This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Ill-posedness of degenerate dispersive equations

David M. Ambrose Department of Mathematics
Drexel University
ambrose@math.drexel.edu
Gideon Simpson Department of Mathematics
University of Toronto
simpson@math.toronto.edu
J. Douglas Wright Department of Mathematics
Drexel University
jdoug@math.drexel.edu
 and  Dennis G. Yang Department of Mathematics
Drexel University
gyang@math.drexel.edu
Abstract.

In this article we provide numerical and analytical evidence that some degenerate dispersive partial differential equations are ill-posed. Specifically we study the K(2,2)K(2,2) equation ut=(u2)xxx+(u2)xu_{t}=(u^{2})_{xxx}+(u^{2})_{x} and the “degenerate Airy” equation ut=2uuxxxu_{t}=2uu_{xxx}. For K(2,2)K(2,2) our results are computational in nature: we conduct a series of numerical simulations which demonstrate that data which is very small in H2H^{2} can be of unit size at a fixed time which is independent of the data’s size. For the degenerate Airy equation, our results are fully rigorous: we prove the existence of a compactly supported self-similar solution which, when combined with certain scaling invariances, implies ill-posedness (also in H2H^{2}).

1. Introduction

Dispersion plays a pivotal role in the analysis of the governing equations for a large number of physical phenomena, ranging from the evolution of surface water waves to the formation of Bose-Einstein condensates. A common theme in such problems is that uniform dispersive effects frequently control nonlinear terms which may otherwise lead to ill-posedness. Consider for instance the following class of quasilinear equations of KdV type studied in [8]:

(1) ut=g3uxxx+g2uxx+g1ux+g0u+hu_{t}=g_{3}u_{xxx}+g_{2}u_{xx}+g_{1}u_{x}+g_{0}u+h

where gj=gj(xju,,u)g_{j}=g_{j}(\partial_{x}^{j}u,...,u) and h=h(x,t)h=h(x,t). There it is shown that such equations are well-posed provided |g3|c>0|g_{3}|\geq c>0 and g20g_{2}\geq 0. The first of these conditions guarantees that dispersive effects from g3uxxxg_{3}u_{xxx} do not vanish while the second prohibits the term g2uxxg_{2}u_{xx} from acting like a backwards heat operator. Similarly, the results on existence of solutions to the quasilinear dispersive equations found in [15] (about Schrödinger equations), [1] (water waves) and [29] (magma dynamics) share the feature that the dispersive effects are nondegenerate, to name just a few examples.

On the other hand, there are a number of physical problems in which the mechanism which causes the dispersive effects is not only nonlinear but degenerate. Some examples can be found in the study of sedimentation [26], shallow water [5], granular media [19, 11, 20], numerical analysis [14] and elastic rods and sheets [9, 16]. While there are a large number of articles in which degenerate dispersive equations (henceforth referred to as DDE) are derived [19, 20, 11, 5, 16, 9, 26], special solutions computed [24, 25, 11, 20, 19] or numerical computation of solutions performed [28, 27, 24, 10, 17, 6], the existence theory for the initial value problem of these equations is largely undeveloped.111The notable exception to this is the well-studied Camassa-Holm equation utuxxt=3uux+2uxuxx+uuxxxu_{t}-u_{xxt}=-3uu_{x}+2u_{x}u_{xx}+uu_{xxx} which when rewritten as a nonlocal evolution equation could be said to be degenerate and dispersive. This equation is integrable and consequently many results for it do not obviously generalize. The articles [10, 3, 13] prove some a priori estimates for DDE. [3] and [2] prove existence (but not uniqueness) of solutions to a very special class of DDE. [12] contains a semi-rigorous discussion of existence issues for these sorts of equations, including defining “δ\delta-entropy solutions” by borrowing ideas from first order conservation laws. None of these articles fully settles the well-posedness issues for degenerate dispersion, and thus our interest is in studying the Cauchy problem for some simple and common DDE.

In particular we study the following two equations:

(2) ut=2uuxxx+6uxuxx+2uuxu_{t}=2uu_{xxx}+6u_{x}u_{xx}+2uu_{x}

and

(3) ut=2uuxxxu_{t}=2uu_{xxx}

for x𝐑x\in{\bf{R}}. The first of these is the K(2,2)K(2,2) equation of Rosenau & Hyman [24]. Their goal was to understand the role of degenerate nonlinear dispersion in the formation of patterns and this equation (together with other members of the K(m,n)K(m,n) family ut=(um)x+(un)xxxu_{t}=(u^{m})_{x}+(u^{n})_{xxx}) has come to serve as the paradigm for degenerate dispersive systems. The most celebrated feature of this equation is the existence of compactly supported traveling waves, a.k.a. “compactons.” These are given by

u(x,t)=Qλ(x+λt):=4λ3cos2(x+λt4)χI(x+λt){u(x,t)=Q_{\lambda}(x+\lambda t):={4\lambda\over 3}\cos^{2}\left({x+\lambda t\over 4}\right)\chi_{I}(x+\lambda t)}

where χI\chi_{I} is the indicator function for the interval I=[2π,2π]I=[-2\pi,2\pi] and λ0\lambda\neq 0. More recently, K(2,2)K(2,2) has been used to model pulses in both elastic and granular media [16, 21]. Equation (3), a “degenerate Airy equation”, appeared in [13] and is amongst the simplest equations which could be said to feature degenerate dispersion. We study it here due to its similarity to the K(2,2)K(2,2) equation and because it is of the form of (1) but does not meet the uniform dispersion hypothesis needed in [8].

Surprisingly, our conclusions are that both equations are ill-posed for data in H2H^{2}. In particular, their solutions do not depend smoothly on their initial conditions. Our results for (2) are computational in nature: we conduct a series of numerical simulations of (2) which demonstrate that data which is very small in H2H^{2} can be of unit size at a fixed time which is independent of the data’s size. For (3) our results are fully rigorous: we prove the existence of a compactly supported self-similar solution which, when combined with certain scaling invariances, implies ill-posedness. Note that our focus on H2H^{2} initial data is due to the fact the compacton solutions of K(2,2)K(2,2) have a discontinuity in their second derivative at the edge of their support [18]. Consequently, they reside in H2H^{2} but not H3H^{3}, and so it is natural to work in this space.

Our suspicion that (2) is ill-posed has its origin in the various numerical simulations of collisions between compactons in K(m,n)K(m,n) equations. A wide variety of numerical methods has been used to compute solutions: pseudo-spectral [24], particle [6], Padé [28, 27], discontinuous Galerkin [17], and finite difference [10]. Each of these simulations shows that the computed collisions between compactons are nearly elastic: the compactons emerge from the collision nearly identical to their original state but phase shifted. As is common in collisions between solitary waves in “nearly” integrable PDE, the simulations show that in addition to the outgoing waves there is a small amplitude disturbance behind the waves. In nondegenerate problems, similar disturbances are sometimes called “dispersive ripples” and are viewed as being linear effects—the tail (approximately) solves the linearized equation [30, 31]. However, since the linearization of (2) about zero is trivial this heuristic cannot apply. The disturbance is described variously as: a compacton/anti-compacton pair in [24]; a numerical artifact in [28, 10]; and as a shock in [27]. No such disturbance is present at all in [6]. Lastly, the authors of [17] describe a “high-frequency oscillation” but go on to state “we do not know what is the source of these oscillations…” In each case, the qualitative features of the disturbance are quite different. Our feeling is that these discrepancies are not related to differences in the numerical methods or the implementation thereof but rather are due to an underlying issue with the equation itself.

Moreover, inspection of the right hand side of (2) provides evidence that the equation may be ill-posed. Observe that when ux<0u_{x}<0, the term 6uxuxx6u_{x}u_{xx} is formally a backwards heat operator. The key question, initially raised in [24], is this: can dispersive effects due to 2uuxxx2uu_{xxx} counteract the instability caused by the backwards heat operator 6uxuxx6u_{x}u_{xx}? Thus we should seek to understand the dispersive effects due to 2uuxxx2uu_{xxx}—hence our interest in (3). Our results show that the answer to the key question is “no”. As we stated above, our results show that 2uuxxx2uu_{xxx} is, in its own right, just as problematic as the backwards heat term. Differentiation of (3) with respect to xx yields for w:=uxw:=u_{x}

wt=2uwxxx+2uxwxx.w_{t}=2uw_{xxx}+2u_{x}w_{xx}.

Note that the second term on the RHS above is also a backwards heat operator when ux<0u_{x}<0. The backwards heat term evident in (2) is, in this way, hidden in (3). Thus we expect blow up of uxu_{x} when uxu_{x} is negative. Moreover, this blow up should be particularly catastrophic when the function crosses the xx-axis with negative slope.

This motivates our approach: the self-similar solution constructed has such a transverse crossing and the resulting ill-posedness is due to blow up in the first and second derivatives of the solution. While our proof that (3) is ill-posed does not naturally carry over to (2), the numerics are highly suggestive. Most tellingly, since the numerics are performed using strictly positive initial data, they suggest the ill-posedness can manifest even in the absence of such a transverse crossing.

Remark 1.

In [7], Craig & Goodman demonstrate that the differential equation ut=xuxxxu_{t}=-xu_{xxx} is ill-posed, whereas ut=xuxxxu_{t}=xu_{xxx} is not only well-posed but the solution map is essentially infinitely smoothing. The proofs of these facts follow from explicit formula for solutions of the initial value problems which can be constructed for each equation by mean of the Fourier transform and the method of characteristics.

The remainder of this paper is organized as follows. In Section 2 we show that the existence of a self-similar solution to (3) gives, as a consequence, the ill-posedness of the equation. In Section 3 we construct the self-similar solution using techniques from spatial dynamics. In Section 4 we discuss our numerical strategy for demonstrating ill-posedness of (2) and present the results of our simulations.

Acknowledgements: We would like to thank Professor Philip Rosenau for many interesting and helpful conversations regarding the degenerate dispersive equations studied in this article. Additionally, the work was supported by NSF Grants DMS 0926378, DMS 1008387, and DMS 1016267 (D.M.A.) and DMS 0908299 and DMS 0807738 (J.D.W.). G. S. was partially funded by NSERC. Finally, we are especially grateful the Louis and Bessie Stein Family Foundation who generously supported this work.

2. Self-similar solutions imply ill-posedness for (3)

We will prove there is a self-similar solution for (3) of the form

u(x,t)=A(x(1t)1/3)u(x,t)=A\left(\displaystyle\frac{x}{(1-t)^{1/3}}\right)

where AA gives the profile. Though other choices for the scaling are possible, we choose this particular set so that our proposed solution exhibits blow up (as t1t\to 1^{-}) in the first and second derivative while the LL^{\infty} norm remains bounded. Inserting this into (3) yields:

(4) A′′′=τA6A.A^{\prime\prime\prime}={\tau A^{\prime}\over 6A}.

Here τ=x(1t)1/3\tau=x(1-t)^{-1/3} is the self-similar coordinate upon which AA depends and we have ()=d/dτ()(\square)^{\prime}=d/d\tau(\square). We specify A(0)=0A(0)=0 and A(0)=1A^{\prime}(0)=-1 so that the self-similar solution crosses the xx-axis transversely with negative slope, the “bad” situation identified in the Introduction.

We prove the following proposition concerning solutions of (4):

Proposition 1.

There exists μ>0\mu_{*}>0, τ>0\tau_{*}>0, C>1C_{*}>1 and a function A(τ)C3([0,τ))A_{*}(\tau)\in C^{3}\left([0,\tau_{*})\right) with the following properties:

  1. (i)

    A(τ)\displaystyle A_{*}(\tau) solves (4) for all τ[0,τ)\tau\in[0,\tau_{*}).

  2. (ii)

    A(0)=0A_{*}(0)=0, A(0)=1A_{*}^{\prime}(0)=-1 and A′′(0)=μA_{*}^{\prime\prime}(0)=\mu_{*}.

  3. (iii)

    A(τ)A_{*}(\tau) is negative in (0,τ)(0,\tau_{*}) and has a unique minimum in the interior of this interval.

  4. (iv)

    limττA(τ)=limττA(τ)=0\displaystyle\lim_{\tau\to\tau_{*}^{-}}A_{*}(\tau)=\lim_{\tau\to\tau_{*}^{-}}A^{\prime}_{*}(\tau)=0.

  5. (v)

    C1|log|ττ|||A′′(τ)|C|log|ττ||C_{*}^{-1}\left|\log\left|\tau-\tau_{*}\right|\right|\leq|A^{\prime\prime}_{*}(\tau)|\leq C_{*}\left|\log\left|\tau-\tau_{*}\right|\right| for all τ(0,τ)\tau\in(0,\tau_{*}) sufficiently close to τ\tau_{*}.

The proof of this proposition is technical and difficult and we postpone it until the next section. First we utilize it to prove our main result for (3):

Theorem 1.

The Cauchy problem for (3), posed in H2H^{2}, does not depend smoothly on the initial data.

Proof.

(Theorem 1) The proof of the theorem is classical: we construct a sequence of solutions which are arbitrarily small in H2H^{2} at t=0t=0 but which are arbitrarily large when t=1t=1. Let

A(τ):={A(τ) if τ[0,τ)A(τ) if τ(τ,0] and 0 otherwise.A(\tau):=\begin{cases}A_{*}(\tau)&\textrm{ if }\tau\in[0,\tau_{*})\\ -A_{*}(-\tau)&\textrm{ if }\tau\in(-\tau_{*},0]\textrm{ and }\\ 0&\textrm{ otherwise}.\end{cases}

This is the profile function for our self-similar solution. See Figure 1.

Refer to caption
Figure 1. Numerically computed profile A(τ)A(\tau). Note μ=μ0.354875\mu=\mu_{\star}\sim 0.354875.

Note that it is compactly supported, not unlike the famous Barenblatt solution of the porous medium equation [4]. Also observe that A′′A^{\prime\prime} is discontinuous at τ=0\tau=0 and diverges logarithmically at ±τ\pm\tau_{*}. Nevertheless, these singularities are sufficiently mild so that AH2A\in H^{2}. Also observe that τA\tau A^{\prime} is continuous for all τ\tau and consequently so is AA′′′AA^{\prime\prime\prime}. Thus the function

u1,1(x,t):=A(x(1t)1/3)u_{1,1}(x,t):=A\left(\displaystyle\frac{x}{(1-t)^{1/3}}\right)

is in fact a classical solution of (3). That is to say, while x3u1,1L2\partial_{x}^{3}u_{1,1}\notin L^{2}, u1,1x3u1,1u_{1,1}\partial_{x}^{3}u_{1,1} is. Compare this with the compacton solution of K(2,2)K(2,2): it is not in H3H^{3}, though its square is.

For any T𝐑T\in{\bf{R}} and λ>0\lambda>0, (3) is invariant under the following transformation

(5) u(x,t)λu(xλ1/3,tT).u(x,t)\longrightarrow\lambda u\left({x\over\lambda^{1/3}},t-T\right).

Thus we have a two-parameter family of self-similar solutions of the form:

uλ,T(x,t):=λA(xλ1/3(Tt)1/3).u_{\lambda,T}(x,t):=\lambda A\left({x\over\lambda^{1/3}(T-t)^{1/3}}\right).

An elementary computation shows that

(6) uλ,T(,t)H2=λ7/6(Tt)1/6AL2+λ5/6(Tt)1/6AL2+λ1/2(Tt)1/2A′′L2.\|u_{\lambda,T}(\cdot,t)\|_{H^{2}}=\lambda^{7/6}(T-t)^{1/6}\|A\|_{L^{2}}\\ +\lambda^{5/6}(T-t)^{-1/6}\|A^{\prime}\|_{L^{2}}+\lambda^{1/2}(T-t)^{-1/2}\|A^{\prime\prime}\|_{L^{2}}.

Fixing 0<ε10<\varepsilon\ll 1, setting λ(ε)=ε2\lambda(\varepsilon)=\varepsilon^{2} and T(ε)=1+ε16T(\varepsilon)=1+\varepsilon^{16} results in

uλ(ε),T(ε)(,0)H2Cεanduλ(ε),T(ε)(,1)H2Cε.\|u_{\lambda(\varepsilon),T(\varepsilon)}(\cdot,0)\|_{H^{2}}\leq C\varepsilon\quad\textrm{and}\quad\|u_{\lambda(\varepsilon),T(\varepsilon)}(\cdot,1)\|_{H^{2}}\geq{C\over\varepsilon}.

(Here C>0C>0 is a nonessential constant.) This pair of inequalities demonstrates that we have constructed a sequence of initial data for (3) which is arbitrarily small but whose solution at time t=1t=1 is arbitrarily large. Thus the equation is ill-posed in H2H^{2}.

Remark 2.

In [13], the authors investigate self-similar solutions of the DDE ut=(uux)xxu_{t}=(uu_{x})_{xx}. In particular they numerically compute profiles for shock and rarefaction solutions which are, in many respects, quite similar to the ones we find here. Namely, the shock solutions cross the xx-axis with negative slope and form a singularity in uxu_{x} at the time of the shock formation. The solutions they find are not in L2(𝐑)L^{2}({\bf{R}})—either they do not converge to zero as xx\to\infty or their decay is not rapid enough to be integrable. Our focus here is on well-posedness issues in H2H^{2} and consequently our strategies and methods will differ from those of [13].

3. Construction of the self-similar solution

This section is devoted to proving Proposition 1. We rewrite (4) as the initial value problem (IVP):

(7) 6AA′′′=τAwithA(0)=0,A(0)=1,A′′(0)=μ.6AA^{\prime\prime\prime}=\tau A^{\prime}\quad{\textrm{with}}\quad A(0)=0,\ A^{\prime}(0)=-1,\ A^{\prime\prime}(0)=\mu.

Our strategy will be as follows: first we will show that the IVP (7) has a unique solution, which depends on μ\mu continuously. Next we will show that for μ\mu sufficiently large the solution of (7) goes to zero in finite time. Then we will show that for μ=0\mu=0 there is a smallest time 0<τ2<0<\tau_{2}<\infty such that the solution of (7) has its first local maximum at τ2\tau_{2} and remains strictly negative for all τ(0,τ2]\tau\in(0,\tau_{2}]. See Figure 2.

Refer to caption
Figure 2. Numerically computed solutions of (7) for increasing values of μ\mu.

Then an open/closed argument will give the special solution AA_{*} for a choice μ=μ>0\mu=\mu_{*}>0.

The differential equation (4) is nonlinear, nonautonomous and singular at A=0A=0. As such, carrying out the details of our strategy is quite challenging. The key observation for proving Proposition 1 is that (4) is, in a loose sense, close to being hamiltonian. Specifically, one can rewrite the right hand side of (4) as τ6ddτln|A(τ)|\displaystyle\frac{\tau}{6}\frac{d}{d\tau}\ln\left|A(\tau)\right|. If we then consider the related equation

(8) B′′′=ddτln|B(τ)|B^{\prime\prime\prime}=\frac{d}{d\tau}\ln\left|B(\tau)\right|

obtained from (4) by replacing the nonautonomous τ/6\tau/6 with the constant 11, one sees that this equation is in fact hamiltonian. Specifically one can integrate both sides to find:

B′′(τ)=ddB((B′′(0)1)B+Bln|B|)B^{\prime\prime}(\tau)=\frac{d}{dB}\left(\left(B^{\prime\prime}(0)-1\right)B+B\ln\left|B\right|\right)

The conserved energy is

12(B)2((B′′(0)1)B+Bln|B|)\displaystyle{1\over 2}(B^{\prime})^{2}-\left(\left(B^{\prime\prime}(0)-1\right)B+B\ln\left|B\right|\right)

and most of the dynamics of (8) can be determined by analyzing this quantity. In the following proof of Proposition 1, we will frequently make use of quantities similar to this energy which, though not conserved for solutions of (4), do provide quite a bit of dynamical information about solutions.

3.1. Well-posedness of (7)

Now we prove (7) has unique solutions and that such solutions depend smoothly upon μ\mu.

Lemma 2.

For any μ0𝐑\mu_{0}\in{\bf{R}}, there exist τ0>0\tau_{0}>0 and δ>0\delta>0 such that for each μ(μ0δ,μ0+δ)\mu\in(\mu_{0}-\delta,\mu_{0}+\delta) the IVP (7) has a unique solution A:[0,τ0]𝐑A:[0,\tau_{0}]\rightarrow{\bf{R}}, which satisfies A(τ)<0A(\tau)<0 for all τ(0,τ0]\tau\in(0,\tau_{0}]. In addition, there exists C>0C>0 such that for any μ1,μ2(μ0δ,μ0+δ)\mu_{1},\mu_{2}\in(\mu_{0}-\delta,\mu_{0}+\delta) the solutions A1(τ)A_{1}(\tau) and A2(τ)A_{2}(\tau) to the IVP (7) with A1′′(0)=μ1A^{\prime\prime}_{1}(0)=\mu_{1} and A2′′(0)=μ2A^{\prime\prime}_{2}(0)=\mu_{2}, respectively, satisfy

|A2(τ)A1(τ)|+|A2(τ)A1(τ)|+|A2′′(τ)A1′′(τ)|C|μ2μ1|\left|A_{2}(\tau)-A_{1}(\tau)\right|+\left|A^{\prime}_{2}(\tau)-A^{\prime}_{1}(\tau)\right|+\left|A^{\prime\prime}_{2}(\tau)-A^{\prime\prime}_{1}(\tau)\right|\leq C|\mu_{2}-\mu_{1}|

for all τ[0,τ0]\tau\in[0,\tau_{0}].

Proof.

Letting F(A,A,A′′,τ):=τA6AF(A,A^{\prime},A^{\prime\prime},\tau):=\displaystyle\frac{\tau A^{\prime}}{6A}, (4) can be written as A′′′=F(A,A,A′′,τ)A^{\prime\prime\prime}=F(A,A^{\prime},A^{\prime\prime},\tau). If F(A,A,A′′,τ)F(A,A^{\prime},A^{\prime\prime},\tau) were continuous in τ\tau and Lipschitz in (A,A,A′′)(A,A^{\prime},A^{\prime\prime}) throughout a neighborhood of (A,A,A′′,τ)=(0,1,μ,0)(A,A^{\prime},A^{\prime\prime},\tau)=(0,-1,\mu,0), then Lemma 2 would be a trivial consequence of the standard theory on existence and uniqueness of solutions and their continuous dependence on initial conditions. Unfortunately, in our case F(A,A,A′′,τ)F(A,A^{\prime},A^{\prime\prime},\tau) is not even defined if A=0A=0. However, notice that if there exists a solution A(τ)A(\tau) to the IVP (7) then L’Hopital’s rule implies:

limτ0+τA(τ)6A(τ)=16.\lim_{\tau\rightarrow 0^{+}}\displaystyle\frac{\tau A^{\prime}(\tau)}{6A(\tau)}=\displaystyle\frac{1}{6}.

Therefore our strategy will be to remove the singularity at (A,τ)=(0,0)(A,\tau)=(0,0) by modifying FF in such a way that the resulting modification is continuous in a neighborhood of the initial data and agrees with FF for solutions of (7)\eqref{IVP}.

Fix μ0𝐑\mu_{0}\in{\bf{R}} and k>0k>0, and let

p(τ):=τ+12(μ0k)τ2andq(τ):=τ+12(μ0+k)τ2.p(\tau):=-\tau+\displaystyle\frac{1}{2}(\mu_{0}-k)\tau^{2}\quad\textrm{and}\quad q(\tau):=-\tau+\displaystyle\frac{1}{2}(\mu_{0}+k)\tau^{2}.

Given the initial conditions in (7), p(τ)p(\tau) and q(τ)q(\tau) formally give crude lower and upper bounds on A(τ)A(\tau) for τ\tau close to 0. Now construct the following function:

F~(A,A,A′′,τ):={τA6p(τ)if τ>0 and Ap(τ),τA6Aif τ>0 and p(τ)<A<q(τ),τA6q(τ)if τ>0 and q(τ)A,A6if τ0.\tilde{F}(A,A^{\prime},A^{\prime\prime},\tau):=\begin{cases}\displaystyle\frac{\tau A^{\prime}}{6p(\tau)}&\text{if $\tau>0$ and $A\leq p(\tau)$},\\ \displaystyle\frac{\tau A^{\prime}}{6A}&\text{if $\tau>0$ and $p(\tau)<A<q(\tau)$},\\ \displaystyle\frac{\tau A^{\prime}}{6q(\tau)}&\text{if $\tau>0$ and $q(\tau)\leq A$},\\ -\displaystyle\frac{A^{\prime}}{6}&\text{if $\tau\leq 0$}.\\ \end{cases}

F~\tilde{F} is well-defined and continuous throughout a small neighborhood of (A,A,A′′,τ)=(0,1,μ,0)(A,A^{\prime},A^{\prime\prime},\tau)=(0,-1,\mu,0) for any μ𝐑\mu\in{\bf{R}}. Then we have the existence of a solution A(τ)A(\tau) to the IVP:

A′′′=F~(A,A,A′′,τ)withA(0)=0,A(0)=1,A′′(0)=μ.A^{\prime\prime\prime}=\tilde{F}(A,A^{\prime},A^{\prime\prime},\tau)\quad{\textrm{with}}\quad A(0)=0,\ A^{\prime}(0)=-1,\ A^{\prime\prime}(0)=\mu.

Furthermore, it is easy to verify that there exist τ0>0\tau_{0}>0 and 0<δ<k0<\delta<k such that for each μ(μ0δ,μ0+δ)\mu\in(\mu_{0}-\delta,\mu_{0}+\delta) the solution A(τ)A(\tau) is defined for all τ[0,τ0]\tau\in[0,\tau_{0}] and satisfies p(τ)<A(τ)<q(τ)<0p(\tau)<A(\tau)<q(\tau)<0 for all τ(0,τ0]\tau\in(0,\tau_{0}]. It follows that the restriction of A(τ)A(\tau) to [0,τ0][0,\tau_{0}] is in fact a solution to the IVP (7).

Note that along any solution of (7), A(τA6A)\displaystyle\frac{\partial}{\partial A}\left(\displaystyle\frac{\tau A^{\prime}}{6A}\right)\rightarrow\infty as τ0+\tau\rightarrow 0^{+}. Consequently, it is not possible to modify FF to a function that is Lipschitz in AA in a neighborhood of (A,A,A′′,τ)=(0,1,μ,0)(A,A^{\prime},A^{\prime\prime},\tau)=(0,-1,\mu,0). Therefore, we have to establish the uniqueness of the solution by other means, namely the nearly hamiltonian structure of (4).

For any μ1,μ2(μ0δ,μ0+δ)\mu_{1},\mu_{2}\in(\mu_{0}-\delta,\mu_{0}+\delta), let A1(τ)A_{1}(\tau) and A2(τ)A_{2}(\tau) be solutions to the IVP (7) with A1′′(0)=μ1A^{\prime\prime}_{1}(0)=\mu_{1} and A2′′(0)=μ2A^{\prime\prime}_{2}(0)=\mu_{2}, respectively. Recall that our choice of δ\delta and τ0\tau_{0} ensures p(τ)<A1,2(τ)<q(τ)<0p(\tau)<A_{1,2}(\tau)<q(\tau)<0 for all τ(0,τ0]\tau\in(0,\tau_{0}]. Thus we can take a sufficiently small T(0,min{1,τ0}]T\in(0,\min\{1,\tau_{0}\}], independent of μ1\mu_{1} and μ2\mu_{2}, such that A1,2(τ)<12τ<0A_{1,2}(\tau)<-\displaystyle\frac{1}{2}\tau<0 for all τ(0,T]\tau\in(0,T] and

M:=maxτ(0,T]|A2(τ)A1(τ)1|<12.M:=\max_{\tau\in(0,T]}\left|\displaystyle\frac{A_{2}(\tau)}{A_{1}(\tau)}-1\right|<\displaystyle\frac{1}{2}.

Then for any τ(0,T]\tau\in(0,T],

(9) |ln(A2(τ)A1(τ))|M1M2M,\left|\ln{\left(\displaystyle\frac{A_{2}(\tau)}{A_{1}(\tau)}\right)}\right|\leq\displaystyle\frac{M}{1-M}\leq 2M,

where the equalities hold if M=0M=0.

Since both A1A_{1} and A2A_{2} are solutions to the ODE, we have

A2′′′A1′′′=τ6(A2A2A1A1)=τ6(A1A2)(A2A1).A^{\prime\prime\prime}_{2}-A^{\prime\prime\prime}_{1}=\displaystyle\frac{\tau}{6}\left(\displaystyle\frac{A^{\prime}_{2}}{A_{2}}-\displaystyle\frac{A^{\prime}_{1}}{A_{1}}\right)=\displaystyle\frac{\tau}{6}\left(\displaystyle\frac{A_{1}}{A_{2}}\right)\left(\displaystyle\frac{A_{2}}{A_{1}}\right)^{\prime}.

Integrating the above equation gives

A2′′(τ)A1′′(τ)=τ6ln(A2(τ)A1(τ))160τln(A2(s)A1(s))𝑑s+δμ,A^{\prime\prime}_{2}(\tau)-A^{\prime\prime}_{1}(\tau)=\displaystyle\frac{\tau}{6}\ln{\left(\displaystyle\frac{A_{2}(\tau)}{A_{1}(\tau)}\right)}-\displaystyle\frac{1}{6}\int_{0}^{\tau}\!\ln{\left(\displaystyle\frac{A_{2}(s)}{A_{1}(s)}\right)}\,ds+\delta\mu,

where δμ:=μ2μ1\delta\mu:=\mu_{2}-\mu_{1}. Then by incorporating (9), we obtain that for all τ(0,T]\tau\in(0,T],

(10) |A2′′(τ)A1′′(τ)|23Mτ+|δμ|.\left|A^{\prime\prime}_{2}(\tau)-A^{\prime\prime}_{1}(\tau)\right|\leq\displaystyle\frac{2}{3}M\tau+|\delta\mu|.

Together with the initial conditions A1(0)=A2(0)=0A_{1}(0)=A_{2}(0)=0 and A1(0)=A2(0)=1A^{\prime}_{1}(0)=A^{\prime}_{2}(0)=-1, (10) implies that for all τ(0,T]\tau\in(0,T],

|A2(τ)A1(τ)|19Mτ3+12|δμ|τ2.\left|A_{2}(\tau)-A_{1}(\tau)\right|\leq\displaystyle\frac{1}{9}M\tau^{3}+\displaystyle\frac{1}{2}|\delta\mu|\tau^{2}.

Recall Tmin{1,τ0}T\leq\min\{1,\tau_{0}\}. It follows that for all τ(0,T]\tau\in(0,T],

|A2(τ)A1(τ)1|\displaystyle\left|\displaystyle\frac{A_{2}(\tau)}{A_{1}(\tau)}-1\right|\leq{} 19Mτ3+12|δμ|τ2|A1(τ)|19Mτ3+12|δμ|τ212τ=29Mτ2+|δμ|τ\displaystyle\displaystyle\frac{\displaystyle\frac{1}{9}M\tau^{3}+\displaystyle\frac{1}{2}|\delta\mu|\tau^{2}}{|A_{1}(\tau)|}\leq\displaystyle\frac{\displaystyle\frac{1}{9}M\tau^{3}+\displaystyle\frac{1}{2}|\delta\mu|\tau^{2}}{\displaystyle\frac{1}{2}\tau}=\displaystyle\frac{2}{9}M\tau^{2}+|\delta\mu|\tau
\displaystyle\leq{} 29MT2+|δμ|T29M+|δμ|.\displaystyle\displaystyle\frac{2}{9}MT^{2}+|\delta\mu|T\leq\displaystyle\frac{2}{9}M+|\delta\mu|.

By the definition of MM, we have M29M+|δμ|M\leq\displaystyle\frac{2}{9}M+|\delta\mu|, which implies

(11) M97|δμ|.M\leq\displaystyle\frac{9}{7}|\delta\mu|.

When δμ=0\delta\mu=0 (i.e., μ1=μ2\mu_{1}=\mu_{2}), Inequality (11) requires M=0M=0. Consequently, A2(τ)=A1(τ)A_{2}(\tau)=A_{1}(\tau) for all τ[0,T]\tau\in[0,T]. This proves the uniqueness of the solution on the interval [0,T][0,T]. Recall that with μ1(μ0δ,μ0+δ)\mu_{1}\in(\mu_{0}-\delta,\mu_{0}+\delta), A1(τ)<0A_{1}(\tau)<0 for all τ[T,τ0]\tau\in[T,\tau_{0}] by our choice of τ0\tau_{0} and δ\delta. Then A′′′=τA6AA^{\prime\prime\prime}=\displaystyle\frac{\tau A^{\prime}}{6A} with the initial conditions A(T)=A1(T)<0A(T)=A_{1}(T)<0, A(T)=A1(T)A^{\prime}(T)=A^{\prime}_{1}(T), A′′(T)=A1′′(T)A^{\prime\prime}(T)=A^{\prime\prime}_{1}(T) constitutes a regular initial value problem, to which A1(τ)A_{1}(\tau) is the unique solution throughout the interval [T,τ0][T,\tau_{0}] according to the standard existence and uniqueness theory. Altogether, we have shown the uniqueness of the solution on the full interval [0,τ0][0,\tau_{0}].

Since the choice of TT is independent of μ1\mu_{1} and μ2\mu_{2}, after combining (10) and (11) and integrating we can obtain a uniform constant CT>0C_{T}>0 such that for any μ1,μ2(μ0δ,μ0+δ)\mu_{1},\mu_{2}\in(\mu_{0}-\delta,\mu_{0}+\delta),

|A2(τ)A1(τ)|+|A2(τ)A1(τ)|+|A2′′(τ)A1′′(τ)|CT|μ2μ1|\left|A_{2}(\tau)-A_{1}(\tau)\right|+\left|A^{\prime}_{2}(\tau)-A^{\prime}_{1}(\tau)\right|+\left|A^{\prime\prime}_{2}(\tau)-A^{\prime\prime}_{1}(\tau)\right|\leq C_{T}|\mu_{2}-\mu_{1}|

for all τ[0,T]\tau\in[0,T]. Since A1,2(τ)<q(τ)<0A_{1,2}(\tau)<q(\tau)<0 for all τ[T,τ0]\tau\in[T,\tau_{0}], by applying Gronwall’s inequality on the interval [T,τ0][T,\tau_{0}] we can extend the above inequality to the full interval [0,τ0][0,\tau_{0}] after replacing CTC_{T} with some larger C>0C>0. ∎

Let A0(τ)A_{0}(\tau) be the solution to the IVP (7) with A0′′(0)=μ0𝐑A^{\prime\prime}_{0}(0)=\mu_{0}\in{\bf{R}}. Clearly, we can extend A0(τ)A_{0}(\tau) to any τ¯>τ0\bar{\tau}>\tau_{0} as long as A0(τ)<0A_{0}(\tau)<0 for all τ(0,τ¯]\tau\in(0,\bar{\tau}]. Furthermore, by considering A′′′=τA6AA^{\prime\prime\prime}=\displaystyle\frac{\tau A^{\prime}}{6A} with the initial conditions specified at τ0\tau_{0} and set close to (A0(τ0),A0(τ0),A0′′(τ0))(A_{0}(\tau_{0}),A^{\prime}_{0}(\tau_{0}),A^{\prime\prime}_{0}(\tau_{0})), we can extend Lemma 2 to the interval [0,τ¯][0,\bar{\tau}]. In particular, by applying Gronwall’s inequality on the interval [τ0,τ¯][\tau_{0},\bar{\tau}], we obtain the following corollary.

Corollary 3.

For any τ¯>0\bar{\tau}>0 such that A0(τ)<0A_{0}(\tau)<0 for all τ(0,τ¯]\tau\in(0,\bar{\tau}], there exist C¯>0\bar{C}>0 and a sufficiently small δ¯>0\bar{\delta}>0 such that for any μ(μ0δ¯,μ0+δ¯)\mu\in(\mu_{0}-\bar{\delta},\mu_{0}+\bar{\delta}) the solution A(τ)A(\tau) with A′′(0)=μA^{\prime\prime}(0)=\mu remains negative for all τ(0,τ¯]\tau\in(0,\bar{\tau}] and satisfies

(12) |A(τ)A0(τ)|+|A(τ)A0(τ)|+|A′′(τ)A0′′(τ)|C¯|μμ0|\left|A(\tau)-A_{0}(\tau)\right|+\left|A^{\prime}(\tau)-A^{\prime}_{0}(\tau)\right|+\left|A^{\prime\prime}(\tau)-A^{\prime\prime}_{0}(\tau)\right|\leq\bar{C}|\mu-\mu_{0}|

for all τ[0,τ¯]\tau\in[0,\bar{\tau}].

3.2. General behavior of solutions

Now that we know (7) is well-posed, we demonstrate some properties of solutions when μ0\mu\geq 0. First, its solutions reach a first minimum in finite time. Let

(13) τ1:=sup{τ>0:sup0<s<τA(s)<0}.\displaystyle\tau_{1}:=\sup\left\{\tau>0:\sup_{0<s<\tau}A^{\prime}(s)<0\right\}.

Note that τ1\tau_{1} depends upon μ\mu, though we supress this for notational simplicity.

Lemma 4.

For any μ0\mu\geq 0 we have 0<τ1<0<\tau_{1}<\infty. In addition, A(τ1)=0A^{\prime}(\tau_{1})=0 and A′′(τ1)>0A^{\prime\prime}(\tau_{1})>0. Moreover there exist μ0,a0,a1>0\mu_{0},a_{0},a_{1}>0 such that μμ0\mu\geq\mu_{0} implies (a) 12μτ11μ\displaystyle\frac{1}{2\mu}\leq\tau_{1}\leq\displaystyle\frac{1}{\mu}, (b) a0μA(τ1)a1μ-\displaystyle\frac{a_{0}}{\mu}\leq A(\tau_{1})\leq-\frac{a_{1}}{\mu} and (c) μA′′(τ1)2μ\mu\leq A^{\prime\prime}(\tau_{1})\leq 2\mu.

Proof.

Clearly, A(τ1)=0A^{\prime}(\tau_{1})=0 if τ1\tau_{1} is finite. For τI1:=(0,τ1)\tau\in I_{1}:=(0,\tau_{1}) (possibly τ1=\tau_{1}=\infty), AA and AA^{\prime} are negative and consequently A′′′>0A^{\prime\prime\prime}>0. Thus A′′A^{\prime\prime} is increasing. This implies that A′′(τ)>μ0A^{\prime\prime}(\tau)>\mu\geq 0 for τI1\tau\in I_{1} and also A′′(τ1)>0A^{\prime\prime}(\tau_{1})>0 if τ1\tau_{1} is finite. Thus AA is concave up on I1I_{1}. This together with the fact that A(0)=0A(0)=0 implies via the mean value theorem that |A(τ)/τ|>|A(τ)||A(\tau)/\tau|>|A^{\prime}(\tau)| for all τI1\tau\in I_{1} and so τA/A<1\tau A^{\prime}/A<1. Thus we have

(14) 0<A′′′<1/60<A^{\prime\prime\prime}<1/6

for τI1\tau\in I_{1}. Integrating this inequality twice gives μτ1Aτ2/12+μτ1\mu\tau-1\leq A^{\prime}\leq\tau^{2}/12+\mu\tau-1 for all τI1\tau\in I_{1}. If μ>0\mu>0, the lower bound on AA^{\prime} becomes positive in finite time, implying that τ1<\tau_{1}<\infty. Examination of the zeroes of the upper and lower bounds gives the estimates for τ1\tau_{1} stated in the lemma. Further integration of (14) gives the estimates on AA and A′′A^{\prime\prime}. If μ=0\mu=0, the lower bound on AA^{\prime} is not sufficient to conclude that τ1<\tau_{1}<\infty. However the fact that A′′′(0)=1/6A^{\prime\prime\prime}(0)=1/6 can be used to improve the lower bound in this case. The details are standard and omitted.

After reaching their first local minimum at τ=τ1\tau=\tau_{1}, solutions AA either reach zero in finite time or alternately reach a local maximum (for which A<0A<0) in finite time, as we now demonstrate. Let

(15) τ2:=sup{τ>τ1:infτ1<s<τA(s)0 and supτ1<s<τA(s)<0}.\displaystyle\tau_{2}:=\sup\left\{\tau>\tau_{1}:\inf_{\tau_{1}<s<\tau}A^{\prime}(s)\geq 0\ {\textrm{ and }}\sup_{\tau_{1}<s<\tau}A(s)<0\right\}.

As with τ1\tau_{1}, τ2\tau_{2} implicitly depends upon μ\mu. If this is finite, then we have either A(τ2)=0A(\tau_{2})=0 and A(τ2)0A^{\prime}(\tau_{2})\geq 0 or A(τ2)<0A(\tau_{2})<0 and A(τ2)=0A^{\prime}(\tau_{2})=0.

Lemma 5.

For all μ0\mu\geq 0 we have τ1<τ2<.\tau_{1}<\tau_{2}<\infty.

Proof.

Clearly τ2>τ1\tau_{2}>\tau_{1} since A′′(τ1)>0A^{\prime\prime}(\tau_{1})>0. Suppose that τ2=\tau_{2}=\infty. Since AA is bounded above and nondecreasing for τ>τ1\tau>\tau_{1}, we have A(τ)A(\tau) convergent as τ\tau\to\infty. Since we have A<0A<0 and A0A^{\prime}\geq 0 for all τ>τ1\tau>\tau_{1}, (7) implies A′′′0A^{\prime\prime\prime}\leq 0 for all τ>τ1\tau>\tau_{1} as well. Thus A′′A^{\prime\prime} is nonincreasing. Since it is initially positive and AA converges, clearly A′′A^{\prime\prime} must eventually become negative. Since A′′A^{\prime\prime} is nonincreasing, once it is negative it remains negative for all subsequent times. Thus there is an ε>0\varepsilon>0 and a value τ3>τ1\tau_{3}>\tau_{1} such that A′′(τ)<εA^{\prime\prime}(\tau)<-\varepsilon for all τ>τ3\tau>\tau_{3}. Integrating this, however, indicates that AA^{\prime} must eventually become negative, a contradiction, and the lemma is shown. ∎

3.3. Energy estimates and behavior of AA for μ=0\mu=0 or μ\mu large

Our next goal is to show that for μ\mu large, solutions go to zero in finite time but when μ=0\mu=0 the solution has a local negative maximum at τ2\tau_{2}. This requires the following energy estimates, similar to those for (8):

Lemma 6.

Fix μ\mu and take τ1\tau_{1} and τ2\tau_{2} as above. Let

V1(a):=τ16aln(aA(τ1))+(A′′(τ1)τ16)(aA(τ1)).V_{1}(a):={\tau_{1}\over 6}a\ln\left({a\over A(\tau_{1})}\right)+\left(A^{\prime\prime}(\tau_{1})-{\tau_{1}\over 6}\right)\left(a-A(\tau_{1})\right).

Then for all τ1<ττ2\tau_{1}<\tau\leq\tau_{2}

12[A(τ)]2<V1(A(τ)).{1\over 2}[A^{\prime}(\tau)]^{2}<V_{1}(A(\tau)).

Moreover, take τ¯:=6A′′(τ1)\bar{\tau}:=6A^{\prime\prime}(\tau_{1}) and set

V2(a):=τ¯6aln(aA(τ1)).V_{2}(a):={\bar{\tau}\over 6}a\ln\left({a\over A(\tau_{1})}\right).

If τ¯τ2\bar{\tau}\geq\tau_{2} then for all τ1<ττ2\tau_{1}<\tau\leq\tau_{2}

V2(A(τ))<12[A(τ)]2.V_{2}(A(\tau))<{1\over 2}[A^{\prime}(\tau)]^{2}.
Proof.

Let I2:=(τ1,τ2)I_{2}:=(\tau_{1},\tau_{2}). By definition of τ2\tau_{2}, we have A(τ)<0A(\tau)<0 and A(τ)0A^{\prime}(\tau)\geq 0 for all τI2\tau\in I_{2}. If τ¯τ2\bar{\tau}\geq\tau_{2} then we have for all τI2\tau\in I_{2},

τ¯A(τ)6A(τ)A′′′(τ)τ1A(τ)6A(τ),{\bar{\tau}A^{\prime}(\tau)\over 6A(\tau)}\leq A^{\prime\prime\prime}(\tau)\leq{\tau_{1}A^{\prime}(\tau)\over 6A(\tau)},

where the equalities hold only if A(τ)=0A^{\prime}(\tau)=0 and the second inequality is true independent of τ¯\bar{\tau}. Integrating this from τ1\tau_{1} to τ\tau gives

A′′(τ1)+τ¯6ln(A(τ)A(τ1))<A′′(τ)<A′′(τ1)+τ16ln(A(τ)A(τ1))A^{\prime\prime}(\tau_{1})+{\bar{\tau}\over 6}\ln\left({A(\tau)\over A(\tau_{1})}\right)<A^{\prime\prime}(\tau)<A^{\prime\prime}(\tau_{1})+{\tau_{1}\over 6}\ln\left({A(\tau)\over A(\tau_{1})}\right)

for all τI2\tau\in I_{2}. Note that we obtain strict inequalities here because with A′′(τ1)>0A^{\prime\prime}(\tau_{1})>0 there is a small α>0\alpha>0 such that A(τ)>0A^{\prime}(\tau)>0 for all τ(τ1,τ1+α)\tau\in(\tau_{1},\tau_{1}+\alpha). If we multiply by A(τ)A^{\prime}(\tau) and integrate once more from τ1\tau_{1} to any τ(τ1,τ2]\tau\in(\tau_{1},\tau_{2}] then the conclusions of the lemma follow. In particular, τ¯\bar{\tau} is irrelevant to the upper bound V1V_{1}. ∎

Lemma 7.

There exists M>0M>0 such that μM\mu\geq M implies A(τ2)=0A(\tau_{2})=0 and A(τ2)>0A^{\prime}(\tau_{2})>0.

Proof.

First we show that if τ¯τ2\bar{\tau}\geq\tau_{2} then we are done. We know by the definition of τ2\tau_{2} that either A(τ2)=0A^{\prime}(\tau_{2})=0 or A(τ2)=0A(\tau_{2})=0. Inspection of V2(a)V_{2}(a) demonstrates that V2(a)>0V_{2}(a)>0 for a(A(τ1),0)a\in(A(\tau_{1}),0) and is zero at the ends of that interval (see Figure 3). Thus, since AA is increasing on (τ1,τ2)(\tau_{1},\tau_{2}) we have A(τ2)(A(τ1),0]A(\tau_{2})\in(A(\tau_{1}),0] and therefore V2(A(τ2))0V_{2}(A(\tau_{2}))\geq 0. We also know for τ(τ1,τ2)\tau\in(\tau_{1},\tau_{2}) that A0A^{\prime}\geq 0. So in particular Lemma 6 implies A(τ2)>2V2(A(τ2))0A^{\prime}(\tau_{2})>\sqrt{2V_{2}(A(\tau_{2}))}\geq 0. So, since A(τ2)0A^{\prime}(\tau_{2})\neq 0 it must be the case that A(τ2)=0A(\tau_{2})=0.

Refer to caption
Figure 3. Sketch of A(τ)A^{\prime}(\tau) vs A(τ)A(\tau) when μ\mu is large.

On the other hand perhaps τ2>τ¯=6A′′(τ1)\tau_{2}>\bar{\tau}=6A^{\prime\prime}(\tau_{1}). We will prove by contradiction that this is impossible when μ\mu is large. The estimate in the proof of Lemma 6 gives:

1A(τ)2V2(A(τ))1\leq{A^{\prime}(\tau)\over\sqrt{2V_{2}(A(\tau))}}

which is valid for τ1<ττ¯\tau_{1}<\tau\leq\bar{\tau} if we assume τ¯<τ2\bar{\tau}<\tau_{2}. (Note that for μ\mu big enough, the estimates on τ1\tau_{1} in Lemma 4 imply that τ¯>τ1\bar{\tau}>\tau_{1}.) Integration from τ1\tau_{1} to τ¯\bar{\tau} gives

τ¯τ1τ1τ¯A(s)ds2V2(A(s))=A(τ1)A(τ¯)da2V2(a).\bar{\tau}-\tau_{1}\leq\int_{\tau_{1}}^{\bar{\tau}}{A^{\prime}(s)ds\over\sqrt{2V_{2}(A(s))}}=\int_{A(\tau_{1})}^{A(\bar{\tau})}{da\over\sqrt{2V_{2}(a)}}.

Substitution in from the definition of V2V_{2} and τ¯\bar{\tau} gives

6A′′(τ1)τ112A′′(τ1)A(τ1)A(τ¯)daaln(a/A(τ1)).6A^{\prime\prime}(\tau_{1})-\tau_{1}\leq{1\over\sqrt{2A^{\prime\prime}(\tau_{1})}}\int_{A(\tau_{1})}^{A(\bar{\tau})}{da\over\sqrt{a\ln\left({a/A(\tau_{1})}\right)}}.

A change of variables and the fact that 1>A(τ¯)/A(τ1)>01>A(\bar{\tau})/A(\tau_{1})>0 implies:

6A′′(τ1)τ1|A(τ1)|2A′′(τ1)A(τ¯)/A(τ1)1db|bln(b)||A(τ1)|2A′′(τ1)01db|bln(b)|.6A^{\prime\prime}(\tau_{1})-\tau_{1}\leq\sqrt{|A(\tau_{1})|\over 2A^{\prime\prime}(\tau_{1})}\int^{1}_{A(\bar{\tau})/A(\tau_{1})}{db\over\sqrt{\left|b\ln(b)\right|}}\leq\sqrt{|A(\tau_{1})|\over 2A^{\prime\prime}(\tau_{1})}\int^{1}_{0}{db\over\sqrt{\left|b\ln(b)\right|}}.

Letting K=01db|bln(b)|<\displaystyle K=\int^{1}_{0}{db\over\sqrt{\left|b\ln(b)\right|}}<\infty we have

6A′′(τ1)τ1K|A(τ1)|2A′′(τ1).6A^{\prime\prime}(\tau_{1})-\tau_{1}\leq K\sqrt{|A(\tau_{1})|\over 2A^{\prime\prime}(\tau_{1})}.

Now we employ the estimates in Lemma 4 to find for all μμ0\mu\geq\mu_{0}:

6μ1μKa0/μ2μ=Kμa02.6\mu-{1\over\mu}\leq K\sqrt{{a_{0}/\mu\over 2\mu}}={K\over\mu}\sqrt{a_{0}\over 2}.

For μ\mu sufficiently large this is impossible and so the lemma is shown.

Lemma 8.

If μ=0\mu=0, then A(τ2)<0A(\tau_{2})<0 and A(τ2)=0A^{\prime}(\tau_{2})=0.

Refer to caption
Figure 4. Sketch of A(τ)A^{\prime}(\tau) vs A(τ)A(\tau) when μ=0\mu=0.
Proof.

Proving the lemma amounts to showing that A(τ2)0A(\tau_{2})\neq 0. Notice that V1(0)=A(τ1)(τ1/6A′′(τ1))V_{1}(0)=A(\tau_{1})\left(\tau_{1}/6-A^{\prime\prime}(\tau_{1})\right). Let ρ(τ)=τ/6A′′(τ)\rho(\tau)=\tau/6-A^{\prime\prime}(\tau). ρ(0)=0\rho(0)=0, since μ=0\mu=0. Also ρ(τ)=1/6A′′′(τ)\rho^{\prime}(\tau)=1/6-A^{\prime\prime\prime}(\tau). In the proof of Lemma 4 we saw that A′′′<1/6A^{\prime\prime\prime}<1/6 for τI1\tau\in I_{1} (see (14)) and so ρ>0\rho^{\prime}>0. Consequently τ1/6A′′(τ1)>0\tau_{1}/6-A^{\prime\prime}(\tau_{1})>0. Thus, since A(τ1)<0A(\tau_{1})<0, we have V1(0)<0.V_{1}(0)<0. Moreover, V1(A(τ1))=0V_{1}(A(\tau_{1}))=0 and V1(A(τ1))=A′′(τ1)>0V^{\prime}_{1}(A(\tau_{1}))=A^{\prime\prime}(\tau_{1})>0. Thus there is a point k(A(τ1),0)k\in(A(\tau_{1}),0) at which V1(k)=0V_{1}(k)=0. See Figure 4.

Since for τ[τ1,τ2]\tau\in[\tau_{1},\tau_{2}] we have 0A(τ)A(τ1)0\geq A(\tau)\geq A(\tau_{1}) and 0A(τ)2V1(A(τ))0\leq A^{\prime}(\tau)\leq\sqrt{2V_{1}(A(\tau))} we must therefore have 0>kA(τ)0>k\geq A(\tau). Since τ2\tau_{2} is finite for μ=0\mu=0 and AA is bounded from zero for all τ1<τ<τ2\tau_{1}<\tau<\tau_{2}, the definition (15) of τ2\tau_{2} requires A(τ2)=0A^{\prime}(\tau_{2})=0.

3.4. Openness of qualitative behavior

The next pair of results show that slightly varying μ\mu does not change the qualitative behavior of AA at τ2\tau_{2}. That is to say, if μ0\mu_{0} is such that A(τ2)=0A(\tau_{2})=0 then there is an open neighborhood of μ0\mu_{0} where the corresponding solutions enjoy this same property. Likewise, if μ0\mu_{0} such that A(τ2)=0A^{\prime}(\tau_{2})=0, then there is an open neighborhood of μ0\mu_{0} where the corresponding solutions do the same. Let A^(τ)\widehat{A}(\tau) be the solution to the IVP (7) with A^′′(0)=μ^0\widehat{A}^{\prime\prime}(0)=\hat{\mu}\geq 0, and define τ^2\widehat{\tau}_{2} for A^(τ)\widehat{A}(\tau) as in (15).

Lemma 9.

Suppose A^(τ^2)=0\widehat{A}(\widehat{\tau}_{2})=0 and A^(τ^2)=2h>0\widehat{A}^{\prime}(\widehat{\tau}_{2})=2h>0. Then for any ϵ>0\epsilon>0 there exists a δ>0\delta>0 such that for any μ(μ^δ,μ^+δ)\mu\in(\hat{\mu}-\delta,\hat{\mu}+\delta), the solution A(τ)A(\tau) to the IVP (7) with A′′(0)=μA^{\prime\prime}(0)=\mu reaches zero at τ2<τ^2+ϵ\tau_{2}<\widehat{\tau}_{2}+\epsilon.

Proof.

Since A^(τ^2)=0\widehat{A}(\widehat{\tau}_{2})=0, the right hand side of (4) is undefined at this point and so we cannot directly use our results about continuous dependence on μ\mu from Corollary 3 to prove this result. Instead we will have to rely instead on the nearly hamiltonian structure of (4).

Integrating the ODE of (7) with the initial conditions A(0)=0A(0)=0, A(0)=1A^{\prime}(0)=-1, and A′′(0)=μA^{\prime\prime}(0)=\mu, we obtain that

(16) A(τ)A′′(τ)12(A(τ))2+12=16τA(τ)160τA(s)𝑑sA(\tau)A^{\prime\prime}(\tau)-\displaystyle\frac{1}{2}\left(A^{\prime}(\tau)\right)^{2}+\displaystyle\frac{1}{2}=\displaystyle\frac{1}{6}\tau A(\tau)-\displaystyle\frac{1}{6}\int_{0}^{\tau}A(s)\,ds

as long as A<0A<0 on the interval (0,τ)(0,\tau).

Since A^(τ^2)=0\widehat{A}(\widehat{\tau}_{2})=0 and A^(τ^2)=2h>0\widehat{A}^{\prime}(\widehat{\tau}_{2})=2h>0, we can choose a small α>0\alpha>0 such that |A^(τ)||h(ττ^2)||\widehat{A}(\tau)|\geq|h(\tau-\widehat{\tau}_{2})| and A^(τ)<3h\widehat{A}^{\prime}(\tau)<3h for all τ[τ^2α,τ^2]\tau\in[\widehat{\tau}_{2}-\alpha,\widehat{\tau}_{2}]. It follows that

|A^′′(τ)A^′′(τ^2α)|τ^2ατ|sA^(s)6A^(s)|𝑑sτ^2ατ|τ^23h6h(sτ^2)|𝑑s=12τ^2lnατ^2τ.\left|\widehat{A}^{\prime\prime}(\tau)-\widehat{A}^{\prime\prime}(\widehat{\tau}_{2}-\alpha)\right|\leq\int_{\widehat{\tau}_{2}-\alpha}^{\tau}\left|\displaystyle\frac{s\widehat{A}^{\prime}(s)}{6\widehat{A}(s)}\right|ds\leq\int_{\widehat{\tau}_{2}-\alpha}^{\tau}\left|\displaystyle\frac{\widehat{\tau}_{2}3h}{6h(s-\widehat{\tau}_{2})}\right|ds=\displaystyle\frac{1}{2}\widehat{\tau}_{2}\ln{\displaystyle\frac{\alpha}{\widehat{\tau}_{2}-\tau}}.

Therefore, A^(τ)A^′′(τ)0\widehat{A}(\tau)\widehat{A}^{\prime\prime}(\tau)\rightarrow 0 as ττ^2\tau\rightarrow\widehat{\tau}_{2}^{-}. Applying (16) to A^\widehat{A} and taking the limits of both sides as ττ^2\tau\rightarrow\widehat{\tau}_{2}^{-}, we obtain

2h2+12=160τ^2A^(s)𝑑s.-2h^{2}+\displaystyle\frac{1}{2}=-\displaystyle\frac{1}{6}\int_{0}^{\widehat{\tau}_{2}}\widehat{A}(s)\,ds.

Suppose there exist an ϵ>0\epsilon>0 and a sequence {μn}μ^\{\mu_{n}\}\rightarrow\hat{\mu} such that for any nn, the solution An(τ)A_{n}(\tau) to the IVP (7) with An′′(0)=μnA_{n}^{\prime\prime}(0)=\mu_{n} is below zero for all τ(0,τ^2+ϵ)\tau\in(0,\widehat{\tau}_{2}+\epsilon). Since A^(τ)<0\widehat{A}(\tau)<0 for all τ(0,τ^2)\tau\in(0,\widehat{\tau}_{2}), we can apply (12) of Corollary 3 to obtain a subsequence {μnk}\{\mu_{n_{k}}\} and some Q>0Q>0 such that for each k>Qk>Q,

|Ank(τ)A^(τ)|+|Ank(τ)A^(τ)|+|Ank′′(τ)A^′′(τ)|<1k|A_{n_{k}}(\tau)-\widehat{A}(\tau)|+|A^{\prime}_{n_{k}}(\tau)-\widehat{A}^{\prime}(\tau)|+|A^{\prime\prime}_{n_{k}}(\tau)-\widehat{A}^{\prime\prime}(\tau)|<\displaystyle\frac{1}{k}

for all τ[0,τ^21k]\tau\in[0,\widehat{\tau}_{2}-\displaystyle\frac{1}{k}]. Clearly, Ank(τ^21k)A^(τ^2)=2hA^{\prime}_{n_{k}}(\widehat{\tau}_{2}-\displaystyle\frac{1}{k})\rightarrow\widehat{A}^{\prime}(\widehat{\tau}_{2})=2h. Without loss of generality, we assume that Ank(τ^21k)>hA^{\prime}_{n_{k}}(\widehat{\tau}_{2}-\displaystyle\frac{1}{k})>h for all k>Qk>Q. Let

ωk:=sup{τ(τ^21k,τ^2+ϵ):infτ^21k<s<τAnk(s)h}τ^2.\omega_{k}:=\sup\left\{\tau\in(\widehat{\tau}_{2}-\displaystyle\frac{1}{k},\widehat{\tau}_{2}+\epsilon):\inf_{\widehat{\tau}_{2}-\displaystyle\frac{1}{k}<s<\tau}A^{\prime}_{n_{k}}(s)\geq h\right\}-\widehat{\tau}_{2}.

Note that ωk0\omega_{k}\rightarrow 0 as kk\rightarrow\infty. Otherwise, for some very large kk, Ank(τ)A_{n_{k}}(\tau) would reach zero at some τ(τ^21k,τ^2+ωk)\tau\in(\widehat{\tau}_{2}-\displaystyle\frac{1}{k},\widehat{\tau}_{2}+\omega_{k}) since Ank(τ^21k)A_{n_{k}}(\widehat{\tau}_{2}-\displaystyle\frac{1}{k}) converges to A^(τ^2)=0\widehat{A}(\widehat{\tau}_{2})=0 from below. Without loss of generality, we assume that τ^2+ωk<τ^2+ϵ\widehat{\tau}_{2}+\omega_{k}<\widehat{\tau}_{2}+\epsilon for all k>Qk>Q. Then the definition of ωk\omega_{k} implies Ank(τ^2+ωk)=hA^{\prime}_{n_{k}}(\widehat{\tau}_{2}+\omega_{k})=h and Ank′′(τ^2+ωk)0A^{\prime\prime}_{n_{k}}(\widehat{\tau}_{2}+\omega_{k})\leq 0 for all k>Qk>Q. Applying (16) to AnkA_{n_{k}} at τ=τ^2+ωk\tau=\widehat{\tau}_{2}+\omega_{k} yields

Ank(τ^2+ωk)Ank′′(τ^2+ωk)12h2+12=16(τ^2+ωk)Ank(τ^2+ωk)160τ^2+ωkAnk(s)𝑑s.A_{n_{k}}(\widehat{\tau}_{2}+\omega_{k})A^{\prime\prime}_{n_{k}}(\widehat{\tau}_{2}+\omega_{k})-\displaystyle\frac{1}{2}h^{2}+\displaystyle\frac{1}{2}=\displaystyle\frac{1}{6}(\widehat{\tau}_{2}+\omega_{k})A_{n_{k}}(\widehat{\tau}_{2}+\omega_{k})-\displaystyle\frac{1}{6}\int_{0}^{\widehat{\tau}_{2}+\omega_{k}}\!A_{n_{k}}(s)\,ds.

Since 0>Ank(τ^2+ωk)>Ank(τ^21k)A^(τ^2)=00>A_{n_{k}}(\widehat{\tau}_{2}+\omega_{k})>A_{n_{k}}(\widehat{\tau}_{2}-\displaystyle\frac{1}{k})\rightarrow\widehat{A}(\widehat{\tau}_{2})=0, the right-hand side converges to 160τ^2A^(s)𝑑s=2h2+12-\displaystyle\frac{1}{6}\int_{0}^{\widehat{\tau}_{2}}\widehat{A}(s)\,ds=-2h^{2}+\displaystyle\frac{1}{2}. However, for all k>Qk>Q the left-hand side is greater than or equal to 12h2+12-\displaystyle\frac{1}{2}h^{2}+\displaystyle\frac{1}{2}. This is a contradiction. ∎

In the next lemma, we still assume A^′′(0)=μ^0\widehat{A}^{\prime\prime}(0)=\hat{\mu}\geq 0.

Lemma 10.

Suppose A^(τ^2)<0\widehat{A}(\widehat{\tau}_{2})<0 and A^(τ^2)=0\widehat{A}^{\prime}(\widehat{\tau}_{2})=0. Then there exists a δ>0\delta>0 such that for any μ(μ^δ,μ^+δ)\mu\in(\hat{\mu}-\delta,\hat{\mu}+\delta), the solution A(τ)A(\tau) to the IVP (7) with A′′(0)=μA^{\prime\prime}(0)=\mu satisfies A(τ2)<0A(\tau_{2})<0 and A(τ2)=0A^{\prime}(\tau_{2})=0.

Proof.

Define τ^1\widehat{\tau}_{1} for A^(τ)\widehat{A}(\tau) as in (13). Then A^′′(τ^1)>0\widehat{A}^{\prime\prime}(\widehat{\tau}_{1})>0 since A^′′(0)=μ^0\widehat{A}^{\prime\prime}(0)=\hat{\mu}\geq 0. It follows that A^(τ^1)=A^(τ^2)=0\widehat{A}^{\prime}(\widehat{\tau}_{1})=\widehat{A}^{\prime}(\widehat{\tau}_{2})=0 and A^(τ)>0\widehat{A}^{\prime}(\tau)>0 for all τ(τ^1,τ^2)\tau\in(\widehat{\tau}_{1},\widehat{\tau}_{2}). Thus there is a τ^c(τ^1,τ^2)\widehat{\tau}_{c}\in(\widehat{\tau}_{1},\widehat{\tau}_{2}) such that A^′′(τ^c)=0\widehat{A}^{\prime\prime}(\widehat{\tau}_{c})=0. Furthermore, since A^′′′=τA^6A^<0\widehat{A}^{\prime\prime\prime}=\displaystyle\frac{\tau\widehat{A}^{\prime}}{6\widehat{A}}<0 for all τ[τ^c,τ^2)\tau\in[\widehat{\tau}_{c},\widehat{\tau}_{2}), we have A^′′(τ^2)<0\widehat{A}^{\prime\prime}(\widehat{\tau}_{2})<0. This allows us to choose a small α>0\alpha>0 such that A^(τ)<0\widehat{A}(\tau)<0 for all τ(0,τ^2+α]\tau\in(0,\widehat{\tau}_{2}+\alpha], A^(τ^2+α)<0\widehat{A}^{\prime}(\widehat{\tau}_{2}+\alpha)<0, and A^′′(τ^2+α)<0\widehat{A}^{\prime\prime}(\widehat{\tau}_{2}+\alpha)<0. Take a small β>0\beta>0. Then Corollary 3 guarantees the existence of a δ>0\delta>0 such that for any μ(μ^δ,μ^+δ)\mu\in(\hat{\mu}-\delta,\hat{\mu}+\delta), the solution A(τ)A(\tau) with A′′(0)=μA^{\prime\prime}(0)=\mu is below zero for all τ(0,τ^2+α]\tau\in(0,\widehat{\tau}_{2}+\alpha] and satisfies A(τ^2β)>0A^{\prime}(\widehat{\tau}_{2}-\beta)>0 and A(τ^2+α)<0A^{\prime}(\widehat{\tau}_{2}+\alpha)<0. This implies that A(τ2)<0A(\tau_{2})<0 and A(τ2)=0A^{\prime}(\tau_{2})=0 at τ2(τ^2β,τ^2+α)\tau_{2}\in(\widehat{\tau}_{2}-\beta,\widehat{\tau}_{2}+\alpha). ∎

3.5. Final steps

We are now in a position to complete the proof of Proposition 1.

Proof.

Define the set Ω\Omega as follows:

Ω:={μ¯𝐑:A(τ2)=0 and A(τ2)>0 for any μμ¯}.\Omega:=\left\{\bar{\mu}\in{\bf{R}}:A(\tau_{2})=0\text{ and }A^{\prime}(\tau_{2})>0\text{ for any }\mu\geq\bar{\mu}\right\}.

Lemma 7 guarantees that Ω\Omega is nonempty, and Lemma 8 shows that zero is a lower bound of Ω\Omega. Take μ:=infΩ>0\mu_{*}:=\inf\Omega>0, with the inequality guaranteed by Lemma 8 and Lemma 10. Let A(τ)A_{*}(\tau) be the solution to the IVP (7) with A′′(0)=μA_{*}^{\prime\prime}(0)=\mu_{*}. Then by Lemma 9 and the definition of μ\mu_{*}, it is only possible that either A(τ2)=0A_{*}(\tau_{2})=0 and A(τ2)=0A_{*}^{\prime}(\tau_{2})=0 or A(τ2)<0A_{*}(\tau_{2})<0 and A(τ2)=0A_{*}^{\prime}(\tau_{2})=0. Furthermore, we rule out the second case by the definition of μ\mu_{*} and Lemma 10. This proves (i)–(iv) of Proposition 1.

Since A(τ)<0A_{*}(\tau)<0 and A(τ)>0A_{*}^{\prime}(\tau)>0 for all τ(τ1,τ2)\tau\in(\tau_{1},\tau_{2}), we have

τ2A6A<A′′′<τ1A6A\displaystyle\frac{\tau_{2}A_{*}^{\prime}}{6A_{*}}<A_{*}^{\prime\prime\prime}<\displaystyle\frac{\tau_{1}A_{*}^{\prime}}{6A_{*}}

for all τ(τ1,τ2)\tau\in(\tau_{1},\tau_{2}). Integrating the above inequality gives

τ26ln|A(τ)A(τ1)|+A′′(τ1)<A′′(τ)<τ16ln|A(τ)A(τ1)|+A′′(τ1)\displaystyle\frac{\tau_{2}}{6}\ln\left|\displaystyle\frac{A_{*}(\tau)}{A_{*}(\tau_{1})}\right|+A_{*}^{\prime\prime}(\tau_{1})<A_{*}^{\prime\prime}(\tau)<\displaystyle\frac{\tau_{1}}{6}\ln\left|\displaystyle\frac{A_{*}(\tau)}{A_{*}(\tau_{1})}\right|+A_{*}^{\prime\prime}(\tau_{1})

for all τ(τ1,τ2)\tau\in(\tau_{1},\tau_{2}). Thus A′′(τ)A_{*}^{\prime\prime}(\tau)\rightarrow-\infty as ττ2\tau\rightarrow\tau_{2}^{-}. Then we can take a sufficiently small α>0\alpha>0 such that A′′(τ)<1A_{*}^{\prime\prime}(\tau)<-1 for all τ(τ2α,τ2)\tau\in(\tau_{2}-\alpha,\tau_{2}). By the mean value theorem we have

A(τ2)A(τ)=A′′(s)(τ2τ)<(τ2τ)A_{*}^{\prime}(\tau_{2})-A_{*}^{\prime}(\tau)=A^{\prime\prime}(s)(\tau_{2}-\tau)<-(\tau_{2}-\tau)

for any τ(τ2α,τ2)\tau\in(\tau_{2}-\alpha,\tau_{2}) and some s(τ,τ2)s\in(\tau,\tau_{2}). Since A(τ2)=0A_{*}^{\prime}(\tau_{2})=0, this implies that 0<(τ2τ)<A(τ)0<(\tau_{2}-\tau)<A_{*}^{\prime}(\tau) and consequently 12(τ2τ)2<A(τ2)A(τ)=A(τ)\displaystyle\frac{1}{2}(\tau_{2}-\tau)^{2}<A_{*}(\tau_{2})-A_{*}(\tau)=-A_{*}(\tau) for all τ(τ2α,τ2)\tau\in(\tau_{2}-\alpha,\tau_{2}). Thus ln|A(τ)|=O(ln(|ττ2|))\ln|A_{*}(\tau)|=O(\ln(|\tau-\tau_{2}|)) as ττ2\tau\to\tau_{2}^{-}. This proves (v) of Proposition 1.

4. Numerical Study of K(2,2)K(2,2)

4.1. Regularization and scaling

In this section we numerically assess the ill-posedness of (2). But it is inappropriate to directly simulate an equation that not only lacks a local well-posedness theory but for which we suspect ill-posedness. Thus we regularize the equation. Simulating this regularized problem, we find evidence of the ill-posedness as we let the regularization parameter vanish.

We study the following regularization of (2)

(17) (I+δx4)tu=x(u2)+x3(u2).(I+\delta\partial_{x}^{4})\partial_{t}u=\partial_{x}\left(u^{2}\right)+\partial_{x}^{3}\left(u^{2}\right).

Here II is the identity and δ>0\delta>0 is the small regularization parameter. We choose this particular regularization since its implementation is natural when simulating solutions of (2) via a pseudospectral method.

Inverting the operator on the left-hand side to be able to write it as an evolution equation, we see that x/(I+δx4)\partial_{x}/(I+\delta\partial_{x}^{4}) and x3/(I+δx4)\partial_{x}^{3}/(I+\delta\partial_{x}^{4}) are bounded operators. Indeed,

xI+δx4HsHs+3Cδ1andx3I+δx4HsHs+1Cδ1.\left\|{\frac{\partial_{x}}{I+\delta\partial_{x}^{4}}}\right\|_{H^{s}\to H^{s+3}}\leq C\delta^{-1}\quad{\textrm{and}}\quad\left\|{\frac{\partial_{x}^{3}}{I+\delta\partial_{x}^{4}}}\right\|_{H^{s}\to H^{s+1}}\leq C\delta^{-1}.

The boundedness of these operators make it trivial to prove:

Theorem 11 (Local Well-Posedness of a Regularized Problem).

Fix δ>0\delta>0. Then (17) is locally well-posed in HsH^{s} for any s>12s>\tfrac{1}{2}

  • For all u0Hsu_{0}\in H^{s}, there exists a T>0T>0 and a unique function u(t)C(0,T;Hs)u(t)\in C(0,T;H^{s}) solving the integral form of (17),

  • The solution will depend continuously upon the data,

  • There exists a maximal time of existence, TexistT_{\mathrm{exist}} such that if it is finite,

    limtTexistu(t)Hs=\lim_{t\to T_{\mathrm{exist}}}\lVert u(t)\rVert_{H^{s}}=\infty

The proof of this, which we omit, is an elementary application of the fixed point technique. The operator is bounded in L2L^{2}-based Sobolev spaces. For s>12s>\tfrac{1}{2}, HsH^{s} is an algebra which makes the nonlinearity easy to treat. An impediment to extending Theorem 11 to one which is global in time, or even one which holds on time intervals of a length uniform in δ\delta, is the lack of an obvious coercive conserved quantity associated with (17). The only obvious conserved quantity for (17) is

u(x,t)𝑑x=u(x,0)𝑑x.\int u(x,t)dx=\int u(x,0)dx.

This is formally an invariant for solutions of (2)\eqref{K22} as well.

By simulating the regularized equation (17) we seek evidence of ill-posedness of (2). In particular, we shall find a sequence of vanishing initial conditions for (17), such that as δ0\delta\to 0, the corresponding solutions at t=T>0t=T_{\star}>0 have H2H^{2} norm of (at least) unit size. To that end, we will scale the initial data for (17) using the same scaling given by (5) and link the vanishing of the smoothing parameter δ\delta to the scaling parameter λ\lambda. Specifically we will study (17) with

(18) δ=δλ:=.1λ4andu(x,0)=fλ(x):=λf(xλ1/3)\delta=\delta_{\lambda}:=.1\lambda^{4}{\quad\textrm{and}\quad}u(x,0)=f_{\lambda}(x):=\lambda f\left({x\over\lambda^{1/3}}\right)

where

f(x)=exp(4x2)f(x)=\exp(-4x^{2})

and λ0\lambda\to 0. Note that the calculations leading to (6) show that fλH2Cλ7/6\|f_{\lambda}\|_{H^{2}}\leq C\lambda^{7/6} and so this choice of initial data vanishes in the limit. We denote the solution of (17) with the choices in (18) by uλ(x,t)u_{\lambda}(x,t). Note that we have taken our initial data to be everywhere positive, unlike the self-similar solutions found for (3). We do this to demonstrate that the problematic effects of degeneracy manifest themselves even for solutions which do not cross the xx-axis.

The choices for δλ\delta_{\lambda} and fλ(x)f_{\lambda}(x) are made for the following reason. Instead of scaling just the initial data for (17) suppose that we rescale the whole equation by

(19) uλ(x,t)=λvλ(xλ1/3,t).u_{\lambda}(x,t)=\lambda v_{\lambda}\left({x\over\lambda^{1/3}},t\right).

Then (17) becomes:

(I+δλλ4/3y4)tvλ=λ2/3yvλ2+y3vλ2(I+\delta_{\lambda}\lambda^{-4/3}\partial_{y}^{4})\partial_{t}v_{\lambda}=\lambda^{2/3}\partial_{y}v_{\lambda}^{2}+\partial_{y}^{3}v_{\lambda}^{2}

where y=x/λ1/3y=x/\lambda^{1/3}. Plugging in from (18) we get

vλ(y,0)=f(y)v_{\lambda}(y,0)=f(y)

and

(20) (I+.1λ8/3y4)tvλ=λ2/3yvλ2+y3vλ2.(I+.1\lambda^{8/3}\partial_{y}^{4})\partial_{t}v_{\lambda}=\lambda^{2/3}\partial_{y}v_{\lambda}^{2}+\partial_{y}^{3}v_{\lambda}^{2}.

Of course (20) is not an exact scaling. Notice that as λ0\lambda\to 0 the term yvλ2\partial_{y}v_{\lambda}^{2} term will vanish. The equation will asymptotically be dominated by the third derivative term which is consistent with our expectation that this term is the source of the ill-posedness. Moreover notice that the coefficient of regularization even in these scaled coordinates vanishes. That is to say, we have chosen the regularization parameter so that it vanishes “more rapidly” than scaling effects.

4.2. Computational results

We integrate our problem pseudospectrally with a Crank-Nicholson time stepping scheme. Some additional details are presented in Section 4.3. Simulating (17) with (18) to T=.1T=.1 on the domain [2π,2π)[-2\pi,2\pi) with 81928192 resolved Fourier modes, the data evolves as in Figures 5 and 6. At large values of λ\lambda, there is some development of oscillations. At very small values of λ\lambda, it develops a highly oscillatory structure. Note that these oscillations appear for x>0x>0, where f<0f^{\prime}<0. That is to say in the exactly the place where the term 6uxuxx6u_{x}u_{xx} in (2) acts, heuristically, as a backwards heat operator.

Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 5. Evolution of the regularized K(2,2)K(2,2) equation with data given by (18) with λ=.4\lambda=.4.
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 6. Evolution of the regularized K(2,2)K(2,2) equation with data given by (18) with λ=.05\lambda=.05.

We present the evolution of the L2L^{2}, H˙1\dot{H}^{1}, and H˙2\dot{H}^{2} norms in Figure 7 of the solutions as functions of time. Though there is little growth in time of the L2L^{2} norm, we see orders of magnitude jumps in the H˙1\dot{H}^{1} and H˙2\dot{H}^{2} norms at t=.1t=.1 as we send λ0\lambda\to 0.

Refer to caption
Refer to caption
Refer to caption
Figure 7. Evolution of several norms for the regularized K(2,2)K(2,2) equation with data given by (18).

We contend that this is evidence of ill-posedness. We have a sequence of initial conditions fλf_{\lambda}, which have vanishing H2H^{2} norm. At a fixed T>0T>0, the H2H^{2} norms are growing as we let λ0\lambda\to 0. Thus, there is a loss of continuous dependence of the solution map upon the initial data about the u(x,t)=0u(x,t)=0 solution.

This growth in the H2H^{2} norm corresponds to a spreading in Fourier space. Indeed, Figure 8 shows precisely this. At fixed time, the Fourier support grows as δ0\delta\to 0. These figures also indicate that except for λ=.04\lambda=.04, all simulations are extremely well resolved. Even the λ=.04\lambda=.04 simulation is well resolved through t=.06t=.06, and only slightly under resolved at the final time, t=.1t=.1.

Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 8. Evolution of the regularized K(2,2)K(2,2) equation with data given by (18) with λ=.05\lambda=.05.

4.3. Details of numerical methods

As noted, the simulations presented in Section 4 were generated using a fully de-aliased pseudospectral discretization with Crank-Nicholson time stepping. This was performed in Matlab and fsolve was used to solve the nonlinear system at each time step with a tolerance of 10810^{-8}. 8192 modes were resolved in these simulations, with a time step of Δt=.001\Delta t=.001.

The invariant u(x,t)𝑑x\displaystyle\int u(x,t)dx was conserved to at least eight significant figures in all the computed cases, as shown in Table 1.

Table 1. Numerical approximation of u(x,t)𝑑x\int u(x,t)dx at the outputted times for different values of λ\lambda
λ\lambda
1 0.4 0.2 0.1 0.05
tt 0.00 0.886226925453 0.261191032665 0.103653730004 0.0411350100123 0.0163244395416
0.02 0.886226925477 0.261191032667 0.103653730004 0.0411350100123 0.0163244395416
0.04 0.886226925489 0.261191032674 0.103653730004 0.0411350100123 0.0163244395416
0.06 0.886226925478 0.261191032676 0.103653730004 0.0411350100123 0.0163244395416
0.08 0.886226925497 0.261191032673 0.103653730004 0.0411350100123 0.0163244395416
0.10 0.886226925509 0.261191032673 0.103653730004 0.0411350100123 0.0163244395416

4.4. Convergence of Norms

Since we are concerned with the value of H˙2\dot{H}^{2} at the end of our simulation, it is also important to confirm that this is converging to a fixed value as we refine the spatial and temporal resolution of our simulations. For the initial condition with λ=0.2\lambda=0.2, Table 2 gives the values of the H˙2\dot{H}^{2} norm for various choices of the number of resolved Fourier modes, NN, and the time step, Δt\Delta t. With only 512 resolved modes, we appear to be fully resolved in space, and there is little gain in accuracy when NN is increased. Of course, for other initial conditions, more than 512 modes may be needed to resolve the simulation.

Table 2. Values of H˙2\dot{H}^{2} at t=.1t=.1 corresponding to the λ=0.2\lambda=0.2 simulation with different numbers of resolved Fourier modes, NN, and time steps, Δt\Delta t.
Δt\Delta t
0.004 0.002 0.001
NN 512 675.27623578 692.22147559 696.5301160
1024 675.27623585 692.22147576 696.5301162
2048 675.27623587 692.22147576 696.5301162
4096 675.27623586 692.22147576 696.5301162
8192 675.27623586 692.22147576 696.5301162

For the three values of Δt\Delta t, we appear to have achieved at least one significant digit of accuracy and as we reduce the time step, the variations of H˙2\dot{H}^{2} become smaller. Since we are using a Crank-Nicholson scheme, we expect O(Δt2){O}(\Delta t^{2}) convergence. If we perform Richardson extrapolation, the estimated value of H˙2\dot{H}^{2} based on the Δt=.004\Delta t=.004 and Δt=.002\Delta t=.002 simulations is 697.87. Comparing the Δt=.002\Delta t=.002 and Δt=.001\Delta t=.001 simulations, the estimated value is 697.97. Separate computations, performed with just 512 grid points for expediency, show that with Δt=0.0005\Delta t=0.0005 the norm takes the value 697.96, and for Δt=0.00025\Delta t=0.00025, the value is 697.99.

An examination of the λ=0.1\lambda=0.1 case is similar. For the same three values of Δt\Delta t, the H˙2\dot{H}^{2} norms, given in Table 3, are in agreement on the order of magnitude. Richardson extrapolation using the data at Δt=.004\Delta t=.004 and Δt=.002\Delta t=.002 predicts a value of 15148, while the predicted value based on the Δt=.002\Delta t=.002 and Δt=.001\Delta t=.001 value is 15278. Independent simulations with 2048 grid points yield a value of 15253 when Δt=0.0005\Delta t=0.0005 and a value of 15302 when Δt=0.00025\Delta t=0.00025. Thus, we believe that our choice of discretization parameters, N=8192N=8192 and Δt=0.001\Delta t=0.001, for the results presented in Section 4.2 yield meaningful measurements of H˙2\dot{H}^{2} that are accurate to at least one signfigant figure. This is sufficient for our study of ill-posedness.

Table 3. Values of H˙2\dot{H}^{2} at t=.1t=.1 corresponding to the λ=0.1\lambda=0.1 simulation with different numbers of resolved Fourier modes, NN, and time steps, Δt\Delta t.
Δt\Delta t
0.004 0.002 0.001
NN 2048 12452.791732 14474.328299 15077.20691
4096 12452.791732 14474.328298 15077.20691
8192 12452.791732 14474.328299 15077.20691

hi!

References

  • [1] D. M. Ambrose and N. Masmoudi. Well-posedness of 3D vortex sheets with surface tension. Commun. Math. Sci., 5(2):391–430, 2007.
  • [2] D. M. Ambrose and J. D. Wright. Traveling waves and weak solutions for an equation with degenerate dispersion. Preprint.
  • [3] D. M. Ambrose and J. D. Wright. Preservation of support and positivity for solutions of degenerate evolution equations. Nonlinearity, 23(3):607–620, 2010.
  • [4] G. I. Barenblatt. On self-similar motions of a compressible fluid in a porous medium. Akad. Nauk SSSR. Prikl. Mat. Meh., 16:679–698, 1952.
  • [5] R. Camassa and D. D. Holm. An integrable shallow water equation with peaked solitons. Phys. Rev. Lett., 71(11):1661–1664, 1993.
  • [6] A. Chertock and D. Levy. Particle methods for dispersive equations. J. Comput. Phys., 171(2):708–730, 2001.
  • [7] W. Craig and J. Goodman. Linear dispersive equations of Airy type. J. Differential Equations, 87(1):38–61, 1990.
  • [8] W. Craig, T. Kappeler, and W. Strauss. Gain of regularity for equations of KdV type. Ann. Inst. H. Poincaré Anal. Non Linéaire, 9(2):147–186, 1992.
  • [9] H.-H. Dai and Y. Huo. Solitary shock waves and other travelling waves in a general compressible hyperelastic rod. R. Soc. Lond. Proc. Ser. A Math. Phys. Eng. Sci., 456(1994):331–363, 2000.
  • [10] J. de Frutos, M. Á. López Marcos, and J. M. Sanz-Serna. A finite difference scheme for the K(2,2)K(2,2) compacton equation. J. Comput. Phys., 120(2):248–252, 1995.
  • [11] F. Fraternali, M. A. Porter, and C. Daraio. Optimal Design of Composite Granular Protectors. Mech. Adv. Mater. Struct., 17(1):1–19, 2010.
  • [12] V. A. Galaktionov. Nonlinear dispersion equations: smooth deformations, compactions, and extensions to higher orders. Zh. Vychisl. Mat. Mat. Fiz., 48(10):1859, 2008.
  • [13] V. A. Galaktionov and S. I. Pokhozhaev. Third-order nonlinear dispersion equations: shock waves, rarefaction waves, and blow-up waves. Zh. Vychisl. Mat. Mat. Fiz., 48(10):1819–1846, 2008.
  • [14] J. Goodman and P. D. Lax. On dispersive difference schemes. I. Comm. Pure Appl. Math., 41(5):591–613, 1988.
  • [15] C. E. Kenig, G. Ponce, and L. Vega. The initial value problem for the general quasi-linear Schrödinger equation. In Recent developments in nonlinear partial differential equations, volume 439 of Contemp. Math., pages 101–115. Amer. Math. Soc., Providence, RI, 2007.
  • [16] D. Ketcheson. High Order Strong Stability Preserving Time Integrators and Numerical Wave Propagation Methods for Hyperbolic PDEs. 2009.
  • [17] D. Levy, C.-W. Shu, and J. Yan. Local discontinuous Galerkin methods for nonlinear dispersive equations. J. Comput. Phys., 196(2):751–772, 2004.
  • [18] Y. A. Li, P. J. Olver, and P. Rosenau. Non-analytic solutions of nonlinear wave models. In Nonlinear theory of generalized functions (Vienna, 1997), volume 401 of Chapman & Hall/CRC Res. Notes Math., pages 129–145. Chapman & Hall/CRC, Boca Raton, FL, 1999.
  • [19] V. F. Nesterenko. Dynamics of hetereogeneous materials. Springer, New York, NY, 2001.
  • [20] L. Ponson, N. Boechler, Y. M. Lai, M. A. Porter, P. G. Kevrekidis, and C. Daraio. Nonlinear waves in disordered diatomic granular chains. Phys. Rev. E, 82(2, Part 1), AUG 12 2010.
  • [21] M. Porter, C. Daraio, I. Szelengowicz, E. Herbold, and P. Kevrekidis. Highly nonlinear solitary waves in heterogeneous periodic granular media. Physica D, 2009.
  • [22] P. Rosenau. Nonlinear dispersion and compact structures. Phys. Rev. Lett., 73(13):1737–1741, 1994.
  • [23] P. Rosenau. Compact and noncompact dispersive patterns. Phys. Lett. A, 275(3):193–203, 2000.
  • [24] P. Rosenau and J. Hyman. Compactons: solitons with finite wavelength. Phys. Rev. Lett., (70):564–567, 1993.
  • [25] P. Rosenau and E. Kashdan. Emergence of compact structures in a klein-gordon model. Phys. Rev. Lett., (104), 2010.
  • [26] J. Rubinstein and J. B. Keller. Sedimentation of a dilute suspension. Phys. Fluids A, 1(4):637–643, 1989.
  • [27] F. Rus and F. R. Villatoro. Padé numerical method for the Rosenau-Hyman compacton equation. Math. Comput. Simulation, 76(1-3):188–192, 2007.
  • [28] F. Rus and F. R. Villatoro. Self-similar radiation from numerical Rosenau-Hyman compactons. J. Comput. Phys., 227(1):440–454, 2007.
  • [29] G. Simpson, M. Spiegelman, and M. I. Weinstein. Degenerate dispersive equations arising in the study of magma dynamics. Nonlinearity, 20(1):21–49, 2007.
  • [30] C. E. Wayne and J. D. Wright. Higher order modulation equations for a Boussinesq equation. SIAM J. Appl. Dyn. Syst., 1(2):271–302 (electronic), 2002.
  • [31] J. D. Wright. Higher order corrections to the KdV approximation for water waves. ProQuest LLC, Ann Arbor, MI, 2004. Thesis (Ph.D.)–Boston University.