0 andε >0. In addition, (3.2) kuεkL([0,∞)×R), . . . ,

xk−1uε L

([0,∞)×R)≤ 1

√2ku0kHk(R), for each ε >0.

Proof. Fixt >0. Multiplying the first equation of (2.14) byAk(uε) and integrating overR, we get

Z

R

tuεAk(uε)dx−ε Z

R

x2uεAk(uε)dx (3.3)

=− Z

R

uεxuεAk(uε)dx− Z

R

xPεAk(uε)dx.

Integrating by parts we have for the left-hand side, Z

R

tuεAk(uε)dx−ε Z

R

x2uεAk(uε)dx (3.4)

=1 2

d

dtkuε(t,·)k2Hk(R)

xuε(t,·)

2 Hk(R), and, using the second equation of (2.14), we have for the right-hand side,

− Z

R

uεxuεAk(uε)dx− Z

R

xPεAk(uε)dx (3.5)

=− Z

R

uεxuεAk(uε)dx− Z

R

x(Ak(Pε))uεdx

=− Z

R

uεxuεAk(uε)dx+ Z

R

Ck(uε)uεdx

=−3 Z

R

uεxuεAk(uε)dx− Z

R

u2εAkxuε +

Z

R

uεAk uεxuε dx

=−3 Z

R

uεxuεAk(uε)dx+ Z

R

x(u2ε)Ak(uε) + Z

R

Ak(uε)uεxuεdx= 0.

(8)

Substituting (3.4) and (3.5) in (3.3), d

dtkuε(t,·)k2Hk(R)+ 2ε

xuε(t,·)

2

Hk(R)= 0.

Integrating over [0, t], we get (3.1). Finally, (3.2) is direct consequence of [26,

Theorem 8.5], equations (2.15) and (3.1).

4. Bounds on the source termPε and invariance properties with k= 2 From now on we assume k = 2. We show in Appendix B how to extend the proofs to the general case k >2.

Using Remark 2.1, we may write

(4.1) Pε=P1,ε+P2,ε,

where

P1,ε(t, x) :=

Z

R

G2(x−y)

u2ε(t, y) +1

2(∂xuε(t, y))2−1

2 ∂x2uε(t, y)2 dy, P2,ε(t, x) :=−3

Z

R

G2(x−y)h

x2uε(t, y)2

+∂xuε(t, y)∂x3uε(t, y)i dy.

Moreover, since G2is the Green’s function of the operatorA2, we have

x3P2,ε(t, x) =−3 Z

R

G0002(x−y)h

2xuε(t, y)2

+∂xuε(t, y)∂x3uε(t, y)i dy

=−3∂xuε(t, x)∂x2uε(t, x)

−3 Z

R

(G002(x−y)−G2(x−y))∂xuε(t, y)∂x2uε(t, y)dy.

Hence

(4.2) ∂x3Pε=−3∂xuεx2uε+∂x3P1,ε+P3,ε, where

P3,ε(t, x) :=−3 Z

R

(G002(x−y)−G2(x−y))∂xuε(t, y)∂x2uε(t, y)dy, for eachε >0, t≥0, x∈R.

Lemma 4.1. Assumek= 2,(2.9)and (2.15). The following inequalities hold kPε(t,·)kW2,1(R),kPε(t,·)kW2,∞(R)≤4C0ku0k2H2(R),

(4.3)

kP1,ε(t,·)kW4,1(R),kP1,ε(t,·)kW4,∞(R)≤(6C0+ 1)ku0k2H2(R), (4.4)

kP2,ε(t,·)kW2,1(R),kP2,ε(t,·)kW2,∞(R)≤2C0ku0k2H2(R), (4.5)

k∂x3Pε(t,·)kL1(R)≤(7C0+ 3)ku0k2H2(R), (4.6)

kP3,ε(t,·)kW1,1(R),kP3,ε(t,·)kW1,∞(R)≤12C0ku0k2H2(R), (4.7)

for each t≥0 andε >0.

Proof. Fixt >0. We begin by proving (4.4). Observing that,

xiP1,ε(t, x) = Z

R

diG2 dxi (x−y)

u2ε+1

2(∂xuε(t, y))2−1

2 ∂x2uε(t, y)2

dy, from (2.4) and (3.2),

k∂ixP1,ε(t,·)kLp(R)

diG2

dxi Lp(R)

Z

R

u2ε+1

2 ∂xuε2 +1

2 ∂x2uε2

dy (4.8)

≤C0ku(t,·)k2H2(R)≤C0ku0k2H2(R),

(9)

for each p∈ {1,∞}, i∈ {0, 1,2,3}.Recalling that G2 is the Green’s function of the operatorA2(see Remark 2.1), we find

(4.9) ∂4xP1,ε=∂x2P1,ε−P1,ε+u2ε+1

2 ∂xuε2

−1

2 ∂x2uε2

, hence, (4.4) is a direct consequence of (3.1), (4.8), and (4.9).

We continue by proving (4.5). Observing that,

xjP2,ε(t, x) =−3 Z

R

djG2

dxj (x−y)h

x2uε(t, y)2

+∂xuε(t, y)∂x3uε(t, y)i dy

=−3 Z

R

dj+1G2

dxj+1 (x−y)∂xuε(t, y)∂x2uε(t, y)dy, we conclude, using the H¨older inequality, (2.4) and (3.2), that

k∂xjP2,ε(t,·)kLp(R)

dj+1G2 dxj+1

Lp(

R)

Z

R

|∂xuεx2uε|dy

≤C0k∂xuε(t,·)kL2(R)k∂x2uε(t,·)kL2(R)

≤C0ku(t,·)k2H2(R)≤C0ku0k2H2(R),

for each p∈ {1,∞}, j ∈ {0, 1,2}. This proves (4.5). Clearly, estimates (4.4) and (4.5) imply (4.3).

Finally, using the H¨older inequality, (2.4) and (3.2), we obtain Z

R

|∂xuεx2uε|dx≤ k∂xuε(t,·)kL2(R)k∂x2uε(t,·)kL2(R)

(4.10)

≤ kuε(t,·)k2H2(R)≤ ku0k2H2(R), k∂xiP3,ε(t,·)kLp(R)≤3

d2+iG2

dx2+i Lp(

R)

+

diG2

dxi Lp(

R)

Z

R

|∂xuεx2uε|dx (4.11)

≤6C0ku0k2H2(R),

for p∈ {1,∞}and i∈ {0,1}. The estimates (4.4), (4.10), and (4.11) imply (4.6)

and (4.7).

Next we turn to estimates of time derivatives. Introduce the notation ΠT := [0, T]×R,

forT positive.

Lemma 4.2. Assumek= 2,(2.9)and (2.15). The following inequalities hold k∂tuε(t,·)kL2(R)≤ 1

2ku0k2H2(R)+ 4C0ku0k2H2(R)+ku0kH2(R), (4.12)

k∂txuεkL2T)≤√

2Tku0k2H2(R)+ 4C0ku0k2H2(R)

√ T+ 1

√2ku0kH2(R), (4.13)

for each T, t >0 and0< ε <1.

Proof. LetT, t >0 and 0< ε <1. From (2.17) and Lemma 3.2 and 4.1, k∂tuε(t,·)kL2(R)≤ kuε(t,·)∂xuε(t,·)kL2(R)+k∂xPε(t,·)kL2(R)

+εk∂2xuε(t,·)kL2(R)

≤ kuεkL([0,∞)×R)k∂xuε(t,·)kL2(R)+k∂xPε(t,·)kL2(R)

+εk∂2xuε(t,·)kL2(R)

≤ 1

2ku0k2H2(R)+ 4C0ku0k2H2(R)+εku0kH2(R), this proves (4.12).

(10)

Moreover, differentiating (2.17) with respect tox, we get (4.14) ∂txuε+ ∂xuε2

+uε2xuε+∂x2Pε=ε∂3xuε, then, from Lemmas 3.2 and 4.1 we find that

k∂txuεkL2T)

≤ k∂xuεk2L4T)+kuεx2uεkL2T)+k∂x2PεkL2T)+εk∂x3uεkL2T)

≤ 1

√2ku0k2H2(R)

T+kuεkLk∂2xuεkL2T)+k∂x2PεkL2T)+εk∂x3uεkL2T)

≤√

2Tku0k2H2(R)+ 4C0ku0k2H2(R)

√ T+ 1

2ku0kH2(R),

which proves (4.13).

Lemma 4.3. Assume k = 2, (2.9) and (2.15). Let T > 0. There exists two positive constantsK1,T, K2,T depending only onku0kH2(R)andT and independent of ε, such that

k∂tx3P1,εkL1T),k∂tx3P1,εkLT)≤K1,T, (4.15)

k∂tP3,εkL1T),k∂tP3,εkLT)≤K2,T, (4.16)

for each 0< ε <1.

Proof. Fix 0< ε <1 andT >0. We begin by proving (4.15). Observe that (4.17) ∂tx2uε+ 3∂x2uεxuε+uεx3uε+∂x3Pε=ε∂x4uε,

and, from (4.2),

(4.18) ∂tx2uε+uεx3uε+∂x3P1,ε+P3,ε=ε∂4xuε.

Hence, since G2 is the Green’s function of the operator A2 (see Remark 2.1), we find from the definition of P1,ε and (4.18) that

tx3P1,ε(t, x) = Z

R

G0002(x−y)

xuεtxuε+ 2uεtuε−∂x2uεtx2uε dy (4.19)

= Z

R

G0002(x−y) [∂xuεtxuε+ 2uεtuε]dy +

Z

R

G0002(x−y)

2xuεx3P1,ε+∂x2uεP3,ε dy +

Z

R

G0002(x−y)

uε2xuεx3uε−ε∂4xuεx2uε dy

= Z

R

G0002(x−y) [∂xuεtxuε+ 2uεtuε]dy +

Z

R

G0002(x−y)

2xuεx3P1,ε+∂x2uεP3,ε dy +1

2uε2xuε2

−ε∂x3uεx2uε +

Z

R

(G002−G2) (x−y) 1

2uε2xuε2

−ε∂x3uεx2uε

dy

− Z

R

G0002(x−y) 1

2∂xuε2xuε2

−ε ∂x3uε2

dy.

Using the H¨older inequality, k∂tx3P1,εkL1T)≤C0

Z

ΠT

h|∂xuεtxuε|+ 2|uεtuε|

(11)

+|∂x2uεx3P1,ε|+|∂x2uεP3,ε|+|uε| ∂2xuε2 (4.20)

+ 2ε|∂x3uεx2uε|+1

2|∂xuε| ∂2xuε2

+ε ∂x3uε2i dtdx +

Z

ΠT

1

2|uε| ∂x2uε

2

+ε|∂3xuεx2uε|

dtdx

≤C0

hk∂xuεkL2k∂txuεkL2+ 2kuεkL2k∂tuεkL2

+k∂2xuεkL2k∂x3P1,εkL2+k∂x2uεkL2kP3,εkL2

+kuεkLk∂x2uεk2L2+ 2εk∂x3uεkL2k∂x2uεkL2

+1

2k∂xuεkLk∂x2uεk2L2+εk∂x3uεk2L2

i

+1

2kuεkLk∂x2uεk2L2+εk∂x3uεkL2k∂x2uεkL2.

Then, the estimate (4.15) is consequence of (2.4), (3.1), (3.2), (4.4), (4.7), (4.12), and (4.13).

We continue by proving (4.16). Observing that, P3,ε(t, x) =−3

2 Z

R

(G0002(x−y)−G02(x−y)) (∂xuε(t, y))2dy,

tP3,ε(t, x) =−3 Z

R

(G0002(x−y)−G02(x−y))∂xuε(t, y)∂txuε(t, y)dy, we have

k∂tP3,εkLpT)≤3 kG0002kLp(R)+kG02kLp(R)

Z

ΠT

|∂xuεtxuε|dx

≤3 kG0002kLp(R)+kG02kLp(R)

k∂xuεkL2T)k∂txuεkL2T), (4.21)

forp∈ {1,∞}. Hence, the estimate (4.16) follows from (2.4), (3.1) and (4.13).

Now we look for invariance properties of the problem (2.17).

Lemma 4.4. Assumek= 2,(2.9)and (2.15). The following estimate holds (4.22) k∂2xuε(t,·)kLp(R)≤ k∂x2u0,εkLp(R)eK1t+K2

eK1t−1 K1 , for each t≥0,2≤p <∞andε >0, where

K1:= 1 p√

2ku0kH2(R), K2:= (18C0+ 1)2ku0k6H2(R). Proof. Let 2≤p <∞. Denote

qε:=∂x2uε, there results

(4.23) ∂tqε+ 3qεxuε+uεxqε+∂3xPε=ε∂x2qε, and, from (4.2),

(4.24) ∂tqε+uεxqε+Peε=ε∂2xqε, where

(4.25) Peε:=∂x3P1,ε+P3,ε. Multiplying (4.24) bypqε|qε|p−2 there results

t(|qε|p) +uεx(|qε|p) +pPeεqε|qε|p−2=pεqε|qε|p−2x2qε

=ε∂x2(|qε2|)−εp(p−1)(∂xqε)2. (4.26)

(12)

By (3.2), (4.4), and (4.7), pkqε(t,·)kp−1Lp(R)

d

dtkqε(t,·)kLp(R)= d dt

Z

R

|qε|pdx

≤ Z

R

xuε|qε|pdx+p Z

R

|Peε||qε|p−1dx

≤K1

Z

R

|qε|pdx+p

Peε(t,·) Lp(

R)

kqε(t,·)kp−1Lp(R)

≤K1kqε(t,·)kpLp(R)+pK2kqε(t,·)kp−1Lp(R), hence

d

dtkqε(t,·)kLp(R)≤K1

p kqε(t,·)kLp(R)+K2.

The claim is a direct consequence of the Gronwall inequality.

Lemma 4.5. Assumek= 2,(2.9)and (2.15). The following estimate holds k∂x3uε(t,·)k2L2(R)+ 2ε

Z t 0

eK3(t−τ)k∂x4uε(τ,·)k2L2(R)dτ (4.27)

≤ k∂3xu0,εk2L2(R)eK3t+K4eK3t−1 K3

, for each t≥0 andε >0, where

K3:= 1

√2ku0kH2(R)+7

2, K4:=

3

4 + 16C02

ku0k4H2(R). Proof. Using the notation from the proof of Lemma 4.4, we have (4.28) ∂txqε+∂xuεxqε−1

2q2ε+uεx2qε+1 2 ∂xuε

2

+u2ε+∂x2Pε−Pε=ε∂x3qε. By (4.28), (3.2), and (4.3),

1 2

d dt

Z

R

(∂xqε)2dx= Z

R

xqεtxqεdx

=ε Z

R

x3qεxqεdx− Z

R

xuε(∂xqε)2dx +1

2 Z

R

qε2xqεdx− Z

R

uε2xqεxqεdx

−1 2 Z

R

xuε

2

xqεdx− Z

R

u2εxqεdx

− Z

R

x2Pεxqεdx+ Z

R

Pεxqεdx

=−ε Z

R

2xqε

2 dx−1

2 Z

R

xuε(∂xqε)2dx

−1 2 Z

R

xqεxuε2

dx− Z

R

u2εxqεdx

− Z

R

x2Pεxqεdx+ Z

R

Pεxqεdx

≤ −ε Z

R

2xqε

2

dx+ 1

2k∂xuεkL([0,∞)×R)+7 4

Z

R

(∂xqε)2dx +1

4 Z

R

xuε4 dx+1

2 Z

R

u4εdx+1 2

Z

R

x2Pε2 dx+1

2 Z

R

Pε2dx

≤ −ε Z

R

2xqε2

dx+K3

2 Z

R

(∂xqε)2dx+K4

2 ,

Referanser

RELATERTE DOKUMENTER

shallow water equation, Burgers equation, conservation law, entropy condition, singular limit, compensated compactness.. The research

In the following example we will qualitatively compare how the variational scheme (34) and the conservative multipeakon scheme (20) handle smooth initial data which leads to

More recently, these metrics have been used with success to show uniqueness past the blow-up time for multidimensional aggregation equations [14] using gradient flow solutions.. It

All the α -dissipative solutions are global weak solutions of the same equation in Eulerian coordinates, yet they exhibit rather distinct behavior at wave breaking.. The solutions

From our point of view, this generalization is of special interest, not only because it has been derived in the context of shallow water waves [15], but also because weak solutions

In the context of dissipative solutions we cannot hope that we can approximate dissipative solutions of the CH equation by solutions of the 2CH system which do not enjoy wave

Benjamin–Ono equation; Godunov splitting; Strang Splitting; Error estimate; Convergence.. Supported in part by the Research Council of Norway and the Alexander von

In 1999, the Research Council worked to further the tripartite collaboration established in 1998 with the Norwegian Industrial and Regional Development Fund (SND) and the