You are on page 1of 33

ht t p: / / www. ut a.

edu/ mat h/ preprint /


Technical Report 2008-10
Wave Breaking Phenomena and
Stability of Peakons for the
Degasperis-Procesi
Yue Liu
Wave Breaking Phenomena and Stability
of Peakons for the Degasperis-Procesi
Equation
Yue Liu

This article is dedicated to Professor Guangchang Dong for his 80th Birthday
Abstract
We survey some recent results concerning with the Degasperis-Procesi
equation, which can be derived as a member of a one-parameter family
of asymptotic shallow-water wave approximations to the Euler equa-
tions with the same asymptotic accuracy as that of the Camassa-Holm
equation. We will focus on some important results including wave
breaking phenomena, blow-up structure, global weak solutions and the
orbital stability of the peaked solitons.
AMS subject classication (2000): 35G25, 35L05, 35Q35, 35Q51, 58D05
Keywords: Camassa-Holm equation; Degasperis-Procesi equation; Peakons;
Shallow water waves; Wave breaking; Global weak solutions; Blow-up
structure; Stability of peakons
1 Introduction
Considered herein is the Degasperis-Procesi (DP) equation
(1.1) y
t
+y
x
u + 3yu
x
= 0, x R, t > 0,
with y = u u
xx
.
Degasperis and Procesi [27] studied a family of third order dispersive
nonlinear equations
(1.2) u
t

2
u
xxt
+u
xxx
+c
0
u
x
= (c
1
u
2
+c
2
u
2
x
+c
3
uu
xx
)
x
.
with six real constants c
0
, c
1
, c
2
, c
3
, , R. They [27] found that there
are only three equations from this family were asymptotically integrable up
to third order, that is, the Korteweg-de Vries (KdV) equation ( = c
2
=

Department of Mathematics, University of Texas, Arlington, TX 76019, yliu@uta.edu


1
c
3
= 0), the Camassa-Holm (CH) equation (c
1
=
3c
3
2
2
, c
2
=
c
3
2
), and one
new equation (c
1
=
2c
3

2
, c
2
= c
3
), which is called the Degasperis-Procesi
equation. By rescaling, shifting the dependent variable, and nally apply-
ing a Galilean transformation, those three completely integrable
1
equations
can be transformed into the following forms, the Korteweg-de Vries (KdV)
equation
u
t
+u
xxx
+uu
x
= 0,
the Camassa-Holm (CH) shallow water equation [5, 28, 40],
(1.3) y
t
+y
x
u + 2yu
x
= 0, y = u u
xx
,
and the Depasperis-Procesi equation of the form (1.1). These three cases
are all the completely integrable candidates for (1.2) [5, 25, 27]. Applying
a reciprocal transformation to the Degasperis-Procesi equation, Degasperis,
Holm and Hone [25] used the Painleve analysis to show the formal integra-
bility of the DP equation as Hamiltonian systems by constructing a Lax pair
and a bi-Hamiltonian structure.
Equation (1.1) was also derived as, in dimensionless space-time variables
(x, t), an approximation to the incompressible Euler equations for shallow
water under the Kodama transformation [29, 38, 39] and its asymptotic
accuracy is the same as that of the Camassa-Holm (CH) shallow water
equation, where u(t, x) is considered as the uid velocity at time t in the
spatial x-direction with momentum density y. More interestingly, the DP
equation is recently observed as a model supporting shock waves [49].
More recently, Constantin and Lannes [18] give a rigorous proof of both
the CH equation and the DP equation are valid approximation to the gov-
erning equations for water waves (see also [40] for the formal asymptotic
procedures) and also show the relevance of these two equations as models
for the propagation of shallow water waves.
To see this rigorous justication of the derivation, one can consider the
water wave equations for one-dimensional surfaces in nondimensionalized
form
_

2
x
+
2
z
= 0, in
t
,

z
= 0, at z = 1,

t

1

(
x

x
+
z
) = 0, at z = ,

t
+

2
(
x
)
2
+

2
(
z
)
2
= 0, at z = ,
where x (t, x) parameterizes the elevation of the free surface at time
t,
t
= (x, z); 1 < z < (t, x) is the uid domain delimited by the
1
Integrability is meant in the sense of the innite-dimensional extension of a classical
completely integrable Hamiltonian system: there is a transformation which converts the
equation into an innite sequence of linear ordinary dierential equations which can be
trivially integrated [52].
2
free surface and the at bottom z = 1, (t, ) is the velocity potential
associated to the ow, and and are two dimensionless parameters dened
by
=
a
h
, =
h
2

2
,
where h is the mean depth, a is the typical amplitude, and is the typical
wavelength of the waves.
Dene the vertically averaged horizontal component of the velocity by
u(t, x) =
1
1 +
_

1

x
(t, x, z)dz.
In the shallow-water scaling ( 1), one can derive the Green-Naghdi
equations [1, 35] for one-dimensional surfaces and at bottoms without any
assumption on ( = O(1)). These equations couple the free surface elevation
to the vertically averaged horizontal component of the velocity u and can
be written as
_

t
+ ((1 +)u)
x
= 0
u
t
+
x
+uu
x
=

3
1
1+
_
(1 +)
3
(u
xt
+uu
xx
u
2
x
)
_
x
,
where O(
2
) terms have been neglected. In the so-called long-wave regime
1, = O(),
the right-going wave should satisfy the KdV equation
u
t
+u
x
+
3
2
uu
x
+
1
6
u
xxx
= 0
with = u + O(, ), or a wider class of equations, referred as the BBM
equations [2](sometimes also called the regularized long-wave equations),
which provide an approximation of the exact water wave equations of the
same accuracy as the KdV equation.
u
t
+u
x
+
3
2
uu
x
+(u
xxx
+u
xxt
) = 0, with =
1
6
.
Consider now the so-called Camassa-Holm scaling, that is
1, = O(

).
With this scaling, one still has 1, the dimensionless parameter is, how-
ever, larger here than in the long wave scaling, and the nonlinear eects are
therefore stronger and it is possible that a stronger nonlinearity could allow
the appearance of breaking waves, which is a fundamental phenomenon in
the theory of water waves that is not captured by the BBM equations.
3
Dene the horizontal velocity u

( [0, 1]) at the level line of the uid


domain by
v u

(x) =
x

z=(1+)1
.
Let p R and =
1
2
(
2

1
3
), with [0, 1]. Assume
= p +, = p
1
6
+, =
2
3
p
1
6

3
2
, =
9
2
p
23
24

3
2
.
Under the Camassa-Holm scaling, one should have the following class of
equations for v u

( [0, 1]), namely


() v
t
+v
x
+
3
2
vv
x
+(v
xxx
+v
xxt
) = (vv
xxx
+v
x
v
xx
),
where O(
4
,
2
) terms have been discarded. The vertically averaged hori-
zontal velocity u and the free surface satisfy
u = u

+u

xx
+ 2u

xx
,
= u +

4
u
2
+
1
6
u
xt

_
1
6
uu
xx
+
5
48
u
2
x
_
.
By rescaling, shifting the dependent variable, and applying a Galilean
transformation, the Camassa-Holm equation
U
t
+U
x
+ 3UU
x
U
txx
= 2U
x
U
xx
+UU
xxx
can be obtained from () if the following conditions hold
< 0, ,= , = 2, = 2,
where p =
1
3
,
2
=
1
2
. The solution u

of () is transformed to the solution


U of the CH equation by
U(t, x) =
1
a
u

_
x
b
+

c
t,
t
c
_
,
with a =
2

(1 ), b
2
=
1

, =

, and c =
b

(1 ).
On the other hand, the DP equation
U
t
+U
x
+ 4UU
x
U
txx
= 3U
x
U
xx
+UU
xxx
can also be derived if the following conditions hold
< 0, ,= , =
8
3
, = 3,
where p =
77
216
,
2
=
23
36
. The solution u

of () is also transformed to the


solution U of the DP equation by
U(t, x) =
1
a
u

_
x
b
+

c
t,
t
c
_
,
4
with a =
8
3
(1 ), b
2
=
1

, =

, and c =
b

(1 ). A detailed
derivation of the CH and DP equations can be found in [18, 40].
It is well known that the KdV equation is an integrable Hamiltonian
equation that possesses smooth solitons as traveling waves. In the KdV
equation, the leading order asymptotic balance that connes the traveling
wave solitons occurs between nonlinear steepening linear dispersion. How-
ever, the nonlinear dispersion and nonlocal balance in the CH equation and
the DP equation, even in the absence of linear dispersion, can still produce
a conned solitary traveling waves
u(t, x) = c(x ct) = ce
|xct|
,
traveling at constant speed c > 0, which are called the peakons [5, 25].
Peakons of both equations are true solitons that interact via elastic collisions
under the CH dynamics, or the DP dynamics, respectively. The peakons of
the CH equation and the DP equation are orbitally stable [22, 44]. The result
of the stability of the DP equation will be discussed in the last section.
The DP equation can be rewritten as the form
(1.4) u
t
u
txx
+ 4uu
x
= 3u
x
u
xx
+uu
xxx
, t > 0, x R.
It is noted that the peaked solitons are not classical solutions of (1.4).
They satisfy the Degasperis-Procesi equation in conservation law form
(1.5) u
t
+
x
_
1
2
u
2
+p
_
3
2
u
2
__
= 0, t > 0, x R,
where p(x) =
1
2
e
|x|
, stands for convolution with respect to the spatial
variable x R, and p f = (1
2
x
)
1
f.
Since p(x) = (x), in view of the structure of Eq.(1.5), it is quite clear
why the peakons can be understood as solutions.
Wave breaking is one of the most intriguing long-standing problems of
water wave theory [61]. For models describing water waves we say that wave
breaking holds if the wave prole remains bounded, but its slope becomes
unbounded in nite time [61]. Breaking waves are commonly observed in
the ocean and important for a variety of reasons, but surprisingly little is
known about them. Indeed, breaking waves place large hydrodynamic loads
on man-made structure, transfer horizontal momentum to surface currents,
provide a source of turbulent energy to mix the upper layers of the ocean,
move sediment in shallow water, and enhance the air-sea exchange of gases
and particulate matter [7, 57]. To further understand why waves break and
what happens during and after breaking themselves, we must rst inves-
tigate the dynamics of wave breaking. Research work on breaking waves
can be divided into three categories: those concerning waves (1) before, (2)
during, and (3) after breaking. Although we are now understanding much
5
about the processes leading up to breaking, there are still some aspects of
these questions unanswered, in particular, question (3), what happens after
breaking of those waves. In this review we shall concentrate on some of the
latest results for the DP equation in the rst two categories.
The KdV equation is well-known a model for water-motion on shallow
water with a at bottom and admits interaction for its solitary waves. It,
however, does not describe breaking of wave as physical water waves do
(the KdV equation is globally well-posed for initial data in L
2
[42, 59]). On
the other hand, wave-breaking phenomena have been observed for certain
solutions to the Whitham equation [61],
u
t
+uu
x
+
_
R
K(x )u
x
(t, )d = 0
with the singular kernel
K(x) =
1
2
_
R
_
tanh

_
1/2
e
ix
d,
(see [16] for a rigorous proof). However, the numerical calculations carried
out for the Whitham equation do not support any strong claim that soli-
ton interaction can be expected [26]. As mentioned by Whitham [61], it is
intriguing to know which mathematical models for shallow water waves ex-
hibit both phenomena of soliton interaction and wave breaking. It is found
that both of the CH equation and DP equation could be rst such equations
and have the potential to become the new master equations for shallow wa-
ter wave theory [34], modeling the soliton interaction of peaked traveling
waves, wave breaking, admitting as solutions permanent waves, and being
integrable Hammiltonian systems. For the CH equation, a procedure to un-
derstand the continuation of solutions past wave breaking has been recently
presented by Bressan and Constantin [3].
As far as we know, the case of the Camassa-Holm equation (rst derived
by Fokas and Fuchssteiner [33] using the method of recursion operators as an
abstract bi-Hamiltonian equation) is well understood by now [12, 15, 16, 19]
and the citations therein, while the Degasperis-Procesi equation case is the
subject of this article. The main mathematical questions concerning with
the DP equation are the well-posedness of the initial-value problem, wave-
breaking phenomena, existence of global weak solutions, and stability of
peakons and their role in the dynamics.
Since its discovery, there has been considerable interest in the Degasperis-
Procesi equation, cf. [17, 37, 43, 49, 51, 64, 65] and the citations therein.
For example, Lundmark and Szmigielski [50] presented an inverse scattering
approach for computing n-peakon solutions to Eq.(1.5). Holm and Staley
[38] studied stability of solitons and peakons numerically to Eq.(1.5). More
recently, Liu and Yin [47] proved that the rst blow-up for Eq.(1.4) must
6
occur as wave breaking and shock waves possibly appear afterwards. It is
shown in [47] that the lifespan of solutions of the DP equation is not aected
by the smoothness and size of initial proles, but aected by the shape of
initial proles. This can be viewed as a signicant dierence between the
DP equation (or the CH equation ) and the KdV equation.
Our goal is to present a review of some signicant work of them.
Notation. As above and henceforth, we denote by the convolution. For
1 p , the norm in the Lebesgue space L
p
(R) will be written | |
L
p,
while | |
H
s, s 0 will stand for the norm in the classical Sobolev spaces
H
s
(R).
2 Comparisons between the Camassa-Holm equa-
tion and the Degasperis-Procesi equation
The DP equation is presently of great interest due to its structure (inte-
grability, special solutions presenting interesting features). While Eq.(1.1)
has an apparent similarity to Eq.(1.3), which both are important model
equations for shallow water waves with the breaking phenomena, there are
major structural dierences and it is not much to know about its qualitative
aspects. One of the novel features of the DP equation is that it has not
only peakon solitons [25], u(t, x) = ce
|xct|
, c > 0 but also shock peakons
[11, 49] of the form
u(t, x) =
1
t +k
sgn(x)e
|x|
, k > 0.
It is easy to see from [49] that the above shock-peakon solutions can be
observed by substituting (x, t) (x, t) to Eq.(1.4) and letting 0
so that it yields the derivative Burgers equation (u
t
+uu
x
)
xx
= 0, from
which shock waves form.
In the periodic case of the spatial variable, both the CH equation and
DP equation have periodic peakons [64] of the form
u
c
(t, x) = c
cosh(x ct [x ct]
1
2
)
sinh(
1
2
)
, x R, t 0, c > 0.
However, it is recently shown by Escher, Liu and Yin [31] that the the
periodic DP equation possesses the periodic shock waves given by
u
c
(t, x) =
_
_
_
_
cosh(
1
2
)
sinh(
1
2
)
t +c
_
1
sinh(x[x]
1
2
)
sinh(
1
2
)
, x R Z, c > 0,
0, x Z.
On the other hand, the isospectral problem in the Lax pair for the DP
equation is of third-order instead of second [25], and consequently is not
7
self-adjoint,

xxx
y = 0,
and

t
+
1

xx
+u
x

_
u
x
+
2
3
_
= 0,
while the isospectral problem for the CH equation is of second order [5],

xx

1
4
y = 0
and

_
1
2
u
_

1
2
u
x
= 0
(in both cases y = u u
xx
). The spectral analysis and the inverse spectral
theory for the CH equation are presented by Constantin and McKean [14,
19] and Johnson [41]. Lundmark and Szmigielski [50] presented an inverse
scattering transform (IST) method for computing npeakon solutions of the
DP equation. The approach is similar to that used by Beals, Sattinger and
Szmigielski [4] to obtain npeakon solutions of the CH equation, but the
present case does involve substantially new features as mentioned above.
It is also noted that the CH equation is a re-expression of geodesic ow
on the dieomorphism group [17] or on the Bott-Virasoro group [55], while
no such geometric derivation of the DP equation is available.
Another indication of the fact that there is no simple transformation
of the DP equation into the CH equation is the entirely dierent form of
conservation laws for these two equations [5, 25]. The following are three
useful conservation laws of the DP equation.
E
1
(u) =
_
R
y dx, E
2
(u) =
_
R
yv dx, E
3
(u) =
_
R
u
3
dx,
where y = (1
2
x
)u and v = (4
2
x
)
1
u, while the corresponding three
useful conservation laws of the CH equation are the following.
F
1
(u) =
_
R
y dx, F
2
(u) =
_
R
(u
2
+u
2
x
) dx, F
3
(u) =
_
R
(u
3
+uu
2
x
) dx.
It is observed that the corresponding conservation laws of the DP equation
are much weaker than those of the CH equation. Therefore, the issue of
if and how particular initial data generate a blow-up in nite time is more
subtle.
It is worth noticing the following result obtained by Henry [37] and
Mustafa [56]. which implies that, analogous to the case of the CH equation,
smooth solutions of the DP equation have innite propagation speed.
Proposition 2.1. Assume u
0
is a smooth function with compact support.
If the solution u with initial data u
0
of (1.4) exists on some time interval
[0, ) with > 0 and, at any time instant t [0, ), the solution u(t, ) has
compact support, then u is identically zero.
8
3 The Cauchy problem
Recall p(x) =
1
2
e
|x|
and (1
2
x
)
1
f = pf. Then Eq.(1.1) can be rewritten
as the following form
(3.1)
_
u
t
+uu
x
+
x
p (
3
2
u
2
) = 0, t > 0, x R,
u(0, x) = u
0
(x), x R.
The local well-posedness of the Cauchy problem (3.1) with initial data u
0

H
s
(R), s >
3
2
can be obtained by applying the Katos semigroup theory for
quasilinear evolution equations.
Lemma 3.1. [63] Given u
0
H
s
(R), s >
3
2
, there exist a maximal T =
T(u
0
) > 0 and a unique solution u to initial-value problem (3.1), such that
u = u(, u
0
) C([0, T); H
s
(R)) C
1
([0, T); H
s1
(R)).
Moreover, the solution depends continuously on the initial data, i.e. the
mapping u
0
u(, u
0
) : H
s
(R) C([0, T); H
s
(R)) C
1
([0, T); H
s1
(R)) is
continuous and the maximal time of existence T > 0 can be chosen to be
independent of s.
In view of Lemma 3.1, using the energy method, one can establish the
following precise blow-up scenario of strong solutions to (3.1) [63], which is
similar to the case of the CH equation.
Proposition 3.2. [63] Given u
0
H
s
(R), s >
3
2
, blow up of the solution
u = u(, u
0
) in nite time T < + occurs if and only if
liminf
tT
inf
xR
[u
x
(t, x)] = .
Proposition 3.3. [47] Assume u
0
H
s
(R), s >
3
2
. Let T be the maximal
existence time of the solution u to (3.1) guaranteed by Lemma 3.1. Then we
have
|u(t, x)|
L
3|u
0
(x)|
2
L
2
t +|u
0
(x)|
L
, t [0, T].
Remark. It is observed that if u is the solution of (1.3) with initial data
u
0
H
1
(R), then we have for all t > 0,
|u(t, )|
L

(R)

2|u(t, )|
H
1
(R)

2|u
0
()|
H
1
(R)
.
The advantage of the CH equation in comparison with the KdV equation lies
in the fact that the CH equation has peaked solitons and models wave break-
ing (the wave prole remains bounded, but its slope becomes unbounded in
nite time wave) [6]. However, note that the conservation laws of the DP
9
equation are much weaker than the those of the CH equation. In partic-
ular, one can see that the conservation law E
2
(u) for the DP equation is
equivalent to |u|
2
L
2
. Indeed, by the Fourier transform, we have
(3.2) E
2
(u) =
_
R
yvdx =
_
R
1 +
2
4 +
2
[ u()[
2
d | u|
2
L
2
= |u|
2
L
2
.
It seems that the estimate in Proposition 3.3 is the best way to control
the L

norm of the solution for the DP equation. It then follows from


Proposition 3.2 and 3.3 that the DP equation also models wave breaking in
any nite time and one can expect that all the points at which the wave
breaking occurs should be peaked points [47].
Consider the following dierential equation
(3.3)
_
q
t
= u(t, q), t [0, T),
q(0, x) = x, x R.
Applying classical results in the theory of ordinary dierential equations,
one can obtain the following result on q.
Lemma 3.4. [65] Let u
0
H
s
(R), s 3, and let T > 0 be the maximal
existence time of the corresponding solution u to (3.1). Then Eq.(3.3) has
a unique solution q C
1
([0, T) R, R). Moreover, the map q(t, ) is an
increasing dieomorphism of R with
q
x
(t, x) = exp
__
t
0
u
x
(s, q(s, x))ds
_
> 0, (t, x) [0, T) R.
The following result, analogous to the case of the CH equation, plays a
crucial role in our considerations on global existence and blow-up solutions.
It roughly says that y(t, ) does not change on the time interval where it is
well-dened. We then infer by means of the geometric interpretation of q a
very important invariant for the solutions to (3.1).
Lemma 3.5. [65] Let u
0
H
s
(R), s 3, and let T > 0 be the maximal
existence time of the corresponding solution u to (3.1). Setting y = uu
xx
,
we have
y(t, q(t, x))q
3
x
(t, x) = y
0
(x), (t, x) [0, T) R.
Proof. The result can be obtained by dierentiating the left-hand side with
respect to time and taking inton account (3.1), (3.3) and Lemma 3.4.
Remark. From Lemma 3.5, one can expect that the lifespan of solutions of
the DP equation is not aected by the smoothness and size of initial proles,
but aected by the shape of initial proles.
By Lemma 3.5, one can obtain the criteria for breaking-waves (Theorem
3.6 and Theorem 4.1).
10
Theorem 3.6. [47] Assume u
0
H
s
(R), s >
3
2
and there exists x
0

[, ] such that
_
y
0
(x) 0 if x x
0
,
y
0
(x) 0 if x x
0
.
Then initial-value problem (3.1) has a unique global strong solution
u = u(., u
0
) C([0, ); H
s
(R)) C
1
([0, ); H
s1
(R)).
Moreover, E
2
(u) =
_
R
yv dx = E
2
(u
0
), where y = (1
2
x
)u and v = (4

2
x
)
1
u, and for all t R
+
we have
(i) u
x
(t, ) [u(t, )[ on R,
(ii) |u|
2
1
6|u
0
|
4
L
2
t
2
+ 4|u
0
|
2
L
2
|u
0
|
L
t +|u
0
|
2
1
.
Proof. Note that the solution u of the DP equation satises
(3.4) u(t, x) =
e
x
2
_
x

y(t, )d +
e
x
2
_

x
e

y(t, )d
and
(3.5) u
x
(t, x) =
e
x
2
_
x

y()d +
e
x
2
_

x
e

y()d.
From the above two equations, we deduce that
(3.6)
_
u
x
(t, x) u(t, x) if x q(t, x
0
),
u
x
(t, x) u(t, x) if x q(t, x
0
).
This implies that u
x
(t, ) [u(t, )[ on R for all t [0, T). Therefore, in
view of Proposition 3.2, the global existence result can be obtained by using
the estimate of |u|
L
in Proposition 3.3.
4 Wave breaking phenomena
Taking into account invariant of y in Lemma 3.5, one can establish some
wave breaking results with dierent shapes of initial proles.
Theorem 4.1. [47] Let u
0
H
s
(R), s >
3
2
. Assume there exists x
0
R
such that
_
y
0
(x) = u
0
(x) u
0,xx
(x) 0 if x x
0
,
y
0
(x) = u
0
(x) u
0,xx
(x) 0 if x x
0
,
and y
0
changes sign. Then, the corresponding solution to (3.1) blows up in
nite time T < + and satises liminf
tT
inf
xR
[u
x
(t, x)] = .
11
The proof of the theorem is inspired by the argument of Constantin
[12]. The idea of the proof is to obtain a dierential inequality for the time
evolution of u
x
(t, q(t, x
0
)) which can be used to prove that T < . We give
only an outline and refer to [47] for the full details.
Proof. Firstly, by dierentiating Eq.(1.5) with respect to x, we have
u
tx
+uu
xx
= u
2
x
+
3
2
u
2
p
_
3
2
u
2
_
= u
2
x
+
3
2
u
2
p
_
1
2
u
2
x
+u
2
_
+
1
2
p
_
u
2
x
u
2
_
.
taking into account p
_
1
2
u
2
x
+u
2
_
(t, x)
1
2
u
2
(t, x), it then follows that
(4.1) u
tx
+uu
xx
u
2
x
+u
2
+
1
2
p
_
u
2
x
u
2
_
.
It is then deduced from the assumption of the theorem that for t [0, T)
(4.2)
_
y(t, x) 0 if x q(t, x
0
),
y(t, x) 0 if x q(t, x
0
).
Next we dene
(4.3) M(t, x) := e
x
_
x

y(t, )d, t [0, T), and


(4.4) N(t, x) := e
x
_

x
e

y(t, )d, t [0, T).


Then it is easy to see that
(4.5) M(t, q(t, x
0
))N(t, q(t, x
0
)) = u
2
(t, q(t, x
0
)) u
2
x
(t, q(t, x
0
)),
and
(4.6) M(t, q(t, x
0
))N(t, q(t, x
0
))
=
_
q(t,x
0
)

y(t, )d
_

q(t,x
0
)
e

y(t, )d < 0, t [0, T).


12
On the other hand, we have
dM(t, q(t, x
0
))
dt
= u(t, x
0
)M(t, q(t, x
0
))
1
2
u
2
(t, q(t, x
0
))
u(t, q(t, x
0
))u
x
(t, q(t, x
0
)) +u
2
x
(t, q(t, x
0
)) +e
q(t,x
0
)
_
q(t,x
0
)

3
2
e

u
2
(t, )d
= u
2
x
(t, q(t, x
0
)) +
1
2
e
q(t,x
0
)
_
q(t,x
0
)

_
u
2
(t, ) u
2

(t, )
_
d

3
2
u
2
(t, q(t, x
0
)) +e
q(t,x
0
)
_
q(t,x
0
)

_
1
2
u
2

(t, ) +u
2
(t, )
_
d
M(t, q(t, x
0
))N(t, q(t, x
0
)) +
1
2
e
q(t,x
0
)
_
q(t,x
0
)

_
u
2
(t, ) u
2

(t, )
_
d.
(4.7)
The following estimate is crucial to ensure increase of M at t.
(4.8)
e
q(t,x
0
)
_
q(t,x
0
)

_
u
2
(t, ) u
2

(t, )
_
d M(t, q(t, x
0
))N(t, q(t, x
0
)).
Combining (4.17) with (4.18), we obtain
(4.9)
dM(t, q(t, x
0
))
dt

1
2
M(t, q(t, x
0
))N(t, q(t, x
0
)) > 0.
In an analogous way, we have the estimate of N.
(4.10)
e
q(t,x
0
)
_

q(t,x
0
)
e

_
u
2
(t, ) u
2

(t, )
_
d M(t, q(t, x
0
))N(t, q(t, x
0
))
and
(4.11)
dN(t, q(t, x
0
))
dt

1
2
M(t, q(t, x
0
))N(t, q(t, x
0
)) < 0.
Therefore,
(4.12) M(t, q(t, x
0
))N(t, q(t, x
0
)) < M(0, x
0
)N(0, x
0
) < 0, t [0, T).
On the other hand, it follows from (4.8) and (4.10) that
p (u
2
u
2
x
)(t, q(t, x
0
)) =
1
2
_

e
|q(t,x
0
)|
_
u
2
(t, ) u
2

(t, )
_
d
M(t, q(t, x
0
))N(t, q(t, x
0
)).
(4.13)
In view of (4.12) and (4.13), it is then inferred from (4.1) that
df(t)
dt

1
2
_
u
2
(t, q(t, x
0
)) u
2
x
(t, q(t, x
0
))
_
=
1
2
M(t, q(t, x
0
))N(t, q(t, x
0
)) <
1
2
M(0, x
0
)N(0, x
0
) < 0.
(4.14)
13
where the function f(t) is dened by f(t) = u
x
(t, q(t, x
0
)).
Suppose now the solution u(t) of (3.1) exists globally in time t [0, ),
that is, T = . We show this leads to a contradiction. We rst claim that
there exists t
1
> 0 such that
(4.15) f
2
(t) 2u
2
(t, q(t, x
0
)), t t
1
.
Applying Gronwalls inequality to (4.9) and (4.11) yields that
M(t, q(t, x
0
)) M(0, x
0
)e

1
2
N(0,x
0
)t
> 0,
N(t, q(t, x
0
)) N(0, x
0
)e
1
2
M(0,x
0
)t
> 0.
The above two estimates then imply that
(4.16)
u
2
x
(t, q(t, x
0
)) u
2
(t, q(t, x
0
)) M(0, x
0
)N(0, x
0
)e
1
2
(M(0,x
0
)N(0,x
0
))t
.
The proof of estimate (4.15) then follows from Proposition 3.3.
Combining (4.14) with (4.15), we obtain
(4.17)
d
dt
f(t)
1
2
u
2
(t, q(t, x
0
))
1
2
f
2
(t)
1
4
f
2
(t), t [t
1
, ).
Note that
f(0) = u
x
(0, x
0
) =
1
2
e
x
0
_
x
0

y
0
()d +
1
2
e
x
0
_

x
0
e

y
0
()d < 0.
The proof of blow-up T < + then results from the classical inequality
(4.17).
Remark. Assume the initial prole y
0
is odd. Then the solution of (3.1)
blows up in nite time if u

0
(0) < 0 [63]. One can see that the assumption
for blow-up in Theorem 4.1, that is, y
0
0 on R

and y
0
0 on R
+
, implies
u

0
(0) < 0.
The following blow-up result improves Theorem 4.1 and may also include
the case of u

0
(0) 0.
Theorem 4.2. [48] Assume u
0
H
s
(R), s >
3
2
and y
0
(x) = u
0
(x)u
0,xx
(x)
is odd. If there is only one point x
0
(0, ) such that y
0
(x
0
) = 0, then
the corresponding solution u to initial-value problem (3.1) blows up in nite
time.
Proof. The key point is how to use the assumptions to derive a dierential
inequality for the time evolution equation of u
x
(t, q(t, x
0
)). By the symmetry,
we only need to consider the case that there is a x
0
> 0 such that
_
y
0
(x) > 0 x (, x
0
),
y
0
(x) < 0 x (x
0
, 0),
14
and y
0
(x
0
) = 0. Since y
0
is odd, it follows that u
0
= p y
0
is odd. Thus, the
corresponding solution u(t, ) and y(t, ) are odd for any t [0, T). Let q(t, )
be dened in (3.3). Then q(t, ) is also odd for any t [0, T). By symmetry
of the solution, we then deduce that
(4.18)
_
y(t, x) > 0 x (, q(t, x
0
)),
y(t, x) < 0 x (q(t, x
0
), 0),
and y(t, q(t, x
0
)) = 0 for all t [0, T). In view of (4.18), we have for all
t [0, T)
(4.19) (u u
x
)(t, q(t, x
0
)) = e
q(t,x
0
)
_
q(t,x
0
)

y(t, )d > 0
and
(u
x
+u)(t, q(t, x
0
)) = e
q(t,x
0
)
_

q(t,x
0
)
e

y(t, )d
= e
q(t,x
0
)
_
_
0
q(t,x
0
)
[e

]y(t, )d +
_

q(t,x
0
)
e

y(t, )d
_
e
q(t,x
0
)
_

q(t,x
0
)
e

y(t, )d < 0.
(4.20)
From the above two relations (4.18) and (4.19), we may also obtain
(4.21) u
x
(t, q(t, x
0
)) < 0.
On the other hand, for (, q(t, x
0
)), t 0 we have
u
2
(t, ) u
2
x
(t, ) =
_

y(t, )d
_
_
q(t,x
0
)

y(t, )d +
_

q(t,x
0
)
e

y(t, )d
_

_
_
q(t,x
0
)

y(t, )d
_
q(t,x
0
)

y(t, )d
_
_

q(t,x
0
)
e

y(t, )d

_
q(t,x
0
)

y(t, )d
_

q(t,x
0
)
e

y(t, )d
= u
2
(t, q(t, x
0
)) u
2
x
(t, q(t, x
0
)).
(4.22)
15
Hence we infer from the above inequality that
d
dt
(u u
x
)(t, q(t, x
0
))
= q
t
(t, x
0
)(u u
x
)(t, q(t, x
0
)) +e
q(t,x
0
)
_
q(t,x
0
)

y
t
(t, )d
= u
2
x
(t, q(t, x
0
)) +
1
2
e
q(t,x
0
)
_
q(t,x
0
)

_
u
2
(t, ) u
2

(t, )
_
d

3
2
u
2
(t, q(t, x
0
)) +
1
2
e
q(t,x
0
)
_
q(t,x
0
)

_
u
2

(t, ) + 2u
2
(t, )
_
d
(u
2
x
u
2
)(t, q(t, x
0
)) +
1
2
e
q(t,x
0
)
_
q(t,x
0
)

_
u
2
(t, ) u
2

(t, )
_
d

1
2
[(u u
x
)(u +u
x
)](t, q(t, x
0
)) > 0.
(4.23)
It then follows from (4.19) and (4.23) that
d
dt
_
e
q(t,x
0
)
(u u
x
)(t, q(t, x
0
))
_
= e
q(t,x
0
)
q
t
(u u
x
)(t, q(t, x
0
)) +e
q(t,x
0
)
d
dt
(u u
x
)(t, q(t, x
0
))
e
q(t,x
0
)
(u
2
uu
x
)(t, q(t, x
0
)) +
1
2
e
q(t,x
0
)
(u
2
x
u
2
)(t, q(t, x
0
))

1
2
e
q(t,x
0
)
(u u
x
)(t, q(t, x
0
))[(u u
x
)(0, x
0
)] > 0.
(4.24)
This in turn implies that
_
e
q(t,x
0
)
(u u
x
)(t, q(t, x
0
))
_
e
x
0
[(u u
x
)(0, x
0
)]e
1
2
[(uu
x
)(0,x
0
)]t
.
Consequently, we deduce that
u
x
(t, q(t, x
0
))
[(u u
x
)(0, x
0
)]e
(
1
2
[(uux)(0,x
0
)]t+q(t,x
0
)x
0)
u(t, q(t, x
0
))
[(u u
x
)(0, x
0
)]e
(
1
2
[(uux)(0,x
0
)]tx
0)

_
3|u
0
|
2
L
2
t +|u
0
|
L

_
.
(4.25)
The rest of the proof can be obtained by following the outline of proof in
Theorem 4.1.
Remark. It was shown by McKean [53, 54] that the solutions of the CH
equation breaks down if and only if some portion of the positive part of
y = u
2
x
u initially lies to the left of some portion of its negative part. The
16
problem whether or not the DP equation has these wave breaking phenom-
ena still remains open. Because of the structural dierence between these
two equations, it is dicult to use the machinery of McKean [Mc2] in study
of the associated spectral problem with the corresponding eigenvalues. The
issue of if and how these particular initial data generate a global solution or
blow-up in nite time is more subtle.
In contrast with the conditions of the blow-up solution of the DP equa-
tion dened on the line R, one can see that the criteria of blow-up for
periodic solutions of the DP equation are quite dierent. Let us consider
periodic solutions of (3.1), i.e, u : S [0, T) R where S is the unit circle
and T > 0 is the maximal existence time of the solution. The interest in
periodic solutions is motivated by the observation that the majority of the
waves propagating on a channel are approximately periodic.
Dene G(x) by G(x) =
cosh(x[x]
1
2
)
2 sinh(
1
2
)
, where [x] stands for the integer part
of x R, then (1
2
x
)
1
f = G f for all f L
2
(S). Using this identity,
we can rewrite (3.1) as a quasi-linear evolution equation of hyperbolic type,
namely,
(4.26)
_

_
u
t
+uu
x
+
x
G (
3
2
u
2
) = 0, t > 0, x R,
u(0, x) = u
0
(x), x R,
u(t, x) = u(t, x + 1), t 0, x R,
Theorem 4.3. [31] Assume that u
0
H
s
(S), s >
3
2
, u
0
, 0, and the
corresponding solution u(t, x) of (4.26) has a zero for any time t 0. Then,
the solution u(t, x) of(4.26) blows up in nite time.
Proof. Proof of blow-up solution for the periodic case is quite dierent from
that of the line case (Teeorem4.1-4.2). An outline of the proof is given in
the following.
By assumption, for each t [0, T) there is a
t
[0, 1] such that u(t,
t
) =
0. Then for x S we have
(4.27) u
2
(t, x) =
__
x
t
u
x
dx
_
2
(x
t
)
_
x
t
u
2
x
dx, x
_

t
,
t
+
1
2
_
.
Hence the above relation and an integration by parts yield
_
t+
1
2
t
u
2
u
2
x
dx
_
t+
1
2
t
(x
t
)u
2
x
__
x
t
u
2
x
_
dx
1
4
_
_
t+
1
2
t
u
2
x
dx
_
2
.
Combining this estimate with a similar estimate on [
t
+
1
2
,
t
+1], we obtain
(4.28)
_
S
u
2
u
2
x
dx
1
4
__
S
u
2
x
dx
_
2
.
17
By (4.19), we also have
(4.29) sup
xS
u
2
(t, x)
1
2
_
S
u
2
x
dx.
Let us assume that the solution u(t, x) exists globally in time. Note that
G(x)
1
2 sinh(
1
2
)
for all x S. It then follows from (4.20) and (4.21) that
d
dt
_
S
u
3
x
dx = 3
_
S
u
2
x
_
u
2
x
uu
xx
+
3
2
u
2
G
_
3
2
u
2
__
dx
= 3
_
S
u
4
x
dx 3
_
S
u
2
x
uu
xx
dx +
9
2
_
S
u
2
x
u
2
dx
9
2
_
S
u
2
x
G
_
u
2
_
dx
2
_
S
u
4
x
dx +
9
8
__
S
u
2
x
dx
_
2

9
2 sinh(
1
2
)
_
S
u
2
x
dx
_
S
u
2
dx.
(4.30)
Applying the Cauchy-Schwartz inequality, we have
(4.31)
9
8
__
S
u
2
x
dx
_
2

9
8
_
S
u
4
x
dx.
On the other hand, in view of (4.20), it is easy to see that
(4.32)
_
S
u
2
x
dx
_
S
u
2
dx 2
__
S
u
2
dx
_
2

1
8
|u
0
|
4
L
2
.
It is thus deduced from (4.21)-(4.23) that
d
dt
_
S
u
3
x
dx
7
8
_
S
u
4
x
dx
9
16 sinh(
1
2
)
|u
0
|
4
L
2
, t 0.
Hence an application of Holders inequality yields
(4.33)
d
dt
_
S
u
3
x
dx
7
8
__
S
u
3
x
dx
_4
3

9
16 sinh(
1
2
)
|u
0
|
4
L
2
, t 0.
If we dene V (t) :=
_
S
u
3
x
(t, x) dx for all t 0, then
V (t) V (0)
9
16 sinh(
1
2
)
|u
0
|
4
L
2
t, t 0.
Since u
0
, 0, the above inequality implies that there exists some t
0
0 such
that V (t) < 0 for all t t
0
. It is then inferred from (4.24) that
d
dt
V (t)
7
8
(V (t))
4
3
, t > t
0
.
18
Thus we have
_
7(t t
0
)
24
+
1
(V (t
0
))
1
3
_
3

1
V (t)
< 0, t t
0
.
Since V (t
0
) < 0, the above inequality will lead to a contradiction as t t
0
is big enough, which implies T < .
As immediate consequences of Theorem 4.3, we have
Corollary 4.4. If u
0
H
3
(S), u
0
, 0 and
_
S
u
0
dx = 0 or
_
S
y
0
dx = 0,
then the corresponding solution u to (4.26) blows up in nite time.
Proof. Note
_
S
u(t, x) dx =
_
S
y(t, x) dx =
_
S
y
0
(x) dx =
_
S
u
0
(x) dx = 0.
The above relation shows that u(t, x) has a zero for all t S. It follows from
Theorem 4.3 that the solution u to (4.26) blows up in nite time.
Corollary 4.5. If u
0
H
s
(S), s >
3
2
, u
0
, 0 and
_
S
u
3
0
dx = 0, then the
corresponding solution u(t, x) of (4.26) blows up in nite time.
Proof. The result can be obtained immediately from the conservation law
E
3
(u) =
_
S
u
3
dx.
Corollary 4.6. If u
0
(x) or y
0
is odd, then the corresponding solution u to
(4.26) blows up in nite time.
Proof. If u
0
(x) or y
0
is odd, then the solution u(t, x) is odd for all t 0.
This also shows that u(t, x) has a zero for all t S.
Remark. Compared with the line case (see Theorem 3.2 and Remark 3.3
in [47]) where the corresponding solution u with u
0
H
s
(R), s > 3/2, and
u
0
(x) being odd, may exist globally in time, we nd from Theorem 4.3 that
there is quite a dierence in the blow-up phenomena of the Degasperis-
Procesi equation between the periodic case and the line case.
5 Blow-up structure
As mentioned above, blow-up can occur only in the form of wave-breaking,
i.e. the wave prole remains bounded but its slope becomes unbounded
in nite time. In this section we give a quite detailed description of its
phenomena. The following theorem shows that there is only one point where
the slope of the solution becomes innity exactly at breaking time.
19
Theorem 5.1. [47] Assume u
0
H
s
(R), s >
3
2
and there exists x
0
R
such that
_
y
0
(x) = u
0
(x) u
0,xx
(x) 0 if x x
0
,
y
0
(x) = u
0
(x) u
0,xx
(x) 0 if x x
0
,
and y
0
changes sign. Let T < be the nite blow-up time of the corre-
sponding solution u to (3.1). Then we have
lim
tT
u
x
(t, q(t, x
0
)) = .
Proof. For any x q(t, x
0
), we have
u
x
(t, x) = u(t, x) +e
x
_
q(t,x
0
)
x
e

y(t, )d +e
x
_

q(t,x
0
)
e

y(t, )d
u(t, x) +u
x
(t, q(t, x
0
)) +u(t, q(t, x
0
)).
On the other hand,
u
x
(t, x) u(t, x) +u
x
(t, q(t, x
0
)) u(t, q(t, x
0
)).
From the above two inequalities we deduce that for (t, x) [0, T) R,
u
x
(t, x) u
x
(t, q(t, x
0
)) 2|u(t, )|
L

u
x
(t, q(t, x
0
)) 2
_
3|u
0
(x)|
2
L
2
t +|u
0
(x)|
L

_
.
(5.1)
In view of T < , it follows from Proposition 3.2 that
liminf
tT
( inf
xR
u
x
(t, x)) = .
Consequently, lim
tT
u
x
(t, q(t, x
0
)) = .
Theorem 5.2. [30] Let T < be the blow-up time of the solution u of
(3.1) with initial data u
0
H
s
(R), s >
3
2
such that the associated potential
y
0
= u
0
u
0,xx
satises y
0
(x) 0 on (, x
0
] and y
0
(x) 0 on [x
0
, )
for some points x
0
R and y
0
does not have a constant sign. Then
lim
tT
_
inf
xR
u
x
(t, x)(T t)
_
= 1 and lim
tT
_
sup
xR
u
x
(t, x)(T t)
_
= 0.
The following result gives some information about the blow-up set of a
breaking wave to (3.1) with a large class of initial data.
Theorem 5.3. [30] Assume that u
0
H
s
(R), s >
3
2
and u
0
, 0 is odd such
that the associated potential y
0
= u
0
u
0,xx
is nonnegative on R

. Then the
solution u to (3.1) with initial data u
0
blows up in nite time only at zero
point.
20
For the periodic solutions of (4.26) we have
Theorem 5.4. [31] Assume that u
0
H
s
(S), s >
3
2
and that the corre-
sponding associated potential y
0
:= u
0
u
0,xx
, 0 is odd.
(a) Suppose that (x
1
2
)y
0
(x) 0 on S. Then the solution to (4.26)) with
initial data u
0
blows up in nite time only at the point x =
1
2
and
lim
tT
u
x
(t,
1
2
) = .
(b) Suppose that (x
1
2
)y
0
(x) 0 on S. Then the solution to (4.26) with
initial data u
0
blows up in nite time only at two points x = 0 and x = 1,
and
lim
tT
u
x
(t, 0) = lim
tT
u
x
(t, 1) = .
6 Global weak solutions
As mentioned in Introduction, the DP equation possesses the peaked solitons
of the form
(6.1) u(t, x) = c(x ct),
where (x) = e
|x|
. It is easy to see that (1
2
x
) = 2 (here is the Dirac
distribution). Note these peakons are not the strong solutions in H
s
, s
3
2
,
but the global weak solutions in H
1
[25, 65]. Using the conservation law of
the DP equation, a partial integration result in the Bochner spaces, and the
Helly theorem, one can obtain the following global weak solutions in H
1
.
Theorem 6.1. [30] Let u
0
H
1
(R). Assume y
0
= (u
0
u
0,xx
) M(R) and
there is a x
0
R such that supp y

0
(, x
0
) and supp y
+
0
(x
0
, ).
Then initial-value problem (3.1) has a unique global weak solution
u W
1,
loc
(R
+
R) L

loc
(R
+
; H
1
(R))
with
y(t, ) = u(t, ) u
xx
(t, )) L

loc
(R
+
, M(R).
Moreover, E
1
(u) and E
2
(u) are two conservation laws.
Example 1.(Peaked solitons) Let u
0
(x) = ce
|x|
, x R, c > 0. Then y
0
=
u
0
u
0,xx
= 2c(x) and u(t, x) = ce
|xct|
is the unique global weak solution
with the initial data u
0
.
Remark. More interestingly, we nd in [32] that there are no traveling-
wave solutions u C([0, ); H
3
) C
1
([0, ); H
2
) to (3.1). Arguing by
21
contradiction, we assume that w H
3
and u(t, x) = w(x ct), c ,= 0 is a
strong solution of (3.1). Then we have
cw

cw

4ww

+ 3w

+ww

= 0 in L
2
.
We nd that
_
cw w
2
2w
2
+ (w)
2
+ww

= 0 in L
2
and therefore
cw w
2
2w
2
+ (w)
2
+ww

= 0 in H
1
or, what is same,
(c w)(w w

)
_
w
2
(w

)
2
_
= 0 in H
1
since w H
3
C
2
0
(R). Multiplying this identity with 2w

yields that
(6.2) (c w)
_
w
2
(w

)
2
_

2w

_
w
2
(w

)
2
_
= 0.
Since w H
3
C
2
0
(R), we have w ,= c, a.e. and w
2
,= (w

)
2
, a.e.
Let w
0
= w() = max
xR
w(x) > 0. Then taking integration for (6.2) in
[, x] yields
_
x

d
_
w
2
(w

)
2
_
w
2
(w

)
2
=
_
x

2dw
c w
, x R.
This implies that
(6.3) (w c)
2

w
2
(w

)
2

= w
2
0
(w
0
c)
2
x R.
If we take into account w, w

0 as x , it is then inferred from relation


(6.3) that
w
2
0
(w
0
c)
2
= 0
which also implies from (6.3) that
(w c)
2
[w
2
(w

)
2
[ = 0, x R.
This leads a contradiction since w H
3
.
Example 2. Let
u
0
(x) = c
1
e
|xx
1
|
+c
2
e
|xx
2
|
, x R,
with c
1
< 0, c
2
> 0 and x
1
< x
2
. It is easily found that
y
0
= u
0
u
0,xx
= 2c
1
(x x
1
) + 2c
2
(x x
2
).
22
By Theorem 6.1, there exists a unique global weak solution u to (3.1) with
the initial data u
0
. It has the explicit form [25]
u(t, x) = p
1
(t)e
|xq
1
(t)|
+p
2
(t)e
|xq
2
(t)|
, (t, x) R
+
R,
for some p
1
, p
2
, q
1
, q
2
W
1,
loc
(R). Indeed, the solution u is the sum of a
peakon and an antipeakon. It is observed that the antipeakon moves o to
the left and the peakon moves o to the right so that no collision occurs.
Remark. In addition, if the initial data
u
0
(x) = c
1
e
|xx
1
|
+c
2
e
|xx
2
|
,
with c
1
> 0, c
2
< 0 and x
1
+ x
2
= 0, x
2
> 0, then the collision occurs at
x = 0 and the solution
u(x, t) = p
1
(t)e
|xq
1
(t)|
+p
2
(t)e
|xq
2
(t)|
,
(x, t) R
+
R, only satises the DP equation for t < T. The unique con-
tinuation of u into an entropy weak solution is then given by the stationary
decaying shockpeakon [49]
u(x, t) =
sgn(x)e
|x|
k + (t T)
for t T.
Existence of these discontinuous (shock waves, [49]) solutions of the DP
equation shows that the DP equation would behave radically dierent from
the Camassa-Holm equation, but similar to the inviscid Burgers equation,
which implies that a well-posedness theory should depend on some functional
spaces which contain discontinuous functions. Indeed, this observation was
conrmed by Coclite and Karlsen [8, 9]. In [8, 9], they [10] proved the global
existence and uniqueness of L
1
BV entropy weak solutions satisfying an
innite family of Kruzkov-type entropy inequalities. Recently, they proved
existence of bounded weak solutions by an Ole

inik-type estimate for L

solutions to (3.1).
Theorem 6.2. [10] Suppose u
0
L

(R). Then there exists a unique en-


tropy weak solution u L

((0, T) R), for any T > 0 to initial-value


problem (3.1) and for each T > 0 there exists a positive constant K
T
such
that the estimate
u(t, x) u)t, y)
x y
K
T
_
1
t
+ 1
_
holds for any x, y R, x ,= y, 0 < t < T.
Indeed, it is recently shown by Coclite and Karlsen [10] that the innite
family of entropy inequality is equivalent to the one-side Lipshitz inequality.
23
Therefore the well-posedness in L
1
L

of the Cauchy problem for the DP


equation can be established [10].
The relevance of these solutions of the DP equation is supported by
Lundmark [49], who found some explicit shock solutions of the DP equation
which are entropy weak solutions. Numerical schemes for computing entropy
weak solutions of the DP equation is developed and analyzed in [11]. On
the other hand, these discontinuous solutions of the DP equation are also
investigated by Feng and Liu [32] using an operator splitting method, which
consists of a second-order total variation diminishing (TVD) scheme and a
second-order linearized nite dierence scheme. Here, we rst show an ex-
ample for symmetric peakon-antipeakon collision, in which initial condition
u
0
is taken as
(6.2) u
0
(x) = e
|x+4|
e
|x4|
.
As shown in Figure 1, when t > 4, a shockpeakon is formed and continues
to decay as a shockpeakon solution mentioned previously. Actually, we
observed that shock formation is generic if u

0
(x) < 0 and u
0
(x) = 0 for
some x. We illustrate this by a numerical example with initial condition
(6.3) u
0
(x) = e
0.5x
2
sin(x),
for x [2, 2], where u
0
is extended periodically outside this interval. Figure
2 shows the numerical results up to t = 0.9. It is seen that two shocks are
formed at t 0.2, then a collision between them produces a stationary
shockpeakon at x = 0.
10 5 0 5 10
1
0.8
0.6
0.4
0.2
0
0.2
0.4
0.6
0.8
1
x
u
(
x
,
t
)
t=0
t=2
t=4
t=6
Figure 1: The numerical solution of a symmetric peakon antipeakon collision
24
2 1.5 1 0.5 0 0.5 1 1.5 2
4
3
2
1
0
1
2
3
4
x
u
(
x
,
t
)
t=0.0
t=0.2
t=0.6
t=0.9
Figure 2: Wave breaking from a smooth initial condition
7 Stability of peakons
The stability of solitary waves is one of the fundamental qualitative prop-
erties of the solutions of nonlinear wave equations. Numerical simulations
[25, 49] suggest that the sizes and velocities of the peakons do not change as
a result of collision so these patterns are expected to be stable. Furthermore,
it is observed that the shape of the peakons remains approximately the same
as time evolves. stability of the peakons for the Camassa-Holm equation is
well understood by now [21, 22], while the the case of the Degasperis-Procesi
equation is recently studied in [44].
As mentioned before, the corresponding conservation laws of the DP
equation are much weaker than those of the CH equation. In particular,
one can see that the conservation law E
2
(u) for the DP equation is only
equivalent to |u|
2
L
2
. Therefore, the stability issue of the peaked solitons of
the DP equation is more subtle.
Probably, one can only expect to obtain the orbital stability of peakons in
the sense of L
2
norm due to a weaker conservation law E
2
. The solutions of
the DP equation usually tend to be oscillations which spread out spatially in
a quite complicated way. In general, a small perturbation of a solitary wave
can yield another one with a dierent speed and phase shift. We dene the
orbit of traveling-wave solutions c to be the set U() = c( +x
0
), x
0

R, and a peaked soliton of the DP equation is called orbitally stable if a
wave starting close to the peakon remains close to some translate of it at all
later times. Let us denote
E
2
(u) = |u|
2
X
.
Theorem 7.1. [44] Let c be the peaked soliton dened in (6.1). Then c
25
is orbitally stable in the following sense. If u
0
H
s
for some s > 3/2,
y
0
= u
0

2
x
u
0
is a nonnegative Radon measure of nite total mass, and
|u
0
c|
X
< c, [E
3
(u
0
) E
3
(c)[ < c
3
, 0 < <
1
2
,
then the corresponding solution u(t) of (3.1) with initial value u(0) = u
0
satises
sup
t0
|u(t, ) c(
1
(t))|
X
< 3c
1/4
,
where
1
(t) R is the maximum point of the function v(t, ) = (4
2
x
)
1
u(t, ).
Moreover, let
M
1
(t) = v(t,
1
(t)) M
2
(t) M
n
(t) 0 and m
1
(t) m
n1
(t) 0
be all local maxima and minima of the nonnegative function v(t, ), respec-
tively. Then
(7.1)

M
1
(t)
c
6

2
and
(7.2)
n

i=2
_
M
2
i
(t) m
2
i1
(t)
_
< 2c
2

.
Remark 1. Under the assumption y
0
= u
0

2
x
u
0
0 in Theorem 7.1, the
existence is global in time [47], that is T = +. For peakons c with c > 0,
we have
_
1
2
x
_
(c) = 2c (here is the Dirac distribution). Hence the
assumption on y
0
that it is a nonnegative measure is quite natural for a
small perturbation of the peakons. Existence of global weak solution in H
1
of the DP equation is also proved in [30]. Note that peakons c are not
strong solutions, since H
s
, only for s < 3/2.
Remark 2. The above theorem of orbital stability states that any solution
starting close to peakons c remains close to some translate of c in the
norm | |
X
, at any later time. More information about this stability is
contained in (7.1) and (7.2). Notice that for peakons c, the function v
c
is
single-humped with the height
1
6
c. So (7.1) and (7.2) imply that the graph
of v(t, ) is close to that of the peakon c with a xed c > 0 for all times.
Remark 3. It is shown in [44] that M
1
= max v
u

_
E
2
(u) /12. For peakons
c, we have max v
c
=
_
E
2
(c) /12 =
1
6
c. So among all waves of a xed
energy E
2
, the peakon is tallest in terms of v
u
.
Remark 4. In addition, it is found in the proof of [44] that the peakons are
energy minimizers with a xed invariant E
3
, which explains their stability,
i. e. if E
3
(u) = E
3
(), then E
2
(u) E
2
(). The same remark also applies
26
to the CH equation and shows that the CH-peakons are energy minima with
xed F
3
.
There are two standard methods to study stability issues of dispersive
wave equations. One is the variational approach which constructs the soli-
tary waves as energy minimizers under appropriate constraints, and the
stability automatically follows. However, without uniqueness of the mini-
mizer, one can only obtain the stability of the set of minima. The variational
approach is used in [21] for the CH equation. It is shown in [21] that the
each peakon c is the unique minimum (ground state) of constrained en-
ergy, from which its orbital stability is proved for initial data u
0
H
3
with y
0
= (1
2
x
)u
0
0. Their proof strongly relies on the fact that the
conserved energy F
2
of the CH equation is the H
1
norm of the solution.
However, for the DP equation the energy E
2
is only the L
2
norm of the so-
lution. Consequently, it is more dicult to use such a variational approach
for the DP equation.
Another approach to study stability is to linearize the equation around
the solitary waves, and it is commonly believed that nonlinear stability is
governed by the linearized equation. However, for the CH and DP equations,
the nonlinearity plays the dominant role rather than being a higher-order
correction to linear terms. Thus it is unclear how one can get nonlinear sta-
bility of peakons by studying the linearized problem. Moreover, the peaked
solitons c are not dierentiable, which makes it dicult to analyze the
spectrum of the linearized operator around c.
To establish the stability result for the DP equation, we extend the
approach in [22] for the CH equation. The idea in [22] is to directly use the
energy F
2
as the Liapunov functional. By expanding F
2
around the peakon
c, the error term is in the form of the dierence of the maxima of c and
the perturbed solution u. To estimate this dierence, they establish two
integral relations
_
g
2
= F
2
(u) 2 (max u)
2
and
_
ug
2
= F
3
(u)
4
3
(max u)
3
with a function g. Relating these two integrals, one can get
F
3
(u) MF
2
(u)
2
3
M
3
, M = max u(x)
and the error estimate [M max [ then follows from the structure of the
above polynomial inequality.
To extend the above approach to nonlinear stability of the DP peakons,
one has to overcome several diculties. By expanding the energy E
2
(u)
around the peakon c, the error term turns out to be max v
c
max v
u
,
with v
u
= (4
2
x
)
1
u. We can derive the following two integral relations
27
for M
1
= max v
u
, E
2
(u) and E
3
(u) by
_
g
2
= E
2
(u) 12M
2
1
and
_
hg
2
= E
3
(u) 144M
3
1
with some functions g and h related to v
u
. To get the required polynomial
inequality from the above two identities, we need to show h 18 max v
u
.
However, since h is of the form
2
x
v
u
6
x
v
u
+ 16v
u
, generally it can
not be bounded by v
u
. This new diculty is due to the more complicated
nonlinear structure and weaker conservation laws of the DP equation. To
overcome it, we introduce a new idea. By constructing g and h piecewise
according to monotonicity of the function v
u
, we then establish two new
integral identities related to E
2
, E
3
and all local maxima and minima of
v
u
. The crucial estimate h 18 max v
u
can now be shown by using this
monotonicity structure and properties of the DP solutions. Then one can
obtain not only the error estimate [M
1
max v
c
[ but more precise stability
information from (7.2). It is observed that the same approach can also be
used for the CH equation to gain more stability information.
The detailed proof of stability of peakons for the DP equation can be
found in [44].
References
[1] B. Alvarez-Samaniego and D. Lannes, Large time existence for 3D water-
waves and asymptotics, Invent. Math., 171 (2008), 485541.
[2] T.b. Benjamin, J. L. Bona and J. J. Mahoney, Lmodel equations for
long waves in nonlinear dispersive system, Philos. trans. R. Soc. Lond. A,
227 (1972), 4778.
[3] A. Bressan and A. Constantin, Global conservative solutions of the
Camassa-Holm equation, Arch. Rat. Mech. Anal., 183 (2007), 215239.
[4] R. Beals, D.H. Sattinger and J. Szmigielski, Multipeakons and the
classical moment problem, Adv. Math., 154 (2000), 229257.
[5] R. Camassa and D. Holm, An integrable shallow water equation with
peaked solitons, Phys. Rev. Letters, 71 (1993), 16611664.
[6] R. Camassa, D. Holm and J. Hyman, A new integrable shallow water
equation, Adv. Appl. Mech., 31 (1994), 133.
[7] E. D. Cokelet, Breaking waves, Nature, 267 (1997), 769-774.
[8] G. M. Coclite and K. H. Karlsen, On the well-posedness of the
Degasperis-Procesi equation, J. Funct. Anal., 233 (2006), 6091.
28
[9] G. M. Coclite and K. H. Karlsen, On the uniqueness of discontinuous
solutions to the Degasperis-Procesi equation, J. Dierential Equations, 234
(2007), 142160.
[10] G. M. Coclite and K. H. Karlsen, Bounded solutions for the Degasperis-
Procesi equation, Boll. Unione Mat. Ital. Sez. B Artic. Ric. Mat. (8), to
appear.
[11] G. M. Coclite, K. H. Karlsen and N. H. Risebro, Numerical schemes
for computing discontinuous solutions of the Degasperis-Procesi equation,
IMA J. Numer. Anal., to appear.
[12] A. Constantin, Global existence of solutions and breaking waves for a shal-
low water equation: a geometric approach, Ann. Inst. Fourier (Grenoble), 50
(2000), 321362.
[13] A. Constantin, Finite propagation speed for the Camassa-Holm equation,
J. Math. Phys., 46 (2005), 023506, 4 pp.
[14] A. Constantin, On the scattering problem for the Camassa-Holm equation,
Proc. R. Soc. lond. A 457 (2001), 953970.
[15] A. Constantin and J. Escher, Global existence and blow-up for a shallow
water equation, Annali Sc. Norm. Sup. Pisa, 26 (1998), 303328.
[16] A. Constantin and J. Escher, Wave breaking for nonlinear nonlocal shal-
low water equations, Acta Mathematica, 181 (1998), 229243.
[17] A. Constantin and B. Kolev, Geodesic ow on the dieomorphism group
of the circle, Comment. Math. Helv., 78 (2003), 787804.
[18] A. Constantin and D. Lannes, The hydrodynamical relevance of the
Camassa-Holm and Degasperis-Procesi equations, Arch. Rantional Mech.
Anal., (2008), to appear.
[19] A. Constantin and H. P. McKean, A shallow water equation on the circle,
Comm. Pure Appl. Math., 52 (1999), 949982.
[20] A. Constantin and L. Molinet, Global weak solutions for a shallow water
equation, Comm. Math. Phys., 211 (2000), 4561.
[21] A. Constantin and L. Molinet, Obtital stability of solitary waves for a
shallow water equation, Physica D, 157 (2001), 7589.
[22] A. Constantin and W. A. Strauss, Stability of peakons, Comm. Pure
Appl. Math., 53 (2000), 603610.
[23] A. Constantin and W. A. Strauss, Stability of the Camassa-Holm soli-
tons, J. Nonlinear Science, 12 (2002), 415422.
[24] H. H. Dai, Model equations for nonlinear dispersive waves in a compressible
Mooney-Rivlin rod, Acta Mechanica, 127(1998), 193207.
29
[25] A. Degasperis, D. D. Holm, and A. N. W. Hone, A New Integral
Equation with Peakon Solutions, Theoretical and Mathematical Physics, 133
(2002), 1463-1474.
[26] R. K. Dodd, J. C. Eilbeck, J. D. Gibbon and H. C. Morris, Solitons
and nonlienar wave equations, Academic Press, New York, 1984.
[27] A. Degasperis and M. Procesi, Asymptotic integrability, in Symmetry and
Perturbation Theory, edited by A. Degasperis and G. Gaeta, World Scientic
(1999), 2337.
[28] H. R. Dullin, G. A. Gottwald, and D. D. Holm, An integrable shallow
water equation with linear and nonlinear dispersion, Phys. Rev. Letters, 87
(2001), 45014504.
[29] H. R. Dullin, G. A. Gottwald, and D. D. Holm, Camassa-Holm,
Korteweg-de Vries-5 and other asymptotically equivalent equations for shallow
water waves, Fluid Dynam. Res., 33 (2003), 7395.
[30] J. Escher, Y. Liu, and Z. Yin, Global weak solutions and blow-up structure
for the Degasperis-Procesi equation, J. Funct. Anal., 241 (2006), 457-485.
[31] J. Escher, Y. Liu, and Z. Yin, Shock waves and blow-up phenomena for
the periodic Degasperis-Procesi equation, Indiana Univ. Math. J., 56 (2007),
87117.
[32] B. Feng and Y. Liu, The non-smooth solutions of the Degasperis-Procesi
equation and its numerical aspects, preprint.
[33] A. Fokas and B. Fuchssteiner, Symplectic structures, their Backlund
transformation and hereditary symmetries, Physica D, 4 (1981), 4766.
[34] B. Fuchssteiner, Some tricks from the symmetrytoolbox for nonlinear equa-
tions: generalizations of the Camassa-Holm equation, Physica D, 95 (1996),
296-343.
[35] A.E. Green and P.M.Naghdi, A derivation of equations for wave propa-
gation in water of variable depth, J. Fluid Mech., 78(1976), 237-246.
[36] G. Gui, Y. Liu, and L. Tian, Global existence and blow-up phenomena for
the peakon b-family of equations, Indiana Univ. Math. J., (2007), accepted.
[37] D. Henry, Innite propagation speed for the Degasperis-Procesi equation. J.
Math. Anal. Appl., 311 (2005), 755759.
[38] D. D. Holm and M. F. Staley, Wave structure and nonlinear balances in a
family of evolutionary PDEs, SIAM J. Appl. Dyn. Syst. (electronic), 2 (2003),
323380.
[39] D. D. Holm and M. F. Staley, Nonlinear balance and exchange of stability
in dynamics of solitons, peakons, ramps/clis and leftons in a 1 1 nonlinear
evolutionary PDE, Phys. Lett. A, 308 (2003), 437-444.
30
[40] R. S. Johnson, Camassa-Holm, Korteweg-de Vries and related models for
water waves, J. Fluid Mech., 455 (2002), 6382.
[41] R. S. Johnson, On solution of the Camassa-Holm equation, Proc. r. Soc.
Lond., A 459 (2003), 16871708.
[42] C. Kenig, G. Ponce, and L. Vega, Well-posedness and scattering results
for the generalized Korteweg-de Veris equation via the contraction principle,
Comm. Pure Appl. Math., 46 (1993), 527620.
[43] J. Lenells, Traveling wave solutions of the Degasperis-Procesi equation, J.
Math. Anal. Appl., 306 (2005), 7282.
[44] Z. Lin and Y. Liu, Stabiltiy of peakons for the Degasperis-Procesi equation,
Comm. Pure Appl. Math., LXI (2008), 1-22.
[45] Y. Li and P. Olver, Well-posedness and blow-up solutions for an integrable
nonlinearly dispersive model wave equation, J. Dierential Equations, 162
(2000), 2763.
[46] Y. Liu, Global existence and blow-up solutions for a nonlinear shallow water
equation, Math. Ann., 335 (2006), 717-735.
[47] Y. Liu and Z. Yin, Global existence and blow-up phenomena for the
Degasperis-Procesi equation, Comm. Math. Phys., 267 (2006), 801-820.
[48] Y. Liu and Z. Yin, On the blow-up phenomena for the Degasperis-Projesi
equation, Int. Math. Res. Not., 117 (2007), 22 pages.
[49] H. Lundmark, Formation and dynamics of shock waves in the Degasperis-
Procesi equation, J. Nonlinear Sci., 17 (2007), 169-198.
[50] H. Lundmark and J. Szmigielski, Multi-peakon solutions of the
Degasperis-Procesi equation, Inverse Problems, 19 (2003), 12411245.
[51] Y. Matsuno, Multisoliton solutions of the Degasperis-Procesi equation and
their peakon limit, Inverse Problems, 21 (2005), 15531570.
[52] H. P. Mckean, Integrable systems and algebraic curves, Global Analysis,
Springer Lecture Notes in Mathematics, 755 (1979), 83200.
[53] H. P. Mckean, Integrable systems and algebraicBreakdown of a shallow
water equation, Asian J. Math., 2 (1998), 867-874.
[54] H. P. Mckean, Breakdown of the Camassa-Holm equation, Comm. Pure
Appl. Math., LVII (2004), 0416-0418.
[55] G. Misiolek, A shallow water equation as a geodesic ow on the Bott-
Virasoro group, J. Geom. Phys., 24 (1998), 203208.
[56] O. G. Mustafa, A note on the Degasperis-Procesi equation, J. Nonlinear
Math. Phys., 12 (2005), 1014.
31
[57] D. H. Peregrine, Breaking waves on beaches, Ann. Rev. Fluid Mech., 15
(1983), 149178.
[58] V. O. Vakhnenko and E. J. Parkes, Periodic and solitary-wave solutions
of the Degasperis-Procesi equation, Chaos Solitons Fractals, 20 (2004), 1059
1073.
[59] T. Tao, Low-regularity global solutions to nonlinear dispersive equations,
Surveys in analysis and operator theory (Canberra,2001), 19-48, Proc. Centre
Math. Appl. Austral. Nat. Univ., 40, Austral. Nat. Univ., Canberra, 2002.
[60] L. Tian, G. Gui, and Y. Liu, On the Cauchy problem and scattering prob-
lem for the Dullin-Gottwald-Holm equation, Comm. Math. Phys., 257 (2005),
667-701.
[61] G. B. Whitham, Linear and Nonlinear Waves, J. Wiley & Sons, New York,
1980.
[62] Z. Xin and P. Zhang, On the weak solutions to a shallow water equation,
Comm. Pure Appl. Math., 53 (2000), 14111433.
[63] Z. Yin, On the Cauchy problem for an integrable equation with peakon solu-
tions, Illinois J. Math., 47 (2003), 649-666.
[64] Z. Yin, Global weak solutions to a new periodic integrable equation with
peakon solutions, J. Funct. Anal., 212 (2004), 182-194.
[65] Z. Yin, Global solutions to a new integrable equation with peakons, Indiana
Univ. Math. J., 53 (2004), 1189-1210.
32

You might also like