Contractive Metrics for Nonsmooth Evolutions

Similar documents
Hyperbolic Systems of Conservation Laws. in One Space Dimension. II - Solutions to the Cauchy problem. Alberto Bressan

Piecewise Smooth Solutions to the Burgers-Hilbert Equation

Hyperbolic Systems of Conservation Laws

On Asymptotic Variational Wave Equations

Global conservative solutions of the Camassa-Holm equation

Hyperbolic Systems of Conservation Laws. I - Basic Concepts

Lecture Notes on Hyperbolic Conservation Laws

On a Lyapunov Functional Relating Shortening Curves and Viscous Conservation Laws

Generic Singularities of Solutions to some Nonlinear Wave Equations

Instability of Finite Difference Schemes for Hyperbolic Conservation Laws

On the Front-Tracking Algorithm

An introduction to Mathematical Theory of Control

Structurally Stable Singularities for a Nonlinear Wave Equation

On the Dependence of Euler Equations on Physical Parameters

The Scalar Conservation Law

On Discontinuous Differential Equations

Shift Differentials of Maps in BV Spaces.

On finite time BV blow-up for the p-system

Existence Theory for Hyperbolic Systems of Conservation Laws with General Flux-Functions

On the Cauchy Problems for Polymer Flooding with Gravitation

Stability of Feedback Solutions for Infinite Horizon Noncooperative Differential Games

Author(s) Huang, Feimin; Matsumura, Akitaka; Citation Osaka Journal of Mathematics. 41(1)

Hyperbolic Systems of Conservation Laws. in One Space Dimension. I - Basic concepts. Alberto Bressan. Department of Mathematics, Penn State University

Extremal Solutions of Differential Inclusions via Baire Category: a Dual Approach

Some asymptotic properties of solutions for Burgers equation in L p (R)

REGULARITY THROUGH APPROXIMATION FOR SCALAR CONSERVATION LAWS

Intersection Models and Nash Equilibria for Traffic Flow on Networks

Équation de Burgers avec particule ponctuelle

A maximum principle for optimally controlled systems of conservation laws

Global Existence of Large BV Solutions in a Model of Granular Flow

The inviscid limit to a contact discontinuity for the compressible Navier-Stokes-Fourier system using the relative entropy method

Singularity formation for compressible Euler equations

Convergence rate estimates for the gradient differential inclusion

On uniqueness of weak solutions to transport equation with non-smooth velocity field

On the Cauchy Problems for Polymer Flooding with Gravitation

Applications of the compensated compactness method on hyperbolic conservation systems

Scalar conservation laws with moving density constraints arising in traffic flow modeling

Recent Trends in Differential Inclusions

L 1 stability of conservation laws for a traffic flow model

SYMMETRY OF POSITIVE SOLUTIONS OF SOME NONLINEAR EQUATIONS. M. Grossi S. Kesavan F. Pacella M. Ramaswamy. 1. Introduction

Rarefaction wave interaction for the unsteady transonic small disturbance equations

1. Introduction. Consider a strictly hyperbolic system of n conservation laws. 1 if i = j, 0 if i j.

Periodic solutions of weakly coupled superlinear systems

M. HERTY, CH. JÖRRES, AND B. PICCOLI

SHOCK WAVES FOR RADIATIVE HYPERBOLIC ELLIPTIC SYSTEMS

arxiv: v2 [math.ap] 1 Jul 2011

A RIEMANN PROBLEM FOR THE ISENTROPIC GAS DYNAMICS EQUATIONS

c 2007 Society for Industrial and Applied Mathematics

EXISTENCE OF SOLUTIONS TO THE CAHN-HILLIARD/ALLEN-CAHN EQUATION WITH DEGENERATE MOBILITY

Non-linear Wave Propagation and Non-Equilibrium Thermodynamics - Part 3

NONCLASSICAL SHOCK WAVES OF CONSERVATION LAWS: FLUX FUNCTION HAVING TWO INFLECTION POINTS

ON WEAKLY NONLINEAR BACKWARD PARABOLIC PROBLEM

Error bounds for a Deterministic Version of the Glimm Scheme

Controlling Mechanical Systems by Active Constraints. Alberto Bressan. Department of Mathematics, Penn State University

Relaxation methods and finite element schemes for the equations of visco-elastodynamics. Chiara Simeoni

Wave breaking in the short-pulse equation

2 Statement of the problem and assumptions

Hyperbolic Conservation Laws Past and Future

COMPARISON PRINCIPLES FOR CONSTRAINED SUBHARMONICS PH.D. COURSE - SPRING 2019 UNIVERSITÀ DI MILANO

A high order adaptive finite element method for solving nonlinear hyperbolic conservation laws

BIHARMONIC WAVE MAPS INTO SPHERES

CODIMENSION-ONE RIEMANN SOLUTIONS: MISSING RAREFACTIONS IN TRANSITIONAL WAVE GROUPS

L 1 Stability for Systems of Hyperbolic Conservation Laws

Nonlinear Control Systems

Coupling conditions for transport problems on networks governed by conservation laws

Alberto Bressan. Department of Mathematics, Penn State University

ASYMPTOTIC THEORY FOR WEAKLY NON-LINEAR WAVE EQUATIONS IN SEMI-INFINITE DOMAINS

i=1 α i. Given an m-times continuously

On the Stability of the Best Reply Map for Noncooperative Differential Games

Lipschitz Metrics for a Class of Nonlinear Wave Equations

ON THE FOLIATION OF SPACE-TIME BY CONSTANT MEAN CURVATURE HYPERSURFACES

at time t, in dimension d. The index i varies in a countable set I. We call configuration the family, denoted generically by Φ: U (x i (t) x j (t))

Fractal Conservation Laws: Global Smooth Solutions and Vanishing Regularization

b i (µ, x, s) ei ϕ(x) µ s (dx) ds (2) i=1

Optimality Conditions for Solutions to Hyperbolic Balance Laws

Ray equations of a weak shock in a hyperbolic system of conservation laws in multi-dimensions

Impulsive Control of Lagrangian Systems and Locomotion in Fluids. Alberto Bressan

STRUCTURAL STABILITY OF SOLUTIONS TO THE RIEMANN PROBLEM FOR A NON-STRICTLY HYPERBOLIC SYSTEM WITH FLUX APPROXIMATION

Hyperbolic Gradient Flow: Evolution of Graphs in R n+1

A Very Brief Introduction to Conservation Laws

EXISTENCE AND REGULARITY RESULTS FOR SOME NONLINEAR PARABOLIC EQUATIONS

Takens embedding theorem for infinite-dimensional dynamical systems

CUTOFF RESOLVENT ESTIMATES AND THE SEMILINEAR SCHRÖDINGER EQUATION

Traffic models on a network of roads

Hysteresis rarefaction in the Riemann problem

EXISTENCE OF SOLUTIONS FOR KIRCHHOFF TYPE EQUATIONS WITH UNBOUNDED POTENTIAL. 1. Introduction In this article, we consider the Kirchhoff type problem

The Riemann problem. The Riemann problem Rarefaction waves and shock waves

INFINITE TIME HORIZON OPTIMAL CONTROL OF THE SEMILINEAR HEAT EQUATION

Propagating terraces and the dynamics of front-like solutions of reaction-diffusion equations on R

Global well-posedness of the primitive equations of oceanic and atmospheric dynamics

Conservative Solutions to a Nonlinear Variational Wave Equation

A TWO PARAMETERS AMBROSETTI PRODI PROBLEM*

Lecture Notes on PDEs

The 2-d isentropic compressible Euler equations may have infinitely many solutions which conserve energy

Universität des Saarlandes. Fachrichtung 6.1 Mathematik

Optima and Equilibria for Traffic Flow on Networks with Backward Propagating Queues

1 Basic Second-Order PDEs

arxiv: v1 [math.ap] 28 May 2009

Differentiability with respect to initial data for a scalar conservation law

Recent developments in elliptic partial differential equations of Monge Ampère type

Transcription:

Contractive Metrics for Nonsmooth Evolutions Alberto Bressan Department of Mathematics, Penn State University, University Park, Pa 1682, USA bressan@mathpsuedu July 22, 212 Abstract Given an evolution equation, a standard way to prove the well posedness of the Cauchy problem is to establish a Gronwall type estimate, bounding the distance between any two trajectories There are important cases, however, where such estimates cannot hold, in the usual distance determined by the Euclidean norm or by a Banach space norm In alternative, one can construct different distance functions, related to a Riemannian structure or to an optimal transportation problem This paper reviews various cases where this approach can be implemented, in connection with discontinuous ODEs on IR n, nonlinear wave equations, and systems of conservation laws For all the evolution equations considered here, a metric can be constructed such that the distance between any two solutions satisfies a Gronwall type estimate This yields the uniqueness of solutions, and estimates on their continuous dependence on the initial data 1 Introduction Consider an abstract evolution equation in a Banach space d u(t) = F (u(t)) (11) dt If F is a continuous vector field with Lipschitz constant L, the classical Cauchy-Lipschitz theory applies For any given initial data u() = ū, (12) the solution of (11) is thus unique, and depends continuously on ū between any two solutions grows at a controlled rate: d dt u(t) v(t) L u(t) v(t) (P1) In this case, the classical Gronwall s estimate yields u(t) v(t) C(t) u() v(), Indeed, the distance (P2) 1

with C(t) = e Lt Semigroup theory has etended the validity of estimates such as (P1), (P2) to a wide class of right hand sides, including differential operators, which generate a continuous flow [19, 24, 3, 35] On the other hand, there are cases (such as the Camassa-Holm equation) where the flow generated by (11) is not Lipschitz continuous wrt the initial data, in any standard Hölder or Sobolev norm In other cases, such as hyperbolic systems of conservation laws, the generated semigroup is globally Lipschitz continuous wrt the L 1 norm but does not satisfy an estimate of the form (P1), for any constant L In all these situations, a natural problem is to seek an alternative distance d (, ), possibly not equivalent to any of the usual norm distances, for which (P1) or (P2) still hold Aim of this note is to discuss a few eamples where this goal can be achieved Typically, the distance d is defined as a Riemann type distance In other words, one starts with a Banach space E and a family Σ of sufficiently regular paths γ : [, 1] E, for which some kind of weighted length γ can be defined This needs not be equivalent to the length derived from the norm distance Given two elements u, v E, one first defines } d (u, v) = inf { γ ; γ Σ, γ() = u, γ(1) = v, (13) and then takes the lower semicontinuous envelope (wrt convergence in norm): d (u, v) = lim inf u u, v v d (u, v ) (14) Besides achieving a proof of uniqueness and continuous dependence, estimates of the form (P2) are useful for establishing error estimates Indeed, adopting a semigroup notation, call t S t ū the solution to the Cauchy problem d u = F (u) u() = ū (15) dt Assume that, for any couple of initial data ū, v, there holds d (S t ū, S t v) C d (ū, v) t [, T ] (16) Then for any Lipschitz continuous trajectory t w(t) one can deduce the error estimate [3] d ( ) T d ( ) w(t + h), S h w(t) w(t ), S T w() C lim inf dt (17) h + h Here the left hand side is the distance at time T between the approimate solution w( ) and the eact solution of (11) with the same initial data w() The right hand side is the integral of an instantaneous error rate The estimate (P1) can be also useful in order to understand which kind of Lipschitz perturbations preserve the well-posedness property In the following sections we shall review three different settings where these ideas can be implemented Section 2 is devoted to discontinuous ODEs in a finite dimensional space [22] Following [5], for a vector field F = F (t, ) having finite directional variation, a general formula yielding a time-dependent contractive Riemann metric can here be given Section 3 reviews two different constructions of a distance functional which satisfies an estimate of the form (P1), in connection with the Camassa-Holm equation [1, 25] Finally, in Section 4 we discuss distance functionals which are contractive for the flow generated by a hyperbolic system of conservation laws [2, 12] 2

2 Discontinuous ODEs To motivate the search for contractive metrics, we start with two elementary eamples Eample 1 The ODE ẋ = 1/2 yields a tetbook case of a Cauchy problem with multiple solutions Yet, there is a simple way to select a unique solution for each initial data () = Let us define a solution t (t) to be admissible if and only if it is strictly increasing These admissible solutions are then unique, and depend continuously on the initial data For =, the corresponding admissible solution is t S t = (sign t) t 2 /4 Notice that the trajectory w(t) is not an admissible solution, but the error w(t) S t w() = t 2 /4 cannot be estimated integrating the instantaneous error rate ẇ(t) w(t) 1/2 On the other hand, a direct computation shows that the Riemann distance d y (, y) = ds s 1/2 is invariant wrt the flow of admissible solutions Namely d (S t, S t ȳ) = d (, ȳ) for every, ȳ, t Using this distance, the error estimate (17) retains its validity Indeed d ( ) w(t ), S T w() = d (, T 2 /4) = T d ( ) w(t + h), S h w(t) { T lim inf h + h dt = lim inf h + T 2 /4 ds s 1/2 = T, d (, h 2 } /4) dt = h T 1 dt = T Eample 2 Consider the discontinuous ODE ẋ = f() = { 1 if < 3 if, (28) For any initial data () =, the Cauchy problem is well posed Indeed, any two solutions satisfy 1 (t) 2 (t) 3 1 () 2 () (29) The estimate (29) alone, however, does not tell for which vector fields g( ) the ODE ẋ = f() + g() generates a continuous semigroup For eample, taking g() 2, the Cauchy problem is well posed, while taking g() 2 it is not The difference between the two above cases becomes apparent by introducing the equivalent distance d (, y) = 3 y if y, y if y, 3 + y if y Notice that d is invariant wrt the flow generated by (28) Denote by S g t and S g t the semigroups generated by the ODEs ẋ = g() and ẋ = g(), respectively Then for every, ȳ IR and t we have d (S g t, Sg t ȳ) d (, ȳ) (21) 3

On the other hand, taking, ȳ >, one has d ( ) S g t, S g t ȳ d (, ȳ) lim h + h = 2 (211) Comparing (21) with (211), we see that the flow generated by g contracts the distance d, while the flow generated by g can increase it, at a rate which does not approach zero as d (, ȳ) f g P P 1 P 2 P 3 P i 1 P i M Γ t Figure 1: Left: the vector field f is transversal to the surface where it is discontinuous The directional variation V M is a bounded function Right: the vector field g is not transversal to the surface where it has a discontinuity Its directional variation is thus unbounded Net, consider a general ODE with bounded, possibly discontinuous right hand side ẋ = f(t, ) IR n (212) In the Euclidean space IR 1+n, consider the cone with opening M: } Γ M = {(τ, y) ; y Mτ Following [4], the total directional variation of the vector field f up to the point (t, ) will be defined as { N } V M (t, ) = sup f(p i ) f(p i 1 ) ; N 1, P i P i 1 Γ M, P N = (t, ) i=1 (213) Notice that, in order that V M be bounded the jumps in f must be located along hypersurfaces which are transversal to the directions in the cone Γ M Otherwise, one can choose arbitrarily many points P i, alternatively on opposite sides of the discontinuity, and render the sum in (213) arbitrarily large (see Fig 1) Given two constants < L < M, we define the weighted length of a Lipschitz continuous path γ : [, 1] IR n at time t as ( 1 γ t = ep V M ) (t, γ(s)) γ(s) ds (214) M L The weighted distance between two points, y IR n at time t is defined as { } d t (, y) = inf γ t ; γ is a Lipschitz path joining with y (215) 4

Notice that, as t increases, the directional variation V M (t, ) increases as well (214), the weighted length of the path γ becomes smaller Hence, by We recall that a Carathéodory solution to the (possibly discontinuous) ODE (212) is an absolutely continuous function t (t) that satisfies (212) for ae time t The main result proved in [5] is as follows Theorem 1 Let f = f(t, ) be a time dependent vector field on IR n Assume that there eist constants L < M such that f(t, ) L for all t,, and the directional variation V M of f defined at (213) is locally bounded Then, for every initial data (t ) =, the ODE (212) has a unique, globally defined Carathéodory solution Any two solutions ( ), y( ) of (212) satisfy d τ ((τ), y(τ)) d t ((t), y(t)) for all t τ (216) Remark 1 Since the Euclidean distance between two nearby trajectories can suddenly increase across a surface where f is discontinuous, to achieve the contractive property (216) this must be compensated by the decrease of the eponential weight inside the integral in (214) 3 Nonlinear wave equations The Camassa-Holm equation can be written as a scalar conservation law with an additional integro-differential term: u t + (u 2 /2) + P =, (317) where P is defined as the convolution P = 1 2 e ( u 2 + u2 2 ) (318) For the physical motivations of this equation we refer to [14, 16, 17] One can regard (317) as an evolution equation on a space of absolutely continuous functions with derivatives u L 2 In the smooth case, differentiating (317) wrt one obtains ( ) u t + uu + u 2 u 2 + u2 2 + P = (319) Multiplying (317) by u and (319) by u one obtains the two balance laws ( ) ( ) ( ) ( ) u 2 u 3 u 2 + 2 3 + u P = u P, uu 2 + 2 2 u3 = u P (32) 3 t t As a consequence, for regular solutions the total energy E(t) = [ ] u 2 (t, ) + u 2 (t, ) d (321) 5

remains constant in time As in the case of conservation laws, because of the strong nonlinearity of the equations, solutions with smooth initial data can lose regularity in finite time For the Camassa-Holm equation (317), however, the uniform bound on u L 2 guarantees that only the L norm of the gradient can blow up, while the solution u itself remains Hölder continuous at all times The equation (317) admits multi-peakon solutions, depending on finitely many parameters These have the form N u(t, ) = p i (t)e qi(t), (322) i=1 where the coefficients p i, q i are obtained by solving the Hamiltonian system of ODEs q i = H(p, q), p i H(p, q) = 1 N p i p j e q i q j (323) ṗ i = 2 i,j=1 H(p, q), q i According to (322), the coefficient p i determines the amplitude of the i-th peakon, while q i describes its location By (32), the H 1 norm is constant in time along regular solutions The space H 1 (IR) = W 1,2 (IR) thus provides a natural domain where to construct global solutions to the Camassa- Holm equation Observe that, for u W 1,p with p < 2, the convolution (318) may not be well defined On the other hand, if p > 2, the Sobolev norm u(t) W 1,p of a solution can blow up in finite time u() u(t) q q 1 2 (t < T) u(t) u( τ) ( τ > T) Given an intial condition Figure 2: Interaction between two peakons having opposite strengths u() = u H 1 (IR), (324) by a solution of the Cauchy problem (317)-(324) on [, T ] we mean a Hölder continuous function u = u(t, ) defined on [, T ] IR with the following properties At each fied t we have u(t, ) H 1 (IR) Moreover, the map t u(t, ) is Lipschitz continuous from [, T ] into L 2 (IR), satisfying the initial condition (324) together with d dt u = uu P (325) 6

for ae t Here (325) is understood as an equality between functions in L 2 (IR) The solution is called conservative if the corresponding energy E(t) in (321) coincides ae with a constant function A globally defined flow of conservative solutions was constructed in [8] One should be aware, however, that the Cauchy problem for the Camassa-Holm equation is not well posed, even in the natural space H 1 (IR) Failure of continuous dependence on initial data can be seen by looking at special solutions with two opposite peakons (Fig 2) In this case we have p 1 (t) + p 2 (t), q 1 (t) + q 2 (t) Let T be the interaction time, so that q 1 (T ) = q 2 (T ) = As t T, one has p 1 (t) +, p 2 (t) q 1 (t) =, q 2 (t), (326) u(t) C, u(t) L 2, q2 (t) q 1 (t) u 2 (t, ) d E, (327) where E is the energy of the solution, which is a constant (ecept at t = T ) For detailed computations we refer to Section 5 in [8] The last limit in (327) shows that, as t T, all the energy is concentrated within the small interval [q 1 (t), q 2 (t)] between the two peakons Net, in addition to this special solution u, consider a family of solutions defined as u ε (t, ) = u(t ε, ) At time t =, as ε we have u ε (, ) u(, ) H 1 However, for any ε >, as t T one has u(t) u ε (t) 2 H = u(t, ) u ε (t, ) 2 d 1 IR ( + IR + IR \ [q 1 (t), q 2 (t)] [q 1 (t), q 2 (t) ) u (t, ) u ε (t, ) 2 d u ε (T, ) 2 d + u ε (T, ) 2 d + E 2 = 2E 2 IR (328) According to (328), solutions u ε which initially start arbitrarily close to u, within finite time split apart at a uniformly positive distance 2 E 31 A metric induced by optimal transportation In order to analyze uniqueness questions for solutions to the Camassa-Holm equations, it is of interest to construct an alternative distance functional J(, ) on H 1 For any two solutions of (317), this distance should satisfy an inequality of the form d J(u(t), v(t)) κ J(u(t), v(t)), (329) dt with a constant κ depending only on the maimum of the two norms u(t) H 1, v(t) H 1 (which remain ae constant in time) For spatially periodic conservative solutions to the Camassa-Holm equation, this goal was achieved in [1], building on insight gained in [7] We describe here the key steps of this construction Consider the unit circle T = [, 2π] with endpoints identified The distance between two angles θ, θ T will be denoted as θ θ Consider the manifold X = IR IR T with 7

distance d ( (, u, θ), (, ũ, θ) ) = ( + u ũ + θ θ ) 1, (33) where a b = min{a, b} Let H 1 per be the space of absolutely continuous periodic functions u, with u() = u( + 1) for every IR, and such that ( 1 1/2 u H 1 per = [u 2 () + u 2 ()] d) < Given u H 1 per(ir), define its etended graph Graph(u) = { (, u(), 2 arctan u ()) ; } IR X Moreover, let µ u be the measure supported on Graph(u), whose projection on the -ais has density 1 + u 2 wrt Lebesgue measure In other words, for every open set A X we require µ u (A) = (1 + u 2 ()) d { ; (,u(),2 arctan u ()) A} The distance J(u, v) between two functions u, v H 1 per is determined as the minimum cost for a constrained optimal transportation problem More precisely, consider the measures µ u, µ v, supported on Graph(u) and on Graph(v), respectively An absolutely continuous strictly increasing map ψ : IR IR satisfying the periodicity condition ψ( + 1) = ψ() + 1 for all IR will be called an admissible transportation plan Given ψ, we can move the mass µ u to µ v, from the point (, u(), 2 arctan u ()) to the point (ψ(), v(ψ()), 2 arctan v (ψ())) In general, however, the measure µ v is not equal to the push forward of the measure µ u determined by the map ψ We thus need to introduce an additional cost, penalizing this discrepancy Using the function { ( ) ( ) } φ 1 () = sup θ [, 1] ; θ 1 + u 2 () 1 + ũ 2 (ψ()) ψ (), the cost associated to the transportation plan ψ is now defined as J ψ (u, v) = 1 [distance] [transported mass] + 1 [ecess mass] 1 = d ( ) (, u(), 2 arctan u ()), (ψ(), ũ(ψ()), 2 arctan ũ (ψ()) φ 1 () (1 + u 2 ()) d (331) + 1 (1 + u 2 ()) (1 + ũ 2 (ψ())) ψ () d Minimizing over all admissible transportation plans, one obtains a distance functional: J(u, v) = inf ψ J ψ (u, v) (332) 8

v u +d ψ() ψ () + ψ ()d Figure 3: Transporting the mass from the graph of u to the graph of v The analysis in [1] shows that this functional is indeed a distance on H 1 per, and grows at the controlled rate (329) along any couple of conservative solutions to the Camassa-Holm equation In turn, this yields the uniqueness of conservative solutions, and a sharp estimate on their continuous dependence on initial data Remark 2 In the definition of the distance functional J, the requirement that the transportation must be achieved in terms of a non-decreasing function ψ plays an essential role Indeed, the topology generated by the distance J is different from the topology of weak convergence of measures, corresponding to the standard transportation distance { } d(µ u, µ v ) = sup f dµ u f dµ v, f Lip 1, (333) where the supremum is taken over all Lipschitz continuous functions with Lipschitz constant 1 For eample, consider the sequence of saw-tooth functions as in Fig 4, where u m is defined as the unique function of period 2 m such that u m () = min{, 2 m } [, 2 m ] Observe that µ um is a measure supported on Graph(u m ), whose projection on the -ais has constant density 1 + (u m ) 2 2 wrt Lebesgue measure As m, one has the weak convergence µ um µ, where µ is the sum of two copies of Lebesgue measure, one on the line {(,, π) ; IR}, and one on the line {(,, π) ; IR} In particular, the sequence (µ um ) m 1 is a Cauchy sequence wrt the distance (333) However, (u m ) m 1 is not a Cauchy sequence wrt the distance (332) 32 A metric induced by relabeling equivalence Net, we discuss an alternative approach to the construction of a distance functional J, having the controlled growth property (329) along solutions to the Camassa-Holm equation Here the starting point is the representation of solutions in terms of new variables, introduced in [8] As independent variables we use time t and an energy variable ξ IR, which is constant along characteristics This means that, in the t- plane, for each fied ξ the curve t y(t, ξ) provides a solution to the Cauchy problem d y(t) = u(t, y(t)), dt y(, ξ) = ȳ(ξ) 9

u n u m 2 m π 2 arctan u n π Figure 4: If one allows transportation plans ψ() which are not monotone, the optimal transportation of the measure µ un to the measure µ um can be achieved with a much smaller cost However, such plans are not allowed by the definition (332) In addition, we use the three dependent variables U = u, v = 2 arctan u, q = (1 + u 2 ) y ξ There is considerable freedom in the parameterization of characteristics A natural way to choose the function ξ ȳ(ξ) is to require that ȳ(ξ) (1 + ū 2 ) d = ξ At time t =, this achieves the identity q(, ξ) 1 t T P Figure 5: Characteristic curves, for a solution to the Camassa-Holm equation It is quite possible that characteristics join together at an isolated time T This happens, for eample, when two peakons cross each other as in Fig 2 In this case, as t T, the measure with density 1 + u 2 approaches a point mass at P However, in the variables (U, v, q), the solution of (334) remains smooth As u ± we simply have 2 arctan u ±π, and the singularity is completely resolved by the variable transformation 1

As proved in [8], for a given initial data u() = ū H 1 (IR), a conservative solution to the Camassa-Holm equation (317) can be constructed as follows As a first step, we solve the Cauchy problem U = P t U(, ξ) = ū(ȳ(ξ)) v t q t = 2(U 2 P ) cos 2 v 2 1 2 sin2 v 2 = P (t, ξ) = 1 2 (U 1 ) 2 P sin v q { ep ξ ξ cos 2 v(s) 2 } q(s) ds v(, ξ) q(, ξ) = 1 [ U 2 (ξ ) cos 2 v(ξ ) 2 = 2 arctan ū (ȳ(ξ)) This can be regarded as a Cauchy problem for an ODE on the Banach space E = H 1 L L (334) + 1 ] 2 sin2 v(ξ ) q(ξ ) dξ 2 By a fied point argument, one obtains a unique solution, globally defined for all t IR In turn, from a solution (U, v, q)(t, ξ) of (334) one recovers a solution u(t, ) of the Camassa-Holm equation (317) by setting t y(t, ξ) = ȳ(ξ) + U(τ, ξ) dτ and then defining u(t, ) = U(t, ξ) if = y(t, ξ) (335) As proved in [8], this procedure yields a group of solutions continuously depending on the initial data Namely, given a sequence of initial data such that ū n ū H 1, the corresponding solutions u n (t, ) converge to u(t, ) uniformly for t, in bounded sets By itself, this result does not guarantee the uniqueness of conservative solutions In principle one may use a completely different construction procedure (say, by vanishing viscosity approimations as in [33, 34]) and generate different solutions To construct a distance functional providing precise information on the continuous dependence of solutions, the approach developed in [25] is based on a relabeling technique See also [11] for an earlier result in connection with the Hunter-Saton equation As motivation, observe that the same solution u(t, ) of the Camassa-Holm equation (317) corresponds to infinitely many equivalent solutions (U, v, q)(t, ξ) of the system (334) Indeed, here the variable ξ is simply used as a label to identify different characteristics A smooth relabeling ξ ζ(ξ) would produce a different solution (Ũ, ṽ, q) of (334), with Ũ(t, ζ) = U(t, ξ), ṽ(t, ζ) = v(t, ξ), q(t, ζ) = q(t, ξ) ξ ζ However, the corresponding solution u(t, ) would be the same Given u H 1, consider the set of triples { F(u) = (U, v, q) ; there eists y( ) such that U(ξ) = u(y(ξ)), v(ξ) = 2 arctan u (y(ξ)), q(ξ) = (1 + u 2 (y(ξ)) 11 } ξ y(ξ)

One can define the functional J (u, ũ) = inf (U, v, q) (Ũ, ṽ, q) E, where the infimum is taken over all triples such that (U, v, q) F(u), (Ũ, ṽ, q) F(ũ) To achieve the triangle inequality, one needs to introduce a further functional J(u, ũ) = inf { N } J (u i, u i 1 ) ; u = u, u N = ũ (336) i=1 As shown in [25], in connection with spatially periodic solutions to the Camassa-Holm equation, this approach yields an alternative construction of a distance functional which satisfies the crucial property (329) Remark 3 While the well-posedness issue for the Camassa-Holm equation is now well understood, it remains a challenging open problem to establish similar results for the nonlinear wave equation u tt c(u)(c(u)u ) = (337) In this case, for each given data (u, u t ) H 1 L 2, one can define not one but two measures µ u +, µ u, accounting for the energy transported by forward and by backward moving waves Given two couples (u, u t ), (v, v t ), it is not clear how to etend a functional of the form (331) to a double transportation problem, relating the two couples of measures (µ u +, µ u ) and (µ v +, µ v ) As proved in [13], global conservative solutions to the equation (337), continuously depending on the initial data, can also be obtained by a nonlinear transformation of independent and dependent variables However, a relabeling technique here is hard to implement Indeed, in (335) the independent variables are related by (t, ) = (t, y(t, ξ)) On the other hand, for the equation (337), it is convenient to use independent variables X, Y which are constant along forward and backward characteristics, respectively This yields a transformation (X, Y ) (t(x, Y ), (X, Y )) where the time variable has no preferred status In general, for any constant c the set {(X, Y ) ; t(x, Y ) = c} has an awkward structure 4 Hyperbolic conservation laws In this last section we discuss the construction of a contractive metric for the system of conservation laws u t + f(u) = (438) Here u = (u 1,, u n ) IR n is the vector of conserved quantities and f = (f 1,, f n ) : IR n IR n is the flu function [3, 2, 26, 31, 32] For smooth solutions, this can be written in quasilinear form u t + A(u)u = A(u) = Df(u) We recall that the system is strictly hyperbolic if each Jacobian matri A(u) = Df(u) has real distinct eigenvalues λ 1 (u) < λ 2 (u) < < λ n (u) In this case, one can find dual bases of right 12

and left eigenvectors r i (u), l j (u), normalized so that r i (u) 1, l j (u) r i (u) = { 1 if i = j, if i j (439) The eistence and uniqueness of entropy admissible weak solutions to (438) was initially developed relying on the following assumption, stating that the directional derivative of eigenvalue in the direction of the corresponding eigenvector is identically zero, or has always the same sign [23, 28] (La Conditions) For each i {1,, n}, the i-th characteristic field is either linearly degenerate, so that Dλ i r i, or genuinely nonlinear, so that Dλ i r i > at every u IR n In the case of a scalar conservation laws, a fundamental result of Kruzhkov [27] valid also in several space dimensions shows that the L 1 distance between solutions does not increase in time Indeed, u(t, ) ũ(t, ) L 1 u(, ) ũ(, ) L 1, (44) where u(t, ), ũ(t, ) are any two bounded, entropy-admissible solutions of (438) Thanks to this property, solutions to a scalar conservation law can also be constructed relying on the abstract theory of contractive semigroups [18] For systems of two or more conservation laws, however, this contractive property fails In general one cannot even find any constant L for which the property (P1) holds For eample, consider a solution u = u(t, ) which initially contains two shocks, interacting at time τ and producing a third outgoing shock (see Fig 6, left) Let ũ be a perturbed solution, containing the same shocks, but slightly shifted in space As a result, the interaction occur a bit later, say at time τ + h In this case, the L 1 distance between the two solutions remains constant, ecept during the short interval [τ, τ + h] where it increases very rapidly (Fig 6, right) shocks in u shocks in u ~ t u ~ u L 1 τ+ h τ τ τ+h t Figure 6: The L 1 distance between two nearby solutions can increase rapidly, during a short interval of time Under the La conditions, two approaches are now available, in order to construct a distance functional on a domain D of functions with small total variation, 13

41 An eplicit functional In [12] an eplicit formula was introduced, providing a functional Φ with and such that u v L 1 Φ(u, v) C u v L 1, (441) Φ(u(t, ), v(t, )) Φ(u(t, ), v(t, )) t < t, (442) for any couple of entropy-admissible weak solutions u, v to (438), with sufficiently small total variation We review here the basic step of this construction 1 Measuring the strength of shock and rarefaction waves Fi a state u IR n and an inde i {1,, n} As before, let r 1 (u),, r n (u) be the right eigenvectors of the Jacobian matri A(u) = Df(u), normalized as in (439) The integral curve of the vector field r i through the point u is called the i-rarefaction curve through u It is obtained by solving the Cauchy problem in state space: du dσ = r i(u), u() = u (443) This curve, parameterized by arc-length, will be denoted as σ R i (σ)(u ) (444) Net, for a fied u IR n and i {1,, n}, one can show that there eists (locally, in a neighborhood of u ) a unique smooth curve of states u which can be connected to the right of u by an i-shock, satisfying the Rankine-Hugoniot equations λ(u u ) = f(u) f(u ), (445) for some scalar speed λ, with λ λ i (u ) as u u This will be called the i-shock curve through the point u and parameterized by arc-length: σ S i (σ)(u ) (446) It is well known that the two curves R i, S i have a second order contact at the point u More precisely, the following estimates hold { Ri (σ)(u ) = u + σr i (u ) + O(1) σ 2, S i (σ)(u ) = u + σr i (u ) + O(1) σ 2 (447), R i (σ)(u ) S i (σ)(u ) = O(1) σ 3, (448) Here and throughout the following, the Landau symbol O(1) denotes a quantity whose absolute value satisfies a uniform bound, depending only on the system (33) Notice that the orientation of the unit vector r i (u ) determines an orientation of the curves R i, S i Recalling the La conditions, if the i-th characteristic field is genuinely nonlinear, the orientation is chosen so that the characteristic speed λ i increases along the curves, as the parameter σ increases On the other hand, if the i-th field is linearly degenerate, one can prove that R i (σ) = S i (σ) for every σ, and that λ i is constant along these curves In this case, the orientation can be chosen arbitrarily 14

S i R i u r (u ) i Figure 7: Shock and rarefaction curves through the point u 2 The interaction potential It will be convenient to work within a special class of functions, which we call PCS, consisting of all piecewise constant functions u : IR IR n, with simple jumps We say that the jump at is simple if either u(+) = R i (σ)(u( )) for some σ > or u(+) = S i (σ)(u( )) for some σ < In both cases, we regard σ as the strength of the jump at For a piecewise constant function u PCS, let α, α = 1,, N, be the locations of the jumps in u( ) Moreover, let σ α be the strength of the wave-front at α, say of the family k α {1,, n} Following [23], we consider the two functionals V (u) = N σ α, (449) α=1 measuring the total strength of waves in u, and Q(u) = (α,β) A σ α σ β, (45) measuring the wave interaction potential In (45), the summation ranges over the set A of all couples of approaching wave-fronts More precisely, two fronts, located at points α < β and belonging to the characteristic families k α, k β {1,, n} respectively, are approaching if k α > k β or else if k α = k β and at least one of the wave-fronts is a shock of a genuinely nonlinear family Roughly speaking, two fronts are approaching if the one behind has the larger speed (and hence it can collide with the other, at a future time) If now u = u(t, ) is a piecewise constant approimate solution, a key observation is that the total strength of waves V (u(t)) can increase in time, but the interaction potential is monotone decreasing Indeed, consider a time τ where two fronts of strength σ, σ collide Then the changes in V, Q are estimated by V (τ) = V (τ+) V (τ ) = O(1) σ σ, (451) Q(τ) = Q(τ+) Q(τ ) = σ σ + O(1) σ σ V (τ ) σ σ, (452) 2 provided that V (τ ) is sufficiently small Indeed (Fig 8), after time τ the two colliding fronts σ, σ are no longer approaching Hence the product σ σ is no longer counted within the summation (45) On the other hand, the new waves σ k emerging from the interaction (having strength O(1) σ σ ) can approach all the other fronts not involved in the interaction (which have total strength V (τ ) ) 15

t σ i σ k σ j τ σ σ" σ α Figure 8: Estimating the change in the total variation at a time where two fronts interact By (451) and (452) we can thus choose a constant C large enough so that the quantity V (u(t)) + C Q(u(t)) is monotone decreasing, provided that V remains sufficiently small In turn, this yields an estimate on the total variation, globally in time: TotVar{u(t)} V (u(t)) V (u()) + C Q(u()) (453) By Helly s theorem, this provides a crucial compactness property, toward a proof of the eistence of globally defined weak solutions [23] 3 A weighted distance functional Relying on the concepts and notations developed above, we can now describe the construction of a functional Φ(u, v), measuring the distance between solutions to the hyperbolic system (438) and satisfying the key properties (441)- (442) Given two piecewise constant functions with simple jumps u, v : IR R n, recalling the construction of shock curves at (446), consider the scalar functions q i defined implicitly by v() = S n (q n ()) S 1 (q 1 ())(u()) (454) ω = v() 3 q 3 λ 3 σ α α v ω 2 λ 1 ω 1 ω = u() q 1 α u Figure 9: Decomposing a jump (u(), v()) in terms of n (possibly non-admissible) shocks Defining the intermediate states ω i = Si (q i ()) S 1 (q 1 ())(u()) i =, 1, 2,, n, this means that each couple of states ω i 1, ω i is connected by an i-shock of size q i () We regard q i () as the strength of the i-th component in the jump v() u(), measured along 16

shock curves Since these curves are parameterized by arc length, as long as u(), v() vary in a small neighborhood of the origin one clearly has v() u() n q i () C 1 v() u() (455) i=1 for some constant C 1 We can now define the functional Φ(u, v) = n q i () W i () d, (456) i=1 where the weights W i are defined by setting: W i () = 1 + κ 1 [total strength of waves in u and in v which approach the i-wave q i ()] +κ 2 [wave interaction potentials of u and of v] = 1 + κ 1 A i () + κ 2 [Q(u) + Q(v)] (457) The quantity A i (), accounting for the total strength of waves approaching an i-wave located at, is defined as follows If the i-th characteristic field is linearly degenerate, we simply take A i () = + σ α (458) α<, i<k α n α>, 1 k α<i The summations here etend to waves both of u and of v Here k α {1,, n} is the family of the jump located at α with size σ α On the other hand, if the i-th field is genuinely nonlinear, the definition of A i contains an additional term, accounting for waves in u and in v of the same i-th family: A i () = α J (u) J (v) α<, i<kα n + + α J (u) J (v) α>, 1 kα<i kα=i α J (u), α< kα=i α J (v), α< + + σ α kα=i α J (v), α> kα=i α J (u), α> σ α if q i () <, σ α if q i () > Here J (u) and J (v) denote the sets of all jumps in u and in v, while J = J (u) J (v) (459) As soon as the functional Φ is defined for piecewise constant functions, it can be etended to all functions u L 1 (IR IR n ) having suitably small total variation, by taking the lower semicontinuous envelope: Φ (u, v) = lim inf Φ(u, v ) (46) u u, v v u,v P CS By choosing the constants κ 2 >> κ 1 >> 1 in (457) sufficiently large, if the total variation of the functions u, v remains small, the analysis in [12] shows that this functional is equivalent 17

to the L 1 distance and is non-increasing in time along couples of entropy-weak solutions to the system (438) We remark that the functional Φ (, ) in (46) is still not a distance, because it may not satisfy the triangle inequality To achieve a distance, as in (336) one should define d Φ (u, v) = inf { N } Φ(u i, u i 1 ) ; u = u, u N = ũ i=1 In practice, it is more convenient to work out all the estimates on piecewise constant approimate solutions, using the eplicit formula (457) The limits of approimate solutions, providing eact solutions, are taken only at the end An etension of these ideas to the initial-boundary value problem can be found in [21] 42 A Riemann type distance With this approach, introduced in [2], one considers a family of sufficiently regular paths γ : [, 1] L 1, for which a weighted length can be defined For any couple of functions u, ũ, the weighted distance d (u, ũ) is then defined as the infimum of lengths of all paths connecting u with ũ In connection with the system (438) we say that a function u : IR IR n is in the class PLSD (Piecewise Lipschitz with Simple Discontinuities) if u is piecewise Lipschitz continuous with finitely many jumps, each jump consisting of a single, entropy admissible shock In other words, at each point α where u has a jump, the left and right stated are related by u( α +) = S i (σ α )(u( α )), (461) for some genuinely characteristic field i, and for some σ α < The condition on the sign of σ α guarantees that the shock is admissible If u is in PLSD and has N discontinuities at the points 1 < < N, the space of generalized tangent vectors at u is defined as T u = L 1 IR N Adopting the point of view of differential geometry, elements in T u can be interpreted as first order tangent vectors as follows On the family Σ u of all continuous paths γ : [, ε ] L 1 with γ() = u, define the equivalence relation γ γ 1 iff lim ε + ε γ(ε) γ (ε) L 1 = (462) We say that a continuous path γ Σ u generates the tangent vector (v, ξ) T u if γ is equivalent to the path γ (v,ξ;u) defined as γ (v,ξ;u) (ε) = u + εv + ξ α< [ u( + α ) u( α ) ] χ [α+εξ α, α] ξ α> [ u( + α ) u( α ) ] χ [α, α+εξ α], (463) where χ I denotes the characteristic function of the interval I Up to higher order terms, γ(ε) is thus obtained from u by adding εv and by shifting the points α, where the discontinuities of u occur, by εξ α 18

To define a norm on each tangent space T u, we proceed as follows Let u be a function in the class PLSD, with jumps at the points 1 < 2 < < N For any (v, ξ) T u = L 1 IR N, define the scalar components v i () = l i (u()) v() (464) Here l 1,, l n are the left eigenvectors of the Jacobian matri A(u) = Df(u), normalized as in (439) Following [2], the weighted norm is defined as (v, ξ) u N m = σ α ξ α Wk u α ( α ) + α=1 i=1 v i () W u i () d, (465) where Wi u () is the weight given to an i-wave located at More precisely: where A u i () = W u i () = 1 + κ 1 A u i () + κ 2 Q(u), (466) + u j (y) dy + + σ α j i j i k α i k α i α> α< measures the total amount of waves in u approaching an i-wave located at, while Q(u) is the interaction potential, introduced at (45), and 1 << κ 1 << κ 2 are suitable constants u+ εv u α + εξ α α Figure 1: A piecewise Lipschitz continuous function u and a perturbation described by the tangent vector (v, ξ) T u Let now γ : [, 1] L 1 be a Lipschitz continuous curve such that, for all but finitely many values of θ, the functions γ(θ) is in PLSD and the tangent vector γ = (v(θ), ξ(θ)) T γ(θ) is well defined One can then define the weighted length of γ by integrating the weighted norm of its tangent vector: 1 γ = γ(θ) γ(θ) dθ (467) In turn, this provides a notion of distance between two functions u, ũ, as in (13)-(14) By choosing the constants κ 1, κ 2 large enough, this distance is non-increasing along any couple of entropy-weak solutions to the hyperbolic system (438), having suitably small total variation See [6, 9] for two implementations of this approach Remark 4 All of the previous analysis dealt with solutions having small total variation An etension to large BV data has been achieved in [29] In this case, a contractive metric can be 19

constructed on a domain of functions consisting of small BV perturbations of a (possibly large) Riemann solution While the total strength of waves V (u) here can be large, the interaction potential Q(u) must remain sufficiently small We remark that, for general initial data with large interaction potential, the a priori BV estimates in [23] do not apply and even the global eistence of weak solutions remains an open problem Remark 5 For strictly hyperbolic systems which do not satisfy the La conditions, a Lipschitz semigroup of globally defined, entropy weak solutions was constructed in [1], taking limits of vanishing viscosity approimations In this general case, a distance which is contractive wrt the flow generated by (438) has not yet been constructed Etending the eplicit definition (456)-(457) appears to be a very difficult task On the other hand, since the continuous dependence of viscous approimations was proved in [1] by studying the weighted length of smooth paths of solutions, constructing a Riemann type metric as in (465)-(467) may be a more promising approach References [1] S Bianchini and A Bressan, Vanishing viscosity solutions of nonlinear hyperbolic systems, Annals of Mathematics 161 (25), 223-342 [2] A Bressan, A locally contractive metric for systems of conservation laws, Ann Scuola Normale Sup Pisa, Serie IV, Vol XXII (1995), 19-135 [3] A Bressan, Hyperbolic Systems of Conservation Laws The One Dimensional Cauchy Problem Oford University Press, 2 [4] A Bressan, Unique solutions for a class of discontinuous differential equations, Proc Amer Math Soc 14 (1988), 772-778 [5] A Bressan and G Colombo, Eistence and continuous dependence for discontinuous ODE s, Boll Un Mat Ital 4-B (199), 295-311 [6] A Bressan and R M Colombo, The semigroup generated by 2 2 conservation laws, Arch Rational Mech Anal 133 (1995), 1-75 [7] A Bressan and A Constantin, Global solutions to the Hunter-Saton equations, SIAM J Math Anal 37 (25), 996 126 [8] A Bressan and A Constantin, Global conservative solutions to the Camassa-Holm equation, Arch Rat Mech Anal 183 (27), 215-239 [9] A Bressan, G Crasta, and B Piccoli, Well posedness of the Cauchy problem for n n systems of conservation laws, Amer Math Soc Memoir 694 (2) [1] A Bressan and M Fonte, An optimal transportation metric for solutions of the Camassa- Holm equation, Methods and Applications of Analysis, 12 (25), 191 22 [11] A Bressan, H Holden, and X Raynaud, Lipschitz metric for the Hunter-Saton equation, J Mathématiques Pures Appliqués 94 (21), 68-92 2

[12] A Bressan, T P Liu and T Yang, L 1 stability estimates for n n conservation laws, Arch Rational Mech Anal 149 (1999), 1-22 [13] A Bressan and Y Zheng, Conservative solutions to a nonlinear variational wave equation, Comm Math Phys 266 (26), 471-497 [14] R Camassa and D D Holm, An integrable shallow water equation with peaked solitons, Phys Rev Lett 71 (1993), 1661-1664 [15] R M Colombo and A Corli, A semilinear structure on semigroups in a metric space Semigroup Forum 68 (24), 419444 [16] A Constantin and H P McKean, A shallow water equation on the circle, Comm Pure Appl Math 52 (1999), 949-982 [17] A Constantin and L Molinet, Global weak solutions for a shallow water equation, Comm Math Phys 211 (2), 45-61 [18] M G Crandall, The semigroup approach to first-order quasilinear equations in several space variables, Israel J Math 12 (1972), 18-132 [19] M G Crandall and T M Liggett, Generation of semi-groups of nonlinear transformations on general Banach spaces Amer J Math 93 (1971), 265298 [2] C Dafermos, Hyperbolic Conservation Laws in Continuum Physics, Dafermos, Constantine M Hyperbolic conservation laws in continuum physics Second edition Springer- Verlag, Berlin, 25 [21] C Donadello and A Marson, Stability of front tracking solutions to the initial and boundary value problem for systems of conservation laws Nonlin Diff Equat Appl 14 (27), 569 592 [22] A F Filippov, Differential Equations with Discontinuous Righthand Sides, Kluwer Academic Publishers, Dordrecht, 1988 [23] J Glimm, Solutions in the large for nonlinear hyperbolic systems of equations, Comm Pure Appl Math 18 (1965), 697-715 [24] D Henry, Geometric theory of semilinear parabolic equations Lecture Notes in Mathematics, 84 Springer-Verlag, Berlin-New York, 1981 [25] K Grunert, H Holden, and X Raynaud, Lipschitz metric for the periodic Camassa- Holm equation Preprint, 21 [26] H Holden and N H Risebro Front Tracking for Hyperbolic Conservation Laws, Springer Verlag, New York 22 [27] S Kruzhkov, First order quasilinear equations with several space variables, Math USSR Sbornik 1 (197), 217-243 [28] P La, Hyperbolic systems of conservation laws II, Comm Pure Appl Math 1 (1957), 537-566 [29] M Lewicka, Well-posedness for hyperbolic systems of conservation laws with large BV data Arch Rational Mech Anal 173 (24), 415 445 21

[3] R H Martin Jr, Nonlinear operators and differential equations in Banach spaces Wiley- Interscience, New York, 1976 [31] D Serre, Systems of Conservation Laws I, II, Cambridge University Press, 2 [32] J Smoller, Shock Waves and Reaction-Diffusion Equations, Springer-Verlag, New York, 1983 [33] Z Xin and P Zhang, On the weak solutions to a shallow water equation, Comm Pure Appl Math 53 (2), 1411-1433 [34] Z Xin and P Zhang, On the uniqueness and large time behavior of the weak solutions to a shallow water equation, Comm Partial Differential Equations 27 (22), 1815-1844 [35] K Yosida, Functional Analysis Reprint of the sith (198) edition Classics in Mathematics Springer-Verlag, Berlin, 1995 22