19
rspa.royalsocietypublishing.org Research Cite this article: Szmigielski J, Zhou L. 2013 Colliding peakons and the formation of shocks in the Degasperis–Procesi equation. Proc R Soc A 469: 20130379. http://dx.doi.org/10.1098/rspa.2013.0379 Received: 6 June 2013 Accepted: 23 July 2013 Subject Areas: applied mathematics, differential equations, mathematical physics Keywords: peakons, shocks, wave breaking Author for correspondence: Jacek Szmigielski e-mail: [email protected] Colliding peakons and the formation of shocks in the Degasperis–Procesi equation Jacek Szmigielski 1 and Lingjun Zhou 2 1 Department of Mathematics and Statistics, University of Saskatchewan, 106 Wiggins Road, Saskatoon, Saskatchewan, Canada S7N 5E6 2 Department of Mathematics, Tongji University, Shanghai, People’s Republic of China The Degasperis–Procesi equation (DP) is one of several equations known to model important nonlinear effects such as wave breaking and shock creation. It is, however, a special property of the DP equation that these two effects can be studied in an explicit manner with the help of the multi-peakon ansatz. In essence, this ansatz allows one to model wave breaking as a collision of hypothetical particles (peakons and antipeakons), called henceforth collectively multi- peakons. It is shown that the system of ordinary differential equations describing DP multi-peakons has the Painlevé property which implies a universal wave breaking behaviour, that multipeakons can collide only in pairs, and that there are no multiple collisions other than, possibly simultaneous, collisions of peakon–antipeakon pairs at different locations. Moreover, it is demonstrated that each peakon– antipeakon collision results in creation of a shock thus making possible a multi-shock phenomenon. 1. Introduction The Degasperis–Procesi (DP) equation [1], u t u txx + 4uu x = 3u x u xx + uu xxx , (1.1) belongs to a class of one-dimensional wave equations that have attracted considerable attention over the past decade, following the most studied equation in this class, namely the Camassa–Holm (CH) equation [2] u t u txx + 3uu x = 2u x u xx + uu xxx . Both these equations can be derived from the governing equations for water waves under the assumption of moderate amplitude [3,4]. What makes them special is 2013 The Author(s) Published by the Royal Society. All rights reserved. on April 6, 2018 http://rspa.royalsocietypublishing.org/ Downloaded from

Colliding peakons and the formation of shocks in the Degasperis

  • Upload
    haduong

  • View
    215

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Colliding peakons and the formation of shocks in the Degasperis

rspa.royalsocietypublishing.org

ResearchCite this article: Szmigielski J, Zhou L. 2013Colliding peakons and the formation of shocksin the Degasperis–Procesi equation. Proc RSoc A 469: 20130379.http://dx.doi.org/10.1098/rspa.2013.0379

Received: 6 June 2013Accepted: 23 July 2013

Subject Areas:applied mathematics, differential equations,mathematical physics

Keywords:peakons, shocks, wave breaking

Author for correspondence:Jacek Szmigielskie-mail: [email protected]

Colliding peakons and theformation of shocks in theDegasperis–Procesi equationJacek Szmigielski1 and Lingjun Zhou2

1Department of Mathematics and Statistics, University ofSaskatchewan, 106 Wiggins Road, Saskatoon, Saskatchewan,Canada S7N 5E62Department of Mathematics, Tongji University, Shanghai,People’s Republic of China

The Degasperis–Procesi equation (DP) is one ofseveral equations known to model important nonlineareffects such as wave breaking and shock creation. It is,however, a special property of the DP equation thatthese two effects can be studied in an explicit mannerwith the help of the multi-peakon ansatz. In essence,this ansatz allows one to model wave breaking asa collision of hypothetical particles (peakons andantipeakons), called henceforth collectively multi-peakons. It is shown that the system of ordinarydifferential equations describing DP multi-peakonshas the Painlevé property which implies a universalwave breaking behaviour, that multipeakons cancollide only in pairs, and that there are no multiplecollisions other than, possibly simultaneous, collisionsof peakon–antipeakon pairs at different locations.Moreover, it is demonstrated that each peakon–antipeakon collision results in creation of a shock thusmaking possible a multi-shock phenomenon.

1. IntroductionThe Degasperis–Procesi (DP) equation [1],

ut − utxx + 4uux = 3uxuxx + uuxxx, (1.1)

belongs to a class of one-dimensional wave equationsthat have attracted considerable attention over the pastdecade, following the most studied equation in this class,namely the Camassa–Holm (CH) equation [2]

ut − utxx + 3uux = 2uxuxx + uuxxx.

Both these equations can be derived from the governingequations for water waves under the assumption ofmoderate amplitude [3,4]. What makes them special is

2013 The Author(s) Published by the Royal Society. All rights reserved.

on April 6, 2018http://rspa.royalsocietypublishing.org/Downloaded from

Page 2: Colliding peakons and the formation of shocks in the Degasperis

2

rspa.royalsocietypublishing.orgProcRSocA469:20130379

..................................................

that, on the one hand, both are Lax-integrable, on the other hand, both exhibit wave breakingphenomena not captured by the linear theory or the shallow water, small amplitude, theorysuch as the Korteweg–de Vries equation. For the CH equation, the most relevant study of thebreakdown of solutions was carried out by McKean [5,6]. In these works, it was argued thatthe breakdown of the CH waves is controlled by a kind of caricature of the higher dimensionalvorticity, namely m = u − uxx (see Majda & Bertozzi [7]). One of the fascinating aspects of boththe CH and DP equations is the existence of special solutions, peakons which play the roleof basic building blocks of the underlying full theory. Peakons are a simple superposition ofexponential terms

u(x, t) =n∑

j=1

mj(t) e−|x−xj(t)|

for which the function m referred to earlier is m = 2∑n

j=1 mjδxj . Were we to take the analogy withvorticity at its face value, m for peakons could be viewed as a collection of point vortices, situatedat xj(t), with strengths mj(t) for j = 1, . . . , n respectively, initially ordered in some fixed way, sayx1(0) < x2(0) < · · · < xn(0). The case of CH peakons shows that if the strengths {mk(0) : 1 ≤ k ≤ n}are not of the same sign, then collisions can occur, meaning that xj(tc) = xj+1(tc) for some j andsometime tc. Each collision is accompanied by a blow-up of mj(tc) and mj+1(tc) resulting in thederivative ux(tc) becoming unbounded even though u(tc) remains bounded, in fact, continuous.Thus, peakons can be used to test ideas about wave breaking, the advantage being that the peakondynamics is described by a finite system of ordinary differential equations (ODEs; see §2). Thecomplete analysis of the CH peakon collisions was carried out in Beals et al. [8,9] with the help ofexplicit formulae.

The case of the DP equation is superficially similar to the case of CH. However, deeper analysisshows a remarkable number of new features. For example, the associated spectral problem,termed a cubic string in reference [10], is not self-adjoint, and this has the immediate consequencethat the inverse problem is far more involved. The peakon problem in the case of positivemeasure, that is when all weights mj are positive, was solved explicitly in reference [10]. However,the generalization to the case of a signed measure m is not straightforward, because the spectraldata break up into several types depending on the degeneracy of the spectrum, in contrast to theCH case where the spectrum remains real and simple.

Similar to the CH case, the presence of a DP peakon collision signals an occurrence of wavebreaking; in the DP context, the connection between wave breaking and peakon collisions wasstudied earlier by Lundmark [11] for the case n = 2 and further by us [12] for n = 3.

The DP equation, however, in contrast to the CH equation, also admits shock solutions (see[13,14] for a general, very thorough discussion). It was Lundmark who introduced the concept ofshockpeakons

u(x, t) =n∑

j=1

{mj(t) − sj(t) sgn(x − xj(t))} e−|x−xj(t)|,

for which

m = 2n∑

j=1

{mjδxj + sjδ′xj}, (1.2)

and showed that the solution describing a collision of two peakons has a unique entropy extensionto shockpeakons. He also hypothesized that this might be a general phenomenon valid also forn > 2. Intuitively speaking, the collisions of DP peakons signal both the wave breaking and theonset of shocks.

Let us briefly describe our strategy. Because, in general, the spectral problem for DP peakons isquite involved, we choose instead to concentrate on analytic properties of xj(t), mj(t) as functionsof t insofar as their behaviour can be inferred from the inverse problem without explicitly solvingit. Then, we use the ODEs describing peakons time evolution to refine information about thedetails of collisions of peakons. More concretely, the plan of the paper is as follows: we review

on April 6, 2018http://rspa.royalsocietypublishing.org/Downloaded from

Page 3: Colliding peakons and the formation of shocks in the Degasperis

3

rspa.royalsocietypublishing.orgProcRSocA469:20130379

..................................................

basic facts about the DP equation in §2; in §3, we discuss the inverse problem for peakonsof both signs generalizing the uniqueness result known from the pure peakon case [10] anduse this result to establish analytic properties of positions xj and masses mj. We prove thateach xj(t) must be a holomorphic function at tc, whereas mj(t) is, in general, only meromorphic(theorem 3.5). Then, we perform singularity analysis of the ODEs describing peakons (2.6) andprove their Painlevé property with the help of theorems 3.5 and 3.6. In §4, which is the mainsection of the paper, we analyse the singular behaviour at the time of collisions and establish auniversal singularity pattern, according to which, in the leading term, only the time of the blow-up depends on the initial conditions, whereas the residue remains universal. This fact is provedin theorem 4.5. We furthermore rule out triple collisions in theorem 4.7, and give an exampleof a simultaneous collision, in different positions, of two peakon–antipeakon pairs; finally, in§5, we apply results of §4 to prove theorem 5.1 stating that the distributional limit of collidingpeakons is, indeed, a shockpeakon. More precisely, we prove that the distributional limit of mat the collision point tc produces shockpeakon data (1.2) with positive shock strengths sj thusallowing a unique entropy weak extension (see theorem 5.1 and corollary 5.2), which settles inpositive Lundmark’s conjecture.

Important questions of stability and general analytic results dealing with DP peakons and theDP wave breaking have been addressed [15–18]. A considerable amount of work has been alsocarried out on adapting numerical schemes to deal with the DP equation; we just mention a few:an operator splitting method of Feng & Liu [19], or numerical schemes discussed by Coclite et al.[20] and Hoel [21].

2. Basic facts about the Degasperis–Procesi equationThe nonlinear equation

ut − uxxt + 4uux = 3uxuxx + uuxxx (2.1)

often written as

mt + mxu + 3mux = 0, m = u − uxx (2.2)

was introduced by Degasperis & Procesi [1].Formal integrability for the DP equation was established by Degasperis et al. [22] through the

construction of a Lax pair and a bi-Hamiltonian structure. In particular, it was shown in reference[22] that the DP equation admits the Lax pair:

(∂x − ∂xxx)Ψ = zmΨ , Ψt = [z−1(1 − ∂2x ) + ux − u∂x]Ψ . (2.3)

Moreover, one can impose additional boundary conditions provided they do not violate thecompatibility of these equations. One such a pair of boundary conditions was introduced inreference [10]:

Ψ → ex, as x → −∞, Ψ is bounded as x → +∞, (2.4)

where it was also shown that the spectrum of this boundary value problem will remain timeinvariant (isospectral deformation). It suffices for our purposes to restrict our attention to the casein which m is a finite discrete (signed) measure. Thus, for the remainder of the paper, we will usethe multi-peakon ansatz

u(x, t) =n∑

i=1

mi(t) e−|x−xi(t)|, (2.5)

where x1(0) < x2(0) < · · · < xn(0) and mi(0) can have both positive and negative values. Thisansatz produces m = 2

∑ni=1 miδxi . Moreover, with the proper interpretation of weak solutions to

on April 6, 2018http://rspa.royalsocietypublishing.org/Downloaded from

Page 4: Colliding peakons and the formation of shocks in the Degasperis

4

rspa.royalsocietypublishing.orgProcRSocA469:20130379

..................................................

equation (2.1), we can easily check that u is a weak solution to (2.1) provided xi(t), mi(t) satisfy thefollowing ODEs

xk(t) = u(xk) =n∑

i=1

mi(t) e−|xk(t)−xi(t)| (2.6a)

and

mk(t) = 2 mk(t) 〈ux(xk)〉 = 2 mk(t)n∑

i=1

mi(t) sgn(xk(t) − xi(t)) e−|xk(t)−xi(t)|, (2.6b)

where 〈f (x)〉 = limε→0+ 12 [f (x + ε) + f (x − ε)] := 1

2 [f (x+) + f (x−)] is the average of f at the point x.We will refer to the coefficients mj as masses to emphasize their role in the spectral problem.We also need a bit of terminology regarding the phenomenon of breaking. We will say that acollision occurred at sometime tc if xi(tc) = xi+1(tc) for some i. We can make this concept moregeometric by introducing a configuration space in which to study peakon solutions, namelythe sector X = {x ∈ Rn |x1 < x2 < · · · < xn}. Then, a collision corresponds to the solution xi hittingthe boundary of X.

A very useful property of equations (2.6) is the existence of n constants of motion. This followsreadily from theorem 2.10 in reference [10].

Lemma 2.1. Mp(1 ≤ p ≤ n), given by,

Mp =∑

(I∈[1,k]p )

⎛⎝∏

i∈I

mi

⎞⎠⎛⎝p−1∏

j=1

(1 − exij −xij+1 )2

⎞⎠

are n constants of motion of the system of equations (2.6), where([1,n]

p)

is the set of all p-element subsetsI = {i1 < · · · < ip} of {1, . . . , n}.

3. Inverse problem for multi-peakonsThe boundary value problem (2.3) and (2.4) can be transformed to a finite interval boundary valueproblem, the cubic string problem. Indeed, following Lundmark & Szmigielski [10], the changeof variables (Liouville transformation)

y = tanhx2

, Ψ (x) = 2φ(y)1 − y2 (3.1)

maps the DP spectral problem into the cubic string problem:

− φyyy(y) = zg(y)φ(y), −1 < y < 1 (3.2a)

andφ(−1) = φy(−1) = φ(1) = 0, (3.2b)

where g is the transformation of the measure m induced by the Liouville transformation (3.1).Furthermore, as one can explicitly check, g is also a finite signed measure and its support doesnot include the endpoints if the original measure m is a finite signed measure. More concretely, inthis paper,

g =n∑

i=1

giδyi , −1 < y1 < y2 < · · · < yn < 1, (3.3)

with weights gi ∈ R. The inverse problem is studied with the help of two Weyl functions.

Definition 3.1. Let φ(y; z) denote the solution to the initial value problem (3.2a) with initialconditions φ(−1; z) = φy(−1; z) = 0, φyy(−1; z) = 1. The Weyl functions are ratios:

W(z) = φy(1; z)φ(1; z)

, Z(z) = φyy(1; z)φ(1; z)

.

on April 6, 2018http://rspa.royalsocietypublishing.org/Downloaded from

Page 5: Colliding peakons and the formation of shocks in the Degasperis

5

rspa.royalsocietypublishing.orgProcRSocA469:20130379

..................................................

These two functions encode spectral information required to solve the inverse problem. It iseasy to verify that in the case of (3.3) both W(z) and Z(z) are rational functions that make inversionalgebraic. However, in contrast to the pure peakon case gi > 0, the spectrum of the boundary valueproblem (3.2a) is, in general, complex and not necessarily simple. This makes the inversion morechallenging. Regardless of the complexity of the spectrum, the Weyl functions undergo a simpleevolution under the DP flow. Indeed, using the second member of the Lax pair given by (2.3), onecan find the time evolution of W(z) and Z(z). To wit, using results from theorem 2.3 in reference[12], we obtain the following characterization of the time evolution of W(z) and Z(z).

Theorem 3.2. Let

W(z)z

=∑

j

dj∑k=1

b(k)j

(z − λj)k+ 1

z

be the partial fraction decomposition of W(z)/z, where dj denotes the algebraic degeneracy of the jtheigenvalue. Then, the DP time evolution implies

(1)

b(k)j = p(k)

j (t)et/λj ,

where p(k)j (t) is a polynomial in t of degree dj − k.

(2) M+def= ∑n

k=1 mk exk =∑j b(1)

j .(3)

W = W − 1z

+ M+, Z = (W − 1)M+ + W

We immediately have the following.

Corollary 3.3. Under the DP flow, the Weyl functions W, Z are entire functions of time.

The uniqueness result below plays a major role in the solution to the inverse problem.

Theorem 3.4. Suppose Φ : g → {W(z), Z(z)} is the map that associates with the cubic string problem(3.2) with a finite signed measure g, the Weyl functions W(z), Z(z). Then, Φ is injective.

Proof. The proof relies on remarks made in reference [23]. We will construct a recursive schemeto solve the inverse spectral problem; given W and Z obtained from the map Φ, we willreconstruct the finite, signed measure g whose Weyl functions are W and Z. More precisely, wewill show that the yj and gj in equation (3.3) are uniquely determined from W(z), Z(z). First, werecall that W and Z are constructed from solutions to the initial value problem (see definition 3.1)

− φyyy(y) = zg(y)φ(y), −1 < y < 1 (3.4a)

and

φ(−1) = φy(−1) = 0, φyy(−1) = 1. (3.4b)

Masses gj are situated at yj, 1 ≤ j ≤ n and for convenience let us set y0 = 0, yn+1 = 1 and denote bylj = yj+1 − yj the length of the interval (yj, yj+1). Then, on each interval (yj, yj+1), the solution to(3.4) takes the form

φ(y) = φ(yj+1) + φy(yj+1)(y − yj+1) + φyy(yj+1−)(y − yj+1)2

2, 0 ≤ j ≤ n,

on April 6, 2018http://rspa.royalsocietypublishing.org/Downloaded from

Page 6: Colliding peakons and the formation of shocks in the Degasperis

6

rspa.royalsocietypublishing.orgProcRSocA469:20130379

..................................................

where notation f (a±) denotes the right-hand or left-hand limits at a. Then, the conditionof crossing yj+1 is continuity of φ and φy and the jump condition φyy(yj+1+) − φyy(yj+1−) =−zgj+1φ(yj+1). We establish, for example, by an easy induction,

(φ(yj+1), φy(yj+1), φyy(yj+1−)) = (−z)jj∏

k=1

gkl2k−1

2

(l2j2

, lj, 1

)+ O(zj−1), (3.5)

valid for 0 ≤ j ≤ n, with the convention that for j = 0 the product equals 1 and there is noremainder. Likewise,

(φ(yj), φy(yj), φyy(yj+)) = (−z)jj∏

k=1

gkl2k−1

2(0, 0, 1) + O(zj−1), (3.6)

valid for 1 ≤ j ≤ n.For 0 ≤ j ≤ n, we define (w2j, z2j) = (φy/φ, φyy/φ)|y=yj+1− and (w2j−1, z2j−1) = (φy/φyy,

φ/φyy)|y=yj+. These quantities are essentially the left-hand and the right-hand analogues of Weylfunctions introduced in definition 3.1 and correspond to shorter strings terminating at yj+1 withno mass at the endpoint, or terminating at yj but with the mass gj at the end. Equation (3.4) impliesthat the sequence (w2j, z2j, w2j−1, z2j−1) satisfies the recurrence relations

w2j−1 = w2j

z2j− lj, z2j−1 = 1

z2j− lj

w2j

z2j+

l2j2

(3.7)

and

w2j−2 = w2j−1

z2j−1, z2j−2 = 1

z2j−1+ zgj; (3.8)

the iteration starts at w2n = W(z), z2n = Z(z) and terminates at w−1, z−1. Moreover, based onequations (3.5) and (3.6), we easily establish

w2j = 2lj

+ O(

1z

), z2j = 2

l2j+ O

(1z

), w2j−1 = z2j−1 = O

(1z

), as z → ∞, (3.9)

which implies that the quantities {lj, gj} are determined in each step from the large z asymptoticsof terms known from the previous step. Indeed, if we denote by a(m) the coefficient of z−m in theexpansion of a holomorphic function a(z) at z = ∞, then we obtain the recovery formulae

lj = 2

w(0)2j

, gj = − 1

z(1)2j−1

. (3.10)

Thus, we proved that given a pair of Weyl functions W(z), Z(z) obtained from a cubic stringproblem (3.4) with a finite signed measure g, there exists a unique solution to the recurrencerelations (3.7) subject to (3.9) and thus a unique cubic string corresponding to W(z), Z(z). �

We are now ready to state the main theorem of this section.

Theorem 3.5. Let {xj(t), mj(t)}, j = 1, . . . , n be the positions and masses of the peakon ansatz (2.5)corresponding to an arbitrary signed measure m = 2

∑nj=1 mjδxj , satisfying peakon equations (2.6) on the

time interval (0, tc) and suppose that a collision occurs at tc. Then, the positions x1(t) . . . , xn(t) are analyticfunctions at tc, whereas the masses m1(t) · · · mn(t) are given by meromorphic functions at tc.

Proof. Given the initial conditions {x1(0) < x2(0) < · · · < xn(0)} and {m1(0), m2(0), . . . , mn(0)}, weset up the string problem (3.2a) after mapping m(0) to g(0). This produces the Weyl functionsW(0), Z(0), which under the peakon flow evolve as entire functions of time in view of corollary 3.3.We then set up the recursive scheme (3.7) with W(t), Z(t) as inputs. At each stage of recursion, onlyrational operations are involved, and because the recursion is finite, the formulae (3.10) result infunctions meromorphic in t. Thus, all gj, yj are meromorphic in t. For t < tc, all distances lj > 0and at tc some li vanishes but all lj remain finite, because this is a finite string. Hence, lj(t) is

on April 6, 2018http://rspa.royalsocietypublishing.org/Downloaded from

Page 7: Colliding peakons and the formation of shocks in the Degasperis

7

rspa.royalsocietypublishing.orgProcRSocA469:20130379

..................................................

regular at tc hence analytic there. For a signed measure g, there are no restrictions on individualgj, so, in general, gj remains meromorphic at tc. Mapping back to the real axis is afforded byy = tanh(x/2); hence, positions of individual masses are given by xj = ln((yj + 1)/(yj − 1)). Theonly singular points of this map are for yj = ±1 which means the end of the string or, aftermapping the problem back to the real axis, ±∞. However, based on results in reference [12], noneof the masses can escape to ±∞ in finite time. So, ((yj + 1)/(yj − 1) is in the domain of analyticityof ln and, hence, the xjs are analytic at tc. The relation between the measures m and g appearingin equations (2.3) and (3.2a) is given by mj = ((1 − y2

j )2/8)gj which implies the claim, because gj ismeromorphic and yj analytic. �

Theorem 3.5 establishes that the only singular points of solutions to the peakon ODE system(2.6a) and (2.6b) are poles. Because the inverse problem argument is valid for a fixed orderingx1 < x2 < · · · < xn of masses, the analytic continuation of masses and positions into the complexdomain in t will satisfy equations (2.6a) and (2.6b) in which sgn(xk − xi), e−|(xk−xi)| are replacedwith sgn(k − i), e−sgn(k−i)(xk−xi), respectively, to be consistent with the original ordering. It is forthese equations that we note the absence of movable critical points also known as Painlevé property[24,25]. To facilitate the statement of the theorem 3.5. of this section, we set Xi = exi , 1 ≤ i ≤ n andrewrite the system (2.6a) and (2.6b) in new variables {mi, Xi}.

Theorem 3.6 (Painlevé property). The system of differential equations

Xk = Xk

n∑i=1

mi

(Xi

Xk

)sgn(k−i), mk = 2mk

n∑i=1

mi sgn(k − i)(

Xi

Xk

)sgn(k−i)

has the Painlevé property.

Proof. First, we observe (using the variables appearing in the proof of theorem 3.5) thatXi = ((yi + 1)/(yi − 1)), hence Xi are meromorphic in t because so are yi. The formulae for Xi andmi obtained from the inverse problem are meromorphic in t in the complex plane C and dependon 2n constants (spectral data consisting, in the generic case, of n positions of poles and n residuesof the Weyl function W), which for the cubic string problem, in view of the ordering condition, areconfined to an open set in R2n by continuity of the inverse spectral map. Relaxing that conditionresults in a solution depending on 2n arbitrary constants that comprises a general solution whichis meromorphic in t in the whole complex plane C. �

Remark 3.7. For an excellent tutorial on the Painlevé property and its ties to integrability, seeHone [26], where interested readers can also find an extensive collection of literature on thesubject. In addition, it is worth mentioning that one of the topics discussed in Hone [26] is theweak Painlevé property of the full DP equation. This by no means contradicts theorem 3.6, whichrefers only to that specific system of ODEs, depending on the specific choice of variables.

In the remainder of the paper, we will apply theorem 3.6 to investigate the singularitystructure arising at the time of collisions of peakons. We point out that our analysis addressesthe singularity structure of the peakon ansatz u(x, t) in the t space. For analysis of singularityformation in the space variable x for more general initial data, the reader is referred to Cocliteet al. [27].

4. Blow-up behaviourWe now proceed to establish several theorems on peakon collisions for DP equation. To beginwith, we recall the definition of a peakon collision briefly discussed in the Introduction. We call tc

the collision time if there exists some i such that

limt→t−c

xi(t) = limt→t−c

xi+1(t), (4.1)

on April 6, 2018http://rspa.royalsocietypublishing.org/Downloaded from

Page 8: Colliding peakons and the formation of shocks in the Degasperis

8

rspa.royalsocietypublishing.orgProcRSocA469:20130379

..................................................

where xi(t)s are the position functions in the ansatz (2.5). Equivalently, we say that the ith peakoncollides with the (i + 1)th peakon at the time tc. If there exist more than two position functionsbeing identical at tc, then we will say that a multiple collision happens at tc.

In this section, we describe the behaviour of the peakon dynamical system (2.6) in theneighbourhood of a collision time tc.

To this end, we need to study a special skew-symmetric n × n real matrix An given by

An = [sgn(i − j)aij] (4.2)

whose entries satisfy aij = aji and

aij = ailalj = 0, for all 0 < i < l < j < n. (4.3)

The following propositions hold for such a matrix.

Lemma 4.1. There exists a matrix P with det P = 1 such that PTAnP = Bn, where

Bn =

⎡⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎣

0 −a12a12 0

0 −a34a34 0

. . .0 −an−1 n

an−1 n 0

⎤⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎦

, if n is even,

or

Bn =

⎡⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎣

00 −a23

a23 00 −a45

a45 0. . .

0 −an−1 n

an−1 n 0

⎤⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎦

, if n is odd.

Proof. The conclusion is trivial for n = 1, 2. We assume the conclusion to hold for n − 2; to showthat it holds for n we divide An into four block submatrices

An =[

An−2 −BBT C

], (4.4)

where C =[

0 −an−1 nan−1 n 0

]. Let us set P1 =

[In−2 0

−C−1BT I2

], then a direct computation shows that

PT1 AnP1 =

[An−2 − BC−1BT 0

0 C

].

In view of condition (4.3), B can be written as (a, an−1 na), where a = (a1 n−1, a2 n−1, . . . , an−2 n−1)T.It is now elementary to verify that BC−1BT = 0. By the induction hypothesis, there exists a matrixP2 with det P2 = 1 such that PT

2 An−2P2 = Bn−2, hence if we set

P = P1

[P2 00 I2

],

the conclusion follows. �

Corollary 4.2. If n = 2k, then det A2k =∏ki=1 a2

2i−1 2i > 0. If n = 2k + 1, then the rank of A2k+1 is 2k.

Lemma 4.3. Let E = (1, 1, . . . , 1), n = 2k + 1 and all entries satisfy 0 < aij ≤ 1. Then, the rank of the

matrix[

EA2k+1

]is 2k + 1.

on April 6, 2018http://rspa.royalsocietypublishing.org/Downloaded from

Page 9: Colliding peakons and the formation of shocks in the Degasperis

9

rspa.royalsocietypublishing.orgProcRSocA469:20130379

..................................................

Proof. Let n be any odd number. It suffices to show that the determinant of the matrix

An =

⎡⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎣

1 1 1 · · · · · · · · · 1a12 0 −a23 −a24 · · · · · · −a2 n

a13 a23 0 −a34 · · · · · · −a3 n... a34

. . .. . .

......

. . .. . . −an−2 n−1 −an−2 n

... an−2 n−1 0 −an−1 n

a1 n a2 n a3 n · · · · · · an−1 n 0

⎤⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎦

is positive. For n = 3, direct computation shows that

det A3 = a13(1 − a23) + a223 > 0.

We assume now that the conclusion holds for n − 2. We will show that it also holds for n. First,we divide An into four submatrices by

An =[

An−2 −B

BT C

],

where

C =[

0 −an−1 n

an−1 n 0

], B =

[a1 n−1 a1 n

b an−1 nb

], B =

[−1 −1b an−1 nb

],

and b = (a2 n−1, a3 n−1, . . . , an−2 n−1)T. Because C is invertible, we can factor An into the product ofupper and lower block triangular matrices as follows:

[An−2 −B

BT C

]=[

An−2 + BC−1BT −BC−1

0 I2

][In−2 0

BT C

].

Hence, det An = det(An−2 + BC−1BT) det C = a2n−1 n det(An−2 + BC−1BT).

Direct computation shows that all the entries of BC−1BT vanish except the first row that equals(a−1

n−1 n − 1)(a1 n−1, a2 n−1, . . . , an−2 n−1), therefore

det(An−2 + BC−1BT)

= det An−2 + (a−1n−1 n − 1)

∣∣∣∣∣∣∣∣∣∣∣∣

a1 n−1 a2 n−1 · · · · · · · · · an−2 n−1a12 0 −a23 · · · · · · −a2 n−2a13 a23 0 −a34 · · · −a3 n−2

......

a1 n−2 a2 n−2 · · · · · · an−3 n−2 0

∣∣∣∣∣∣∣∣∣∣∣∣

= det An−2 + (a−1n−1 n − 1)

∣∣∣∣∣∣∣∣∣∣∣∣

a12 0 −a23 · · · · · · −a2 n−2a13 a23 0 −a34 · · · −a3 n−2

......

a1 n−2 a2 n−2 · · · · · · an−3 n−2 0a1 n−1 a2 n−1 · · · · · · an−3 n−1 an−2 n−1

∣∣∣∣∣∣∣∣∣∣∣∣,

on April 6, 2018http://rspa.royalsocietypublishing.org/Downloaded from

Page 10: Colliding peakons and the formation of shocks in the Degasperis

10

rspa.royalsocietypublishing.orgProcRSocA469:20130379

..................................................

where we used that n is odd. Finally, in view of equation (4.3), we can replace the matrixin the second determinant by an upper triangular matrix by performing appropriate columnadditions, obtaining

det An

a2n−1,n

= det An−2 + (a−1n−1 n − 1)

∣∣∣∣∣∣∣∣∣∣∣∣∣∣

a12 *a23

0.. .

an−2 n−1

∣∣∣∣∣∣∣∣∣∣∣∣∣∣= det An−2 + (a−1

n−1 n − 1)a12a23 · · · an−2 n−1 > 0. �

By using the lemmas 4.1 and 4.3, we can obtain the property of mi(t) at the time of blow-up.

Theorem 4.4. If mi blows up at some t0, then mi has a pole of order 1 at t0.

Proof. Because mis are meromorphic in t we can assume that the leading term in the Laurentseries of mi around t0 is Ci/(t − t0)αi , Ci = 0. If the conclusion does not hold, then

α = maxi

{αi} ≥ 2.

Set S = {ij : αij = α} = {i1, . . . , ik}, where i1 < i2 < · · · < ik and k is at least 2 by virtue of lemma 2.1with p = 1. Comparing the leading term of both sides of (2.6b) with ij ∈ S, one can see the leadingterm on the left-hand side is −αCij/(t − t0)α+1, whereas the leading term on the right-hand side is

2Cij

(t − t0)2α

k∑l=1

sgn(ij − il) e−|xij −xil |Cil .

Because 2α > α + 1, the coefficient of (t − t0)−2α must be zero, which leads to a homogeneouslinear equation AkC = 0, where Ak = (sgn(j − l)ajl) is a k × k skew–symmetric matrix with

ajl = exij −xil (1 < j < l) and C = (Ci1 , . . . , Cik )T. Additionally, one can also find

Ci1 + Ci2 + · · · + Cik = 0

by comparing the leading term in M1. It is clear that Ak satisfies (4.3) and the condition inlemma 4.3. Hence, C must be zero according to corollary 4.2 and lemma 4.3, which leads to acontradiction. �

In the proof above, we use only the mi(t)s that are meromorphic. However, we can get strongerconclusions if we also take into account that the xi(t)s are holomorphic.

Theorem 4.5. If mj1 , . . . , mjk (1 ≤ j1 < · · · < jk ≤ n) blow up at tc and all other mi remain bounded, thenthe following conclusions hold.

(1) k must be even.(2) tc must be a collision time. Moreover, for all odd l such that 1 ≤ l < k, the peakon with label jl must

collide with the peakon with label jl+1.(3) The leading term of mjs (t) in the Laurent series around tc must have the form (−1)s/2(t − tc) for

all 1 ≤ s ≤ k.

Proof. Assume that mj1 , mj2 , . . . , mjk blow up at tc. Because M1 = m1 + m2 + · · · + mn isconserved, k is at least 2. Moreover, by theorem 4.4, the leading term in each mij ’s Laurent serieshas the form Cj/(t − tc). Hence, by equations (2.6b), the coefficients C = (C1, . . . , Ck)T satisfy thelinear equations

AkC = − 12 (1, . . . , 1)T, (4.5)

on April 6, 2018http://rspa.royalsocietypublishing.org/Downloaded from

Page 11: Colliding peakons and the formation of shocks in the Degasperis

11

rspa.royalsocietypublishing.orgProcRSocA469:20130379

..................................................

where the matrix Ak = (A(k)lm )1≤l,m≤k = (sgn(l − m) e−|xjm −xjl |)1≤l,m≤k satisfies (4.2) and (4.3).

Likewise, comparing the leading terms of both sides of (2.6a) with the subscript js (1 ≤ s ≤ k),one finds

B(k)C = 0, (4.6)

where the matrix B(k) = (B(k)lm )1≤l,m≤k = (e−|xjm −xjl |)1≤l,m≤k. Now, we prove that the theorem holds

for k = 2. In this case, (4.5) and (4.6) reduce to[0 −a12

a12 0

][C1C2

]= −1

2

[11

],

[1 a12

a12 1

][C1C2

]= 0,

where a12 = exj1 −xj2 . Direct computation shows that the solution of the equations exists iff a12 = 1,and the solution is C2 = −C1 = 1

2 . Because a12 = 1 is equivalent to xj1 (tc) = xj2 (tc), we conclude thatthe peakon with label j1 collides at tc with the peakon with label j2. Suppose now the conclusionsare valid for k − 2. We will show that they also hold for k. Let us use the same block decompositionas in equation (4.4), obtaining

Ak =

⎡⎢⎣

Ak−2 −a −ak−1 ka

aT 0 −ak−1 kak−1 kaT ak−1 k 0

⎤⎥⎦ ,

and

B(k) =

⎡⎢⎣ B(k−2) a ak−1 ka

aT 1 ak−1 kak−1 kaT ak−1 k 1

⎤⎥⎦ ,

where a = (a1 k−1, a2 k−1, . . . , ak−2 k−1)T. Let us now combine the last two rows of (4.5) and (4.6),writing them collectively as

⎡⎢⎢⎢⎣

aT 0 −ak−1 kak−1 kaT ak−1 k 0

aT 1 ak−1 kak−1 kaT ak−1 k 1

⎤⎥⎥⎥⎦C =

⎡⎢⎢⎢⎣

− 12

− 12

00

⎤⎥⎥⎥⎦ .

The latter expression can subsequently be easily reduced to

⎡⎢⎢⎢⎣

aT 0 −ak−1 k0 ak−1 k a2

k−1 k0 1 2ak−1 k0 0 1

⎤⎥⎥⎥⎦

⎡⎢⎢⎢⎢⎢⎢⎣

C1...

Ck−2Ck−1Ck

⎤⎥⎥⎥⎥⎥⎥⎦

= 12

⎡⎢⎢⎢⎣

−1−(1 − ak−1 k)

11

⎤⎥⎥⎥⎦ ,

which implies the condition for the existence of the solution to be a2k−1 k = 1, hence ak−1 k = 1. The

latter condition indicates the collision of mjk with mjk−1 . Furthermore, the solution for the lasttwo components of C is then Ck−1 = − 1

2 and Ck = 12 , which proves the sign statement for the last

two components. Substituting ak−1 k = 1, Ck−1 = − 12 , Ck = 1

2 into (4.5) and (4.6) and denoting thefirst k − 2 components of C by C, we obtain the following equations:

Ak−2C = − 12 (1, . . . , 1)T, B(k−2)C = 0, aTC = 0.

The first two equations hold by the induction hypothesis. To show that the third equation holdsautomatically if the induction hypothesis is satisfied, we observe that as the result of collisions(j1th mass collides with j2th mass, etc.) aT = (a1 k−1, a1 k−1, a3 k−1, a3 k−1, . . . , ak−3 k−1, ak−3 k−1);hence, indeed, the last equation follows from the induction hypothesis. �

on April 6, 2018http://rspa.royalsocietypublishing.org/Downloaded from

Page 12: Colliding peakons and the formation of shocks in the Degasperis

12

rspa.royalsocietypublishing.orgProcRSocA469:20130379

..................................................

The following amplification of item 2 in the above theorem is automatic.

Corollary 4.6. Suppose mj1 , . . . , mjk (1 ≤ j1 < · · · < jk ≤ n) blow up at tc and all other mi remainbounded. Then, for all odd l such that 1 ≤ l < k, the peakon with label jl must collide at tc with the peakonwith label jl + 1 (its neighbour to the right).

So far, we have established that when the masses become unbounded the collisions must occur.The converse also turns out to be valid.

Theorem 4.7. For all initial conditions for which Mn = 0, the following properties are valid:

(1) If the ith peakon collides with another peakon/peakons at tc, mi(t) must blow up at tc.(2) For all i, mi(t) cannot change its sign. In particular, mi(t) = 0 for t < tc.(3) There are no multiple collisions.(4) The distance between colliding peakons has a simple zero at tc.

Proof. Without loss of generality, we can suppose that the labelling is chosen so thatx1(tc) = · · · = xk1 (tc) < xk1+1(tc) = xk1+2(tc) = · · · = xk2 (tc) < · · · < xkl (tc) for all colliding peakons atdistinct positions xk1 (tc) < xk2 (tc) < · · · < xkl (tc). Let us now denote the set indexing all collidingpeakons by I. Because Mn = m1m2(1 − ex1−x2 )2 · · · mn−1mn(1 − exn−1−xn )2 is conserved and non-zero, it is clear that some of the masses must become unbounded. Let us denote the set of labels ofthose masses which blow up at tc by J. By theorem 4.4, any such mass corresponds to a collidingpeakon; thus, J ⊂ I. Moreover, any such a mass has a simple pole at tc. On the other hand, for eachpair of colliding peakons with adjacent indices j and j + 1, tc is a zero of (1 − exj−xj+1 )2 of orderbounded from below by 2. Thus, the order of the zero of all such exponential factors appearingin Mn is bounded from below by 2([k1 − 1] + [k2 − 1] + · · · [kl − 1]) = 2(|I| − l), where |I| denotesthe cardinality of I. Hence, because all unbounded masses have poles of order 1, |J| ≥ 2(|I| − l) toensure that Mn = 0. The maximum of l occurs when the masses collide in pairs, hence l ≤ |I|/2 andthus |J| ≥ 2(|I| − |I|/2) = |I|. This proves that J = I because J ⊂ I and thus (1) is proven. To prove(3), we return to the inequality above which now reads |I| ≥ 2(|I| − l), implying l ≥ |I|/2. Becausethe right-hand side is the maximum of l, l = |I|/2 follows, which, in turn, implies that all collisionsoccur in pairs; hence we infer the absence of multiple collisions. To prove (4), we note that for Mn

to remain bounded the order of the zero of all exponential factors has to be exactly |I| = 2(|I|/2);hence each factor (1 − exj−xj+1 )2 has a zero of order exactly equal 2.

This concluded the proof of (1), (3) and (4). In order to prove (2), we suppose that for some i,mi(t) changes its sign, then there exists some t0 for which mi(t0) = 0 while all mj remain boundedbecause t0 < tc. Hence

Mn = limt→t0

Mn = limt→t0

m1m2 · · · mn(1 − ex1−x2 )2 · · · (1 − exn−1−xn )2 = 0.

This contradicts Mn = 0. �

Remark 4.8. It is now not difficult to verify that the constants of motion M1, . . . , Mn can beextended up until the collision time tc by using lemma 2.1, followed by theorems 4.5 and 4.7.

Corollary 4.9. If mj and mj+1 collide at tc > 0, then mj > 0 and mj+1 < 0 before the collision.

Proof. Because collisions occur only in pairs, the leading terms in mj and mj+1’s Laurent seriesmust be ∓1/2(t − tc), respectively. This implies

limt→t−c

mj = +∞, limt→t−c

mj+1 = −∞.

The conclusion holds because in view of theorem 4.7 mj and mj+1 cannot change their signs. �

The following proposition shows that the simultaneous collisions (several peakon–antipeakonpairs collide at distinct locations at the common time tc) can happen. We indicate below howcertain symmetric initial conditions will lead to simultaneous collisions. To this end, we considerequations (2.6) for n = 4 and a special choice of initial conditions.

on April 6, 2018http://rspa.royalsocietypublishing.org/Downloaded from

Page 13: Colliding peakons and the formation of shocks in the Degasperis

13

rspa.royalsocietypublishing.orgProcRSocA469:20130379

..................................................

time

posi

tion

time

0

1

2

3

4

5

6

7

8

0.1 0.2 0.3 0.4−60

−40

−20

20

40

60

0 0.1 0.2 0.3 0.4

mas

s

0

Figure 1. Two symmetric peakon–antipeakon pairs with massesm1(0)= 10= −m4(0),m2(0)= −1= −m3(0) undergo asimultaneous collision. (Online version in colour.)

Lemma 4.10. If the initial conditions satisfy

m1(0) = −m4(0) > 0, m2(0) = −m3(0) < 0, (4.7a)

and

x1(0) − x2(0) = x3(0) − x4(0) < 0, (4.7b)

then m1(t) = −m4(t), m2(t) = −m3(t), x1(t) − x2(t) = x3(t) − x4(t) will hold for all 0 < t < tc.

Proof. Consider the following ODEs

x1 = m1 + m2 ex1−x2 − m2 ex1−x3 − m1 e2x1−x2−x3 ,

x2 = m1 ex1−x2 + m2 − m2 ex2−x3 − m1 ex1−x3 ,

x3 = m1 ex1−x3 + m2 ex2−x3 − m2 − m1 ex1−x2 ,

m1 = −2m1(−m2 ex1−x2 + m2 ex1−x3 + m1 e2x1−x2−x3 ),

and m2 = −2m2(m1 ex1−x2 + m2 ex2−x3 + m1 ex1−x3 ),

then direct computation shows that {x1, x2, x3, x2 + x3 − x1, m1, m2, −m2, −m1} satisfy the systemof ODEs (2.6) for n = 4. �

The following is then immediate (figure 1).

Corollary 4.11. If the initial conditions (4.7) hold and the peakon–antipeakon pair (m1, m2) collides attc, then so does (m3, m4) and vice versa.

5. Collisions and shocksIn this section, we investigate the behaviour of m and u at the time of collision(s). We start withm and observe that because the collision of peakons occurs in pairs it is sufficient to study a fixedcolliding pair mj, mj+1.

on April 6, 2018http://rspa.royalsocietypublishing.org/Downloaded from

Page 14: Colliding peakons and the formation of shocks in the Degasperis

14

rspa.royalsocietypublishing.orgProcRSocA469:20130379

..................................................

Theorem 5.1. If mj collides with mj+1 at time tc > 0 and the position xc, then

limt→t−c

(mj(t)δ(x − xj(t)) + mj+1(t)δ(x − xj+1(t)))

=(

limt→t−c

(mj(t) + mj+1(t))

)δ(x − xc) + 1

2(xj(tc) − xj+1(tc))δ′(x − xc)

=(

limt→t−c

(mj(t) + mj+1(t))

)δ(x − xc) + 1

2

(lim

t→t−c(u(xj(t), t) − u(xj+1(t), t))

)δ′(x − xc)

in D ′(R).

Proof. For an arbitrary ϕ(x) ∈ D(R),

〈mj(t)δ(x − xj(t)) + mj+1(t)δ(x − xj+1(t)), ϕ(x)〉 = mj(t)ϕ(xj(t)) + mj+1(t)ϕ(xj+1(t)).

Using corollary 4.9, we can write

mj(t) = − 12(t − tc)

+ C0 + O(t − tc), mj+1(t) = 12(t − tc)

+ C0 + O(t − tc)

around tc. Hence

limt→t−c

〈mj(t)δ(x − xj(t)) + mj+i(t)δ(x − xj+1(t)), ϕ(x)〉

= (C0 + C0)ϕ(xc) − limt→tc

ϕ(xj(t)) − ϕ(xj+1(t))

2(t − tc)

=(

limt→t−c

(mj((t) + mj+1(t))

)ϕ(xc) − 1

2

(lim

t→t−c(xj − xj+1)

)ϕ′(xc)

=(

limt→t−c

(mj(t) + mj+1(t))

)ϕ(xc) − 1

2

(lim

t→t−c(u(xj(t), t) − u(xj+1(t), t))

)ϕ′(xc),

where in the last step we used equation (2.6a). The conclusion now follows from the definition ofδ and δ′. �

Because m = u − uxx, we have the immediate corollary.

Corollary 5.2. Suppose m(0) = 2∑n

k=1 mk(0)δ(x − xk(0)) is a multi-peakon at t = 0 for which Mn = 0and such that at tc one, or several of its peakon–antipeakon pairs collide. For any colliding pair k, k + 1, letus denote limt→t−c u(xk(t)) = ul(xk(tc)), limt→t−c u(xk+1(t)) = ur(xk(tc)), respectively. Then,

limt→t−c

m(t) = 2n∑

k=1

mk(tc)δ(x − xk(tc)) + 2∑

k:pairsxk ,xk+1collide

sk(tc)δ′(x − xk(tc)) in D′(R).

The shock strengths are given by

sk(tc) = ul(xk(tc)) − ur(xk(tc))2

,

and they satisfy the (strict) entropy condition ([11]) sk(tc) > 0 .

Proof. It suffices to prove the claim if there is only one colliding pair; the general case followseasily, because masses collide pairwise. For t < tc, the measure evolves as m(t) = 2

∑nk=1 mk(t)δ(x −

xk(t)), where xk(t), mk(t) satisfy equations (2.6a) and (2.6b), respectively. Suppose now that thepair j, j + 1 collides at the point xc. Then, by theorem 5.1 limt→tc m(t) = 2

∑k =j,j+1 mk(tc)δ(x −

xk(tc)) + 2 limt→t−c (mj(t) + mj+1(t))δ(x − xc) + 2 12 (xj(tc) − xj+1(tc))δ′(x − xc). To prove that sj(tc) ≥ 0,

on April 6, 2018http://rspa.royalsocietypublishing.org/Downloaded from

Page 15: Colliding peakons and the formation of shocks in the Degasperis

15

rspa.royalsocietypublishing.orgProcRSocA469:20130379

..................................................

we write sj(tc) = 12 limt→t−c (xj(t) − xj+1(t)) = 1

2 limt→t−c (u(xj(t), t) − u(xj+1(t), t)) and observe

xj(tc) − xj+1(tc) = limt→tc

xj+1(t) − xj(t)

tc − t

which implies the entropy condition sj(tc) ≥ 0 in view of the ordering assumption xj(t) < xj+1(t).The strict inequality follows from item (4) in theorem 4.7. �

The following amplification of the previous theorem brings the issues of the wave breakdownand a shock creation sharply into focus. To put our result into the proper perspective, we firstbriefly review the well-posedness result for L1(R) ∩ BV(R) proved by Coclite & Karlsen [13,section 3]. We state only the core result pertinent to our paper, leaving other aspects, includingthe precise definition of the entropy condition, required for the stability and uniqueness of the DPequation, to an interested reader.

Theorem 5.3 (Coclite–Karlsen). Let u0 ∈ L1(R) ∩ BV(R). Then, there exists a unique entropy weaksolution to the Cauchy problem u|t=0 = u0 for the DP equation (2.1).

It is then proven by Lundmark [11] that the shockpeakon ansatz

u(x, t) =n∑

j=1

{mj(t) − sj(t) sgn(x − xj(t))} e−|x−xj(t)|,

is an entropy-weak solution provided the shock strengths sj ≥ 0. This sets the stage for thenext theorem. First, we will equip the space L1(R) ∩ BV(R) with the norm ‖f‖BV = ‖f‖L1 + Vf (R),where, in general, Vf (U) denotes the variation of f over an open subset U ⊂ R.

Theorem 5.4. Given a multi-peakon solution u(x, t) defined on Rn × [0, tc), then

(1) u(·, t) ∈ L1(R) ∩ BV(R) for all 0 ≤ t < tc,(2) u(·, t) converges in ‖ · ‖L1 to the shockpeakon

u(x, tc) =n∑

i=1

mi(tc) e−|x−xi(tc)| +n∑

i=1

Cixi(tc) sgn(x − xi(tc)) e−|x−xi(tc)|,

u(·, tc) ∈ L1(R) ∩ BV(R), where the Laurent expansion of mi(t) around tc is written as

mi(t) = Ci

t − tc+

∞∑l=0

al(t − tc)l def= Ci

t − tc+ mi(t),

with the proviso that Ci = 0 if the ith mass is not involved in a collision and Ci = − 12 for a colliding

peakon, Ci = 12 for a colliding antipeakon, respectively.

(3) Vu(·,t)(R) → Vu(·,tc)(R) as t → t−c .

Proof. We start with the case n = 2. Then, u(x, t) = m1(t) e−|x−x1(t)| + m2(t) e−|x−x2(t)| and x1(tc) =x2(tc) = xc. According to theorem 4.5, we have

m1(t) = − 12(t − tc)

+ m1(t), m2(t) = 12(t − tc)

+ m2(t),

where m1(t), m2(t) are analytic around tc. It is clear that

m1(t) e−|x−x1(t)| + m2(t) e−|x−x2(t)| ∈ L1(R) ∩ BV(R) for all 0 ≤ t ≤ tc.

on April 6, 2018http://rspa.royalsocietypublishing.org/Downloaded from

Page 16: Colliding peakons and the formation of shocks in the Degasperis

16

rspa.royalsocietypublishing.orgProcRSocA469:20130379

..................................................

By the mean value theorem, we find that

v(x, t) def= 12(t − tc)

(e−|x−x2(t)| − e−|x−x1(t)|)

=

⎧⎪⎪⎪⎪⎪⎪⎨⎪⎪⎪⎪⎪⎪⎩

12

[x1(s) ex−x1(s) − x2(s) ex−x2(s)]|s=t+θ1(tc−t), x < x1(t) < x2(t),

12

[x2(s) ex2(s)−x − x1(s) ex1(s)−x]|s=t+θ2(tc−t), x1(t) < x2(t) < x,

ex−xc − exc−x

2(t − tc)− 1

2(x1(s) ex1(s)−x + x2(s) ex−x2(s))|s=t+θ3(tc−t), x1(t) < x < x2(t),

where 0 < θj < 1, j = 1, 2, 3. Hence, we have the pointwise limit

limt→t−c

v(x, t) =

⎧⎪⎪⎨⎪⎪⎩

sgn(x − xc)(

−12

x1(tc) + 12

x2(tc))

e−|x−xc|, for x = xc,

− x1(tc) + x2(tc)2

, for x = xc.

Let us define

v(x, tc) =

⎧⎪⎨⎪⎩

sgn(x − xc)(

−12

x1(tc) + 12

x2(tc))

e−|x−xc|, for x = xc,

0, for x = xc.

and consider the integral

∫+∞

−∞|v(x, t) − v(x, tc)| dx =

(∫ x1(t)

−∞+

∫ x2(t)

x1(t)+

∫+∞

x2(t)

)|v(x, t) − v(x, tc)| dx.

Then, the first and the last term of the right-hand side converge to zero as t → t−c owing toLebesgue’s dominated convergence theorem.

Observe that the second term satisfies∫ x2(t)

x1(t)|v(x, t) − v(x, tc)| dx ≤

∫ x2(t)

x1(t)|v(x, t)| dx +

∫ x2(t)

x1(t)|v(x, tc)| dx

≤∫ x2(t)

x1(t)|v(x, t)| dx +

∫ x2(t)

x1(t)

|ex−x2(t) − exc−x|2(tc − t)

dx

+ 12

∫ x2(t)

x1(t)|x1(s)|x1(s)−xdx + 1

2

∫ x2(t)

x1(t)|x2(s)| ex−x2(s)dx

=∫ x2(t)

x1(t)|v(x, t)| dx + x2(t) − x1(t)

2(tc − t)| ey−xc − exc−y|

+ 12

∫ x2(t)

x1(t)|x1(s)| ex1(s)−xdx + 1

2

∫ x2(t)

x1(t)|x2(s)| ex−x2(s)dx,

where s ∈ (t, tc) and y ∈ (x1(t), x2(t)). Because |x1(s)| ex1(s)−x and |x2(s)| ex−x2(s) are bounded, and

x2(t) − x1(t) → 0,x2(t) − x1(t)

2(tc − t)→ x1(tc) − x2(tc), | ey−xc − exc−y| → 0

as t → t−c , we have that v(x, t) converges to v(x, tc) in the sense of L1, which shows that theconclusion holds for n = 2.

In general, because collisions can occur only in pairs, we can assume that mj1 (t), mj1+1(t), mj2 (t),mj2+1(t), . . . , mjk (t), mjk+1(t) blow up at tc and all the other mi remain bounded. It is clear thatmi(t) e−|x−xi(t)| lies in L1(R) ∩ BV(R) and converges to mi(tc) e−|x−xi(tc)| in L1 if mi(t) remains

on April 6, 2018http://rspa.royalsocietypublishing.org/Downloaded from

Page 17: Colliding peakons and the formation of shocks in the Degasperis

17

rspa.royalsocietypublishing.orgProcRSocA469:20130379

..................................................

bounded at tc. Meanwhile, according to the proof above, we can easily see that

mjs (t) e−|x−xjs (t)| + mjs+1(t) e−|x−xjs+1(t)| ∈ L1(R) ∩ BV(R) for all 1 ≤ s ≤ k,

whose limit is

mjs (tc) e−|x−xjs (tc)| + mjs+1(tc) e−|x−xjs+1(tc)|

− 12 sgn(x − xjs (tc))(xjs (tc) − xjs+1(tc)) e−|x−xj(tc)|

as t → t−c , which proves the L1 convergence. Let us now compute the variation Vu(·,tc). Becauseu(x, tc) is piecewise smooth, its distributional derivative, which is a Radon measure, equalsux(x, tc) = u(x, tc)cl

x +∑j[u](xj(tc))δxj(tc), where ucl

x means the classical derivative. One easily checksthat the jump [u](xi(tc)) = xi+1(tc) − xi(tc) if the peakons mi and mi+1 collided at xi(tc); otherwise,[u](xj(tc)) = 0. Thus, Vu(·,tc) = ‖ucl

x ‖L1 +∑j |[u](xj(tc)|. We will now compute the total variation

Vu(·,t). For ease of notation, we will write u instead of u(·, t) in the remainder of the proof.Let us denote x0 = −∞, xn+1 = ∞, while, as above, xj = xj(t), 1 ≤ j ≤ n, denotes the positions of

peakons. We then write the total variation Vu(R) =∑ni=0 Vu(xi, xi+1). We will now compute the

limit t → t−c for each Vu(xi, xi+1). There are three cases to consider:

(1) Vu(x0, x1) and Vu(xn, xn+1),(2) Vu(xi, xi+1) when mi is not colliding with mi+1, and(3) Vu(xi, xi+1) for a colliding peakon–antipeakon pair.

In all the three cases, as long as t < tc, the distributional derivative ux is the same as uclx and we

will drop the superscript from the notation; moreover, ux is continuous on (xi, xi+1) and boundedon [xi, xi+1], hence implying Vu(xi, xi+1) = ∫xi+1

xi|ux(x, t)| dx. From the peakon ansatz, (2.5), ux =

±u on (x0, x1) and (xn, xn+1), whereas ux = −∑ij=1 mj exj−x +∑n

j=i+1 mj ex−xj := −(M+)i1 e−x +

(M−)ni+1 ex on (xi, xi+1). Thus, in the first case, Vu(x0, x1) = ∫x1

x0|u(x, t)| dx and an analogous result

holds for Vu(xn, xn+1). Because u(x, t) is exponentially bounded in x and continuous in t, we cantake the limit obtaining limt→t−c Vu(·,t)(x0, x1) = Vu(·,tc)(x0, x1(tc)).

Now, consider the interval (xi, xi+1) for the non-colliding peakon pair mi and mi+1. Thecolliding peakon pairs are either in (M+)i

1 or (M−)ni+1, both of which remain bounded at

tc, because for colliding peakons expressions mj e±xj + mj+1 e±xj+1 have finite limits at tc,implying that limt→t−c |ux| = |ux(x, tc)| which is continuous on [xj(tc), xj+1(tc)], implying that |ux| isuniformly bounded, hence by Lebesgue’s dominated convergence theorem limt→t−c Vu(xi, xi+1) =Vu(·,tc)(xi, xi+1(tc)).

Finally, in the third case, when mi collides with mi+1 and the interval under consideration is[xi, xi+1], we can split ux = −(M+)i−1

1 e−x − mi exi e−x + mi+1 e−xi+1 ex + (M−)ni+2 ex. Because there

are no triple collisions, all the remaining colliding peakon pairs are in (M+)i−11 or (M−)n

i+2, whichare bounded, and hence these terms do not contribute in the limit t → tc for which xi(tc) =xi+1(tc). Thus, limt→tc

∫xi+1xi

|ux| dx = limt→tc

∫xi+1xi

| − mi exi e−x + mi+1 e−xi+1 ex| dx = limt→tc (mi −mi+1)(1 − exi−xi+1 ). The last limit can be computed explicitly using theorem 4.5 giving thefinal result

limt→tc

Vu(xi, xi+1) = xi(tc) − xi+1(tc) < ∞ = |[u](xi(tc))|.

Thus, by taking the limit t → t−c in the sum∑n

i=0 Vu(xi, xi+1), we obtain Vu(·,tc)(R), which concludesthe proof. �

Remark 5.5. Even though we showed the L1 convergence of a multi-peakon u(x, t) ∈ L1(R) ∩BV(R) to a shockpeakon u(x, tc) which is also in L1(R) ∩ BV(R), as well as the convergence ofrespective variations, this does not imply the convergence in the BV topology. Indeed, considerany point x0 for which u(x0, t) → u(x0, tc) and take another arbitrary point x. Denote u(t, x) −u(tc, x) = f (t, x). Then, |f (t, x)| ≤ |f (t, x0)| + |f (x, t) − f (x0, t)| ≤ |f (t, x0)| + Vf (·,t). Because f (t, x0) goesto 0 as t → t−c , lim Vf (·,t) = 0 would imply not only the pointwise convergence of f (t, x) to 0,

on April 6, 2018http://rspa.royalsocietypublishing.org/Downloaded from

Page 18: Colliding peakons and the formation of shocks in the Degasperis

18

rspa.royalsocietypublishing.orgProcRSocA469:20130379

..................................................

hence u(x, t) → u(x, tc), but also the uniform convergence on any compact set, thus contradictingthe existence of discontinuities at the points of collisions of peakons. This is an instance ofthe L1 topology providing more appropriate framework for the BV functions, which ultimatelyreinforces viewing BV(R) as a subspace of L1(R); for another, similar, case the reader might consulttheorem 1 on p. 331, section 4.1, in Giaquinta et al. [28].

In the body of the proof of theorem 5.4, we identified the jump [u](xi(tc) to be equalxi+1(tc) − xi(tc) at any collision point. Thus, si from Lundmark’s shockpeakon ansatz equalssi = (xi(tc) − xi+1(tc))/2 and, as already indicated in corollary 5.2, si > 0. This allows us toconclude with.

Corollary 5.6. The L1 limit of a multi-peakon u(·, t) as t → t−c is a shockpeakon in the sense of Lundmarkand as such it admits a unique entropy weak extension past the collision point tc.

Acknowledgements. We thank Hans Lundmark for numerous perceptive comments. J.S. thanks the CentroInternacional de Ciencias (CIC) in Cuernavaca (Mexico) for hospitality and F. Calogero for making the stay soenjoyable and productive. We also thank the Department of Mathematics and Statistics of the University ofSaskatchewan for making the collaboration possible.

Funding statement. This work was supported by the National Natural Science Funds of China (NSFC11271285 toL.Z.) and the Natural Sciences and Engineering Research Council of Canada (NSERC163953 to J.S.).

References1. Degasperis A, Procesi M. 1999 Asymptotic integrability. In Symmetry and perturbation theory

(eds A Degasperis, G Gaeta), pp. 23–37. River Edge, NJ: World Scientific Publishing.2. Camassa R, Holm DD. 1993 An integrable shallow water equation with peaked solitons. Phys.

Rev. Lett. 71, 1661–1664. (doi:10.1103/PhysRevLett.71.1661)3. Johnson RS. 2002 Camassa–Holm, Korteweg–de Vries and related models for water waves.

J. Fluid Mech. 455, 63–82. (doi:10.1017/S0022112001007224)4. Constantin A, Lannes D. 2009 The hydrodynamical relevance of the Camassa–Holm and

Degasperis–Procesi equations. Arch. Ration. Mech. Anal. 192, 165–186. (doi:10.1007/s00205-008-0128-2)

5. McKean HP. 2004 Breakdown of the Camassa–Holm equation. Commun. Pure Appl. Math. 57,416–418. (doi:10.1002/cpa.20003)

6. McKean HP. 1998 Breakdown of a shallow water equation (Mikio Sato: a great Japanesemathematician of the twentieth century.). Asian J. Math. 2, 867–874.

7. Majda AJ, Bertozzi AL. 2002 Vorticity and incompressible flow. Cambridge Texts in AppliedMathematics, no. 27. Cambridge, UK: Cambridge University Press.

8. Beals R, Sattinger D, Szmigielski J. 2000 Multipeakons and the classical moment problem. Adv.Math. 154, 229–257. (doi:10.1006/aima.1999.1883)

9. Beals R, Sattinger DH, Szmigielski J. 1999 Multi-peakons and a theorem of Stieltjes. InverseProbl. 15, L1–L4. (doi:10.1088/0266-5611/15/1/001)

10. Lundmark H, Szmigielski J. 2005 Degasperis–Procesi peakons and the discrete cubic string.Int. Math. Res. Pap. 2005, 53–116. (doi:10.1155/IMRP.2005.53)

11. Lundmark H. 2007 Formation and dynamics of shock waves in the Degasperis–Procesiequation. J. Nonlinear Sci. 17, 169–198. (doi:10.1007/s00332-006-0803-3)

12. Szmigielski J, Zhou L. 2013 Peakon–antipeakon interaction in the Degasperis–Procesiequation. Contemp. Math. 593, 83–107. (doi:10.1090/conm/593/11873) (http://arxiv.org/abs/1301.0171)

13. Coclite GM, Karlsen KH. 2006 On the well-posedness of the Degasperis–Procesi equation.J. Funct. Anal. 233, 60–91. (doi:10.1016/j.jfa.2005.07.008)

14. Coclite GM, Karlsen KH. 2007 On the uniqueness of discontinuous solutions to theDegasperis–Procesi equation. J. Differ. Equ. 234, 142–160. (doi:10.1016/j.jde.2006.11.008)

15. Liu Y. 2010 Wave breaking phenomena and stability of peakons for the Degasperis–Procesiequation. In Trends in partial differential equations. Adv. Lect. Math. (ALM), no. 10, pp. 265–293.Somerville, MA: International Press.

on April 6, 2018http://rspa.royalsocietypublishing.org/Downloaded from

Page 19: Colliding peakons and the formation of shocks in the Degasperis

19

rspa.royalsocietypublishing.orgProcRSocA469:20130379

..................................................

16. Lin Z, Liu Y. 2009 Stability of peakons for the Degasperis–Procesi equation. Comm. Pure Appl.Math. 62, 125–146.

17. Liu Y, Yin Z. 2006 Global existence and blow-up phenomena for the Degasperis–Procesiequation. Commun. Math. Phys. 267, 801–820. (doi:10.1007/s00220-006-0082-5)

18. Escher J, Liu Y, Yin Z. 2006 Global weak solutions and blow-up structure for the Degasperis–Procesi equation. J. Funct. Anal. 241, 457–485. (doi:10.1016/j.jfa.2006.03.022)

19. Feng B-F, Liu Y. 2009 An operator splitting method for the Degasperis–Procesi equation.J. Comput. Phys. 228, 7805–7820. (doi:10.1016/j.jcp.2009.07.022)

20. Coclite GM, Karlsen KH, Risebro NH. 2008 Numerical schemes for computing discontinuoussolutions of the Degasperis–Procesi equation. IMA J. Numer. Anal. 28, 80–105. (doi:10.1093/imanum/drm003)

21. Hoel HA. 2007 A numerical scheme using multi-shockpeakons to compute solutions of theDegasperis–Procesi equation. Electron. J. Differ. Equ. 100, 22.

22. Degasperis A, Holm DD, Hone ANW. 2002 A new integrable equation with peakon solutions.Theor. Math Phys. 133, 1463–1474. (doi:10.1023/A:1021186408422)

23. Lundmark H, Szmigielski J. 2003 Multi-peakon solutions of the Degasperis–Procesi equation.Inverse Probl. 19, 1241–1245. (doi:10.1088/0266-5611/19/6/001)

24. Ince EL. 1944 Ordinary differential equations. New York, NY: Dover Publications.25. Conte R (ed.) 1999 The Painlevé property: one century later. CRM Series in Mathematical Physics.

New York, NY: Springer.26. Hone ANW. 2009 Painlevé tests, singularity structure and integrability. In Integrability. Lecture

Notes in Phys. no. 767, pp. 245–277. Berlin, Germany: Springer.27. Coclite GM, Gargano F, Sciacca V. 2012 Analytic solutions and singularity formation for the

Peakon b-family equations. Acta Appl. Math. 122, 419–434. (doi:10.1007/s10440-012-9753-8)28. Giaquinta M, Modica G, Soucek J. 1998 Cartesian currents. Cartesian currents in the calculus

of variations. I, vol. 37 of Ergebnisse der Mathematik und ihrer Grenzgebiete. 3. Folge. A seriesof Modern Surveys in Mathematics (Results in Mathematics and Related Areas. 3rd Series. ASeries of Modern Surveys in Mathematics). Berlin, Germany: Springer.

on April 6, 2018http://rspa.royalsocietypublishing.org/Downloaded from