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

Illposedness via degenerate dispersion for generalized surface quasi-geostrophic equations with singular velocities

Dongho Chae Department of Mathematics, Chung-ang University. E-mail: [email protected]    In-Jee Jeong Department of Mathematical Sciences and RIM, Seoul National University. E-mail: [email protected]    Sung-Jin Oh Department of Mathematics, UC Berkeley and School of Mathematics, Korea Institute for Advanced Study. E-mail: [email protected]
Abstract

We prove strong nonlinear illposedness results for the generalized SQG equation

tθ+Γ[θ]θ=0\begin{split}\partial_{t}\theta+\nabla^{\perp}\Gamma[\theta]\cdot\nabla\theta=0\end{split}

in any sufficiently regular Sobolev spaces, when Γ\Gamma is a singular multiplier in the sense that its symbol satisfies |Γ(ξ)||\Gamma(\xi)|\rightarrow\infty as |ξ||\xi|\rightarrow\infty together some mild regularity assumptions. The key mechanism is degenerate dispersion, i.e., the rapid growth of frequencies of solutions around certain shear states, as in the second and third author’s earlier work on Hall-magnetohydrodynamics [20]. The robustness of our method allows one to extend linear and nonlinear illposedness to fractionally dissipative systems, as long as the order of dissipation is lower than that of Γ\Gamma. Our illposedness results are completely sharp in view of various existing wellposedness statements as well as those from our companion paper [10].

Key to our proofs is a novel construction of degenerating wave packets for the class of linear equations

tϕ+ip(t,X,D)ϕ=0\begin{split}\partial_{t}\phi+ip(t,X,D)\phi=0\end{split}

possibly with lower order terms, where p(t,X,D)p(t,X,D) is a possibly time dependent pseudo-differential operator which is formally self-adjoint in L2L^{2}, degenerate, and dispersive. Degenerating wave packets are approximate solutions to the above linear equation with spatial and frequency support localized at (X(t),Ξ(t))(X(t),\Xi(t)), which are solutions to the bicharacteristic ODE system associated with p(t,x,ξ)p(t,x,\xi). These wave packets explicitly show degeneration as X(t)X(t) approaches a point where pp vanishes, which in particular allows us to prove illposedness in topologies finer than L2L^{2}. While the equation for the wave packet can be formally obtained from a Taylor expansion of the symbol near ξ=Ξ(t)\xi=\Xi(t), the difficult part is to rigorously control the error in sufficiently long timescales within which significant degeneration occurs. To achieve this task, we develop a systematic way to obtain sharp estimates for not only degenerating wave packets but also for oscillatory integrals which naturally appear in the error estimate.

1 Introduction

1.1 Generalized SQG equations

In two spatial dimensions, the generalized surface quasi-geostrophic (gSQG) equations are given by

{tθ+uθ=0,u=Γθ,\left\{\begin{aligned} &\partial_{t}\theta+u\cdot\nabla\theta=0,\\ &u=\nabla^{\perp}\Gamma\theta,\end{aligned}\right. (1.1)

where =(x2,x1)\nabla^{\perp}=(-\partial_{x_{2}},\partial_{x_{1}})^{\top} and Γ\Gamma is a Fourier multiplier with a real-valued symbol γ\gamma. Here, θ(t,):Ω\theta(t,\cdot):\Omega\rightarrow\mathbb{R} and u(t,):Ω2u(t,\cdot):\Omega\rightarrow\mathbb{R}^{2} with Ω\Omega a two-dimensional domain without boundaries (e.g. Ω=𝕋2\Omega=\mathbb{T}^{2}, 2\mathbb{R}^{2}, or 𝕋×\mathbb{T}\times\mathbb{R}). The system (1.1) says that the scalar θ\theta is being advected by the flow of uu, which is determined from θ\theta at each moment of time by the “Biot–Savart” law u=Γθu=\nabla^{\perp}\Gamma\theta. For this reason, (1.1) is sometimes referred to as an active scalar system.

The system (gSQG) generalizes a number of important PDEs arising from hydro- and magnetohydrodynamics (MHD), and has been intensively studied in the past few decades. As it is well-known, when Γ=(Δ)1\Gamma=(-\Delta)^{-1}, (1.1) is simply the vorticity equation for the two-dimensional incompressible Euler equations; note that θ=×u\theta=\nabla\times u. On the other hand, the case Γ=Λ1\Gamma=\Lambda^{-1} corresponds to the (usual) SQG equations describing the motion of atmospheric fronts, where Λ=(Δ)12\Lambda=(-\Delta)^{\frac{1}{2}} is the Zygmund operator. This model was derived by Constantin–Majda–Tabak in [13, 12] to describe the Boussinesq dynamics in the boundary of the upper half-space (see [29] for details). In the other extreme, the cases Γ=Λ\Gamma=\Lambda and Λ2\Lambda^{2} appear quite naturally in magnetohydrodynamics and large-scale atmospheric dynamics, respectively. We shall return to these examples below.

The pioneering works [13, 12] already suggested the possibility of rapid small scale creation for the SQG equation, and while significant efforts have been devoted to the question of global regularity for the SQG equation Γ=Λ1\Gamma=\Lambda^{-1}, to the best of our knowledge, there is still no convincing evidence that smooth solutions of the SQG equation blows up in finite time. This is in stark contrast with the Euler case Γ=Λ2\Gamma=\Lambda^{-2} for which global regularity is a classical result. It is expected that as the multiplier Γ\Gamma becomes more singular, smooth solutions are more likely to develop singularities in finite time; the conservation law θ(t,)Lp\|{\theta(t,\cdot)}\|_{L^{p}} becomes relatively weaker compared with the velocity field. (When Γ\Gamma is more regular than the 2D Euler case, the conservation of θ(t,)L\|{\theta(t,\cdot)}\|_{L^{\infty}} guarantees global regularity, as in 2D Euler.) For this reason, the global regularity question has been extensively studied for the generalized models where Γ\Gamma is taken to be Λβ\Lambda^{\beta} for some β>2\beta>-2.

However, as one increases β\beta, even the question of local regularity for smooth solutions becomes non-trivial, already when the multiplier becomes more singular than the SQG case, i.e. β>1\beta>-1. Then, uu is more singular than θ\theta and there is a serious difficulty in closing energy estimates in Sobolev spaces. Namely, one needs to worry about the term where all derivatives in the HmH^{m} estimate falls on uu:

m(Λβθ)θ,mθ=θΛβg,g,\begin{split}\langle{\partial^{m}(\nabla^{\perp}\Lambda^{\beta}\theta)\cdot\nabla\theta,\partial^{m}\theta}\rangle=\langle{\nabla\theta\cdot\nabla^{\perp}\Lambda^{\beta}g,g}\rangle,\end{split}

with g:=mθg:=\partial^{m}\theta and ,\langle{\cdot,\cdot}\rangle representing the usual L2L^{2} inner product. Note that the principal part of the operator θΛβ\nabla\theta\cdot\nabla^{\perp}\Lambda^{\beta} is anti-symmetric, which allows one to (formally) gain a derivative. This observation was used in [8] to obtain local well-posedness for β<0\beta<0.111One may generalize this observation to obtain energy estimates when |γ(ξ)|1|\gamma(\xi)|\lesssim 1 in the limit |ξ|0|\xi|\to 0, assuming some natural regularity assumptions for the derivatives of γ\gamma. Thereafter, behavior of solutions in the regime β(1,0)\beta\in(-1,0) has been investigated by many authors.

1.2 The case of singular multipliers

Turning to the case of singular multipliers, by which we simply mean that γ(ξ)\gamma(\xi)\rightarrow\infty as |ξ||\xi|\rightarrow\infty, one first sees that in the “borderline” case when β=0\beta=0 (Γ=Λβ\Gamma=\Lambda^{\beta}), the nonlinearity vanishes completely: uθ=θθ0u\cdot\nabla\theta=\nabla^{\perp}\theta\cdot\nabla\theta\equiv 0. This could make one speculate that there might be some additional cancellation which gives local regularity even when β>0\beta>0. Furthermore, the (formally) conserved quantity Γ12θ(t,)L2\|{\Gamma^{\frac{1}{2}}\theta(t,\cdot)}\|_{L^{2}} for (1.1) becomes stronger than the other L2L^{2}-based conservation law θ(t,)L2\|{\theta(t,\cdot)}\|_{L^{2}} as soon as β>0\beta>0.

Despite these facts, our main result shows strong illposedness in Sobolev spaces (with arbitrarily high regularity) for singular Γ\Gamma satisfying a few reasonable assumptions. Interestingly, the generalized SQG equations with singular multipliers naturally appear in a variety of situations, as we shall now explain.

  • Ohkitani model. In the papers [26, 27], Ohkitani considered the collective behavior of solutions to (1.1) which are obtained by varying β<0\beta<0 with the same initial data, towards the goal of settling the question of global regularity versus finite time singularity formation. Numerical simulations in [26, 27] did not show any singular behavior of solutions in the limit β0\beta\to 0^{-}, and based on these, Ohkitani conjectured global regularity of the limiting model

    tθln(Λ)θθ=0\begin{split}\partial_{t}\theta-\nabla^{\perp}\ln(\Lambda)\theta\cdot\nabla\theta=0\end{split} (1.2)

    which is (1.1) with Γ=ln(Λ)\Gamma=-\ln(\Lambda). This has been referred to as Ohkitani model in [8]. To see how (1.2) arises, one can simply rewrite (1.1) with Γ=Λβ\Gamma=\Lambda^{\beta} with β<0\beta<0 as

    1βtθ+(Λβ1β)θθ=0,\begin{split}\frac{1}{\beta}\partial_{t}\theta+\nabla^{\perp}\left(\frac{\Lambda^{\beta}-1}{\beta}\right)\theta\cdot\nabla\theta=0,\end{split} (1.3)

    and formally we have that as β0\beta\to 0^{-}, (1.3) converges to (1.2) in the rescaled timescale βt\beta t. This limit was made rigorous in our companion paper [10].

  • E-MHD system. The electron magnetohydrodynamics (E-MHD) system takes the form

    {tB+×((×B)×B)=0,B=0,\left\{\begin{aligned} &\partial_{t}B+\nabla\times((\nabla\times B)\times B)=0,\\ &\nabla\cdot B=0,\end{aligned}\right. (1.4)

    where B(t,x):×33B(t,x):\mathbb{R}\times\mathbb{R}^{3}\rightarrow\mathbb{R}^{3}. This is the magnetic part of the well-known Hall–MHD system ([28]). Under the so-called 2+122+\frac{1}{2}-dimensional assumption, BB can be written as

    B=ψ×ez+Λϕez\begin{split}B=\nabla\psi\times e_{z}+\Lambda\phi\,e_{z}\end{split}

    for some scalar functions ψ,ϕ\psi,\phi independent of the third coordinate zz, and (1.4) reduces to the following system in two dimensions ([23]):

    {tψ+Λϕψ=0,tϕ+Λ1(Λ(Λψ)ψ)=0.\left\{\begin{aligned} &\partial_{t}\psi+\nabla^{\perp}\Lambda\phi\cdot\nabla\psi=0,\\ &\partial_{t}\phi+\Lambda^{-1}(\nabla^{\perp}\Lambda(\Lambda\psi)\cdot\nabla\psi)=0.\end{aligned}\right. (1.5)

    Up to leading order, the ansatz ψϕ\psi\simeq\phi propagates in time, which simply corresponds to (1.1) with Γ=Λ\Gamma=\Lambda.

  • Asymptotic model for the large-scale quasi-geostrophic equation. The following equation is referred to as the asymptotic model (AM) for the large-scale quasi-geostrophic equation (see [22, 4, 36, 5]):

    tψ+Δψψ=0.\begin{split}\partial_{t}\psi+\nabla^{\perp}\Delta\psi\cdot\nabla\psi=0.\end{split} (1.6)

    Notice that this is nothing but (1.1) with the choice Γ=Δ\Gamma=\Delta. The AM equation has received quite a bit of attention from physicists as the solutions exhibit very different features from the usual 2D turbulence [5]. One can arrive at (1.6) by starting from the so-called Charney–Hasegawa–Mima (CHM) equation

    tq+ψq=0,\begin{split}\partial_{t}q+\nabla^{\perp}\psi\cdot\nabla q=0,\end{split} (1.7)

    which is relevant for shallow water quasi-geostrophic dynamics. It is argued that this equation governs ocean front dynamics and planetary atmospheric pattern including Great Red Spot ([34, 29, 18]). Here, qq denotes the potential vorticity and is related by the stream function ψ\psi by q=(ΔLD2)ψq=(\Delta-L_{D}^{-2})\psi (so that (1.7) is nothing but (1.1) with Γ=(ΔLD2)1\Gamma=(\Delta-L_{D}^{-2})^{-1}). Here, LDL_{D} is the so-called Rossby deformation length and it is argued in [5] that when the characteristic length-scale of the flow LL satisfies LLDL\gg L_{D}, (1.6) can be obtained from (1.7) in the rescaled timescale tLD2tL_{D}^{2}.

1.3 Rough statement of the results

Our main result, which is stated roughly for now, gives strong illposedness for a large class of singular symbols, including all of the above three examples.

Theorem 1.1.

Consider the following symbols γ\gamma and pairs of exponents s,ss,s^{\prime}:

Multiplier Sobolev regularity exponents
γ=ξβ\gamma=\langle{\xi}\rangle^{\beta}, β>1\beta>1 s=s>3+32βs^{\prime}=s>3+\frac{3}{2}\beta
γ=ξ\gamma=\langle{\xi}\rangle s>92s^{\prime}>\frac{9}{2}
γ=ξβ\gamma=\langle{\xi}\rangle^{\beta}, β<1\beta<1 11βs>max{s+β22(1β),322+β1β}\frac{1}{1-\beta}s^{\prime}>\max\{s+\frac{\beta^{2}}{2(1-\beta)},\frac{3}{2}\frac{2+\beta}{1-\beta}\}
γ=logβ(10+|ξ|)\gamma=\log^{\beta}(10+|{\xi}|), β>0\beta>0 s=s>3s^{\prime}=s>3
γ=logβ(10+log(10+|ξ|))\gamma=\log^{\beta}(10+\log(10+|{\xi}|)), β>0\beta>0 s=s>3s^{\prime}=s>3

In each of the above cases, the Cauchy problem for (1.1) on the domain Ω=𝕋2\Omega=\mathbb{T}^{2} or 𝕋×\mathbb{T}\times\mathbb{R} is HsH^{s}-HsH^{s^{\prime}} ill-posed in the following sense: For any ϵ,δ,A>0\epsilon,\delta,A>0, there exists initial data θ0Cc(Ω)\theta_{0}\in C^{\infty}_{c}(\Omega) with θ0Hs<ϵ\|{\theta_{0}}\|_{H^{s}}<\epsilon such that either

  • there exists no solution θL([0,δ];Hs)\theta\in L^{\infty}([0,\delta];H^{s^{\prime}}) to (1.1) with θ|t=0=θ0\left.\theta\right|_{t=0}=\theta_{0}, or

  • any solution θ\theta belonging to L([0,δ];Hs)L^{\infty}([0,\delta];H^{s^{\prime}}) satisfy the growth

    supt[0,δ]θ(t,)Hs>A.\begin{split}\sup_{t\in[0,\delta]}\|{\theta(t,\cdot)}\|_{H^{s^{\prime}}}>A.\end{split}

Theorem 1.1 is a norm inflation result. When Ω=𝕋×\Omega=\mathbb{T}\times\mathbb{R} and s=ss=s^{\prime}, we may furthermore establish a non-existence result.

Theorem 1.2.

Consider a symbol γ\gamma and a real number ss such that γ\gamma and s=ss=s^{\prime} satisfy the hypothesis of Theorem 1.1. Then the Cauchy problem for (1.1) on the domain Ω=𝕋×\Omega=\mathbb{T}\times\mathbb{R} is ill-posed in the following sense: There exists an initial data set θ0Hs\theta_{0}\in H^{s} with arbitrarily small HsH^{s} norm, for which there does not exist a solution to the Cauchy problem for (1.1) in Lt([0,δ],Hs)L^{\infty}_{t}([0,\delta],H^{s^{\prime}}).

The above illposedness results suggest that one needs to be careful when working with models with singular multipliers: either an appropriate dissipative term must be supplied,222We note that while these models are often written without any dissipative terms in many physics texts, numerical simulations are always performed by adding very strong dissipative terms. or one should restrict to an appropriate class of functions (see Remark 1.4 below for more discussion). Indeed, local wellposedness of all the above singular examples, namely (1.2), (1.4), (1.5), and (1.6), has been obtained with appropriate dissipation terms: consider now

{tθ+uθ+κΥ(θ)=0,u=Γ[θ],\left\{\begin{split}&\partial_{t}\theta+u\cdot\nabla\theta+\kappa\Upsilon(\theta)=0,\\ &u=\nabla^{\perp}\Gamma[\theta],\end{split}\right. (1.8)

where κ>0\kappa>0 and Υ\Upsilon is a multiplier with strictly positive symbol. The authors of [8] have shown local regularity of the Ohkitani model with arbitrarily fractional dissipation (i.e. Υ=(Δ)ϵ\Upsilon=(-\Delta)^{\epsilon} for any ϵ>0\epsilon>0), which have been improved to any super-logarithmic dissipation in [10]. In the case of the E-MHD (and Hall-MHD), the works [9, 11] obtained local regularity with magnetic dissipation stronger than Λ\Lambda. A similar computation can be done for the AM, which then requires a dissipation term strictly stronger than Λ2\Lambda^{2}.

While these wellposedness results are obtained by rather standard Sobolev and commutator estimates, our illposedness results for the dissipative systems (1.8) show that these existing results are completely sharp:

Theorem 1.3.

Consider the following pairs of symbols γ,υ\gamma,\upsilon and exponents s,ss,s^{\prime}:

Multiplier Γ\Gamma Dissipation Υ\Upsilon Sobolev regularity exponents
γ=ξβ\gamma=\langle{\xi}\rangle^{\beta}, β>1\beta>1 υ=ξα\upsilon=\langle{\xi}\rangle^{\alpha}, α<β\alpha<\beta s1(βα)>s+β(βα)2(1(βα))\frac{s^{\prime}}{1-(\beta-\alpha)}>s+\frac{\beta(\beta-\alpha)}{2(1-(\beta-\alpha))}
γ=ξ\gamma=\langle{\xi}\rangle υ=ξα\upsilon=\langle{\xi}\rangle^{\alpha}, α<1\alpha<1 sα>s+1α2α\frac{s^{\prime}}{\alpha}>s+\frac{1-\alpha}{2\alpha}
γ=ξβ\gamma=\langle{\xi}\rangle^{\beta}, β<1\beta<1 υ=ξα\upsilon=\langle{\xi}\rangle^{\alpha}, α<β\alpha<\beta s1(βα)>s+β(βα)2(1(βα))\frac{s^{\prime}}{1-(\beta-\alpha)}>s+\frac{\beta(\beta-\alpha)}{2(1-(\beta-\alpha))}
γ=logβ(10+|ξ|)\gamma=\log^{\beta}(10+|{\xi}|), β>0\beta>0 υ=logα(10+|ξ|)\upsilon=\log^{\alpha}(10+|{\xi}|), α<β\alpha<\beta s=ss^{\prime}=s

The restrictions on ss and ss^{\prime} in the table are in addition to those from the nondissipative case. In each of the above cases, the Cauchy problem for (1.8) on the domain Ω=𝕋2\Omega=\mathbb{T}^{2} or 𝕋×\mathbb{T}\times\mathbb{R} is HsH^{s}-HsH^{s^{\prime}} ill-posed in the same sense as in Theorem 1.1.

Remark 1.4.

Our results establish illposedness of (1.1) for a large class of singular symbols in standard function spaces near the trivial solution θ̊=0\mathring{\theta}=0. Nevertheless, wellposedness in standard function spaces may still be possible around nontrivial background solutions, which are sometimes physically motivated. For instance, small data local wellposedness of E-MHD in weighted Sobolev space around θ̊=x2\mathring{\theta}=x_{2} – corresponding to a uniform magnetic field – may be established as an application of the techniques in [24].

1.4 Degenerate dispersive equations

It turns out that the linearization around degenerate shear steady states for (1.1) shows degenerate dispersion, which is the mechanism behind strong illposedness in the singular regime. Indeed, a large part of this work is devoted to the construction of degenerating wave packets for the class of linear equations

tϕ+ip(t,X,D)ϕ=0\begin{split}\partial_{t}\phi+ip(t,X,D)\phi=0\end{split} (1.9)

possibly with lower order terms, where pp is a time-dependent real pseudo-differential operator which is degenerate and dispersive. Once degenerating wave packets are constructed for all large frequencies, their time evolution essentially governs in which topologies the initial value problem for (1.9) could be well-posed. While our framework works for the general class of equations (1.9), we have chosen to focus on its applications towards the family of singular generalized SQG equations in this work, as the class of linearized equations arising from this family by varying Γ\Gamma and the steady state profile forms a representative class of (1.9).

Degenerate dispersive equations appear in a variety of physical contexts, besides those related with the gSQG equations described in the above. Primary examples include shallow water wave ([7, 3, 17]) and sedimentation models ([2, 6, 37, 32, 33]). Many of these models, most notably the Camassa–Holm and abcdabcd-Boussinesq equations (introduced in [7], [3] respectively and extensively studied since) feature principal terms which involve non-local and non-linear dispersion. More comprehensive list of physical systems involving degenerate dispersion (as well as related mathematical progress) is given in [1, 16, 21]. We shall review a few recent developments on the Cauchy problem for these type of equations, which are most directly relevant for the current work.

Well/Illposedness of K(m,n)K(m,n) equations and their variants. The family of K(m,n)K(m,n) equations introduced in [31, 30] is given by

tu+(un)xxx+(um)x=0.\begin{split}\partial_{t}u+(u^{n})_{xxx}+(u^{m})_{x}=0.\end{split} (1.10)

For n>1n>1, this model could be considered as the simplest equations featuring a quasilinear dispersive principal term (see [37] where this type of term appears for a model of particle suspensions). Various numerical simulations for this equation hinted at illposedness in strong topologies (see [1, 19, 15]). In the case of degenerate Airy equation tu+2uuxxx=0\partial_{t}u+2uu_{xxx}=0 (which is a further simplified model for n=2n=2), [1] gave illposedness in H2H^{2} of the initial value problem using explicit self-similar solution with scaling symmetries of the equation. We note that (uni-directional in time) illposedness for tu±xuxxx=0\partial_{t}u\pm xu_{xxx}=0 was obtained earlier in [14], based on the explicit solution formula. A general illposedness result, which works not only for (1.10) but also for many variants, was obtained in our recent work [21]. Here, the illposedness is deduced from the construction of degenerating wave packets for the linearized equation around degenerate solutions. On the other hand, [16] obtained a well-posedness result for certain variant of (1.10) in the case of “subcritically” degenerate (cf. [16, Section 1.5]) data. This well-posedness result is not contradictory to illposedness results from [21]; the solutions of [16] live in a suitably weighted space, which takes into account the rate of degeneracy of the solution.

Illposedness of the Hall- and electron-MHD systems. Based on construction of degenerating wave packets, the system (1.4) (as well as the Hall–MHD system) was shown to be strongly ill-posed near degenerate shear magnetic backgrounds, in the recent work [20] of the second and third authors. Indeed, the mechanism of illposedness for (1.4) is the same with the current paper. Additional difficulties arising in this work is that the system is non-local (opposed to (1.4)) and the symbol of Γ\Gamma could be only slightly singular, where the issue of well/illposedness becomes very delicate (as demonstrated explicitly in Theorem E).


The remainder of the introduction is organized as follows. In Section 1.5, we give precise statements of the main results of this paper, of which Theorems 1.11.2 are special cases. In Section 1.6, we present a toy model for the linearized equation which is almost explicitly solvable in the Fourier space yet contains the main features of the linearized dynamics. The model is obtained simply by dropping the sub-principal term and replacing the principal coefficient with a linear function. This solvable toy model demonstrates that the illposedness behavior is caused by degenerate dispersion, and gives the optimal growth rate of Sobolev norms that can be achieved. Furthermore, by comparing the toy model with actual linear equations, we explain the main difficulties in understanding the dynamics of linear equations. Then we end the introduction with an outline of the Organization of the Paper.

1.5 Main Results

We now give precise formulation of the main results of this paper.

Assumptions on Γ\Gamma. In what follows, we assume that γ\gamma is a smooth even positive symbol that satisfies the following properties for some Ξ0>1\Xi_{0}>1:

  1. 1.

    γ\gamma is elliptic and slowly varying: |ξIγ(ξ)|Iξ|I|γ(ξ)|{\partial_{\xi}^{I}\gamma(\xi)}|\lesssim_{I}\langle{\xi}\rangle^{-|{I}|}\gamma(\xi) for any multi-index II and ξ2\xi\in\mathbb{R}^{2}.

  2. 2.

    γ\gamma\nearrow\infty as |ξ||{\xi}|\nearrow\infty: γ(ξ1,ξ2)\gamma(\xi_{1},\xi_{2})\to\infty as |ξ||{\xi}|\to\infty.

  3. 3.

    ξ2ξ2γ\xi_{2}\partial_{\xi_{2}}\gamma is elliptic and slowly varying: |ξI(ξ2ξ2γ(ξ))|Iξ|I|ξ2ξ2γ(ξ)|{\partial_{\xi}^{I}(\xi_{2}\partial_{\xi_{2}}\gamma(\xi))}|\lesssim_{I}\langle{\xi}\rangle^{-|{I}|}\xi_{2}\partial_{\xi_{2}}\gamma(\xi) for any multi-index II and Ξ0|ξ1||ξ2|\Xi_{0}\leq|{\xi_{1}}|\leq|{\xi_{2}}|.

  4. 4.

    ξ2ξ2γ\xi_{2}\partial_{\xi_{2}}\gamma is almost comparable to γ\gamma: ξ2ξ2γ(ξ)1(log|ξ2|)2γ(ξ)\xi_{2}\partial_{\xi_{2}}\gamma(\xi)\gtrsim\frac{1}{(\log|{\xi_{2}}|)^{2}}\gamma(\xi) for Ξ0|ξ1||ξ2|\Xi_{0}\leq|{\xi_{1}}|\leq|{\xi_{2}}|.

A brief explanation of each assumption is in order:

  • Assumption 1 is a natural assumption that justifies, in particular, symbolic calculus (see Section 3).

  • Assumption 2 is a basic requirement for an arbitrarily fast frequency growth.

  • Assumptions 3 and 4 arise naturally in the control of the focusing of nearby bicharacteristics, which are put together to construct a suitable approximate solution to the linearized equation, called degenerating wave packets (see Sections 36). In particular, Assumption 4 allows us to quantify the scale of the degenerating wave packets (denoted by μ1\mu^{-1} later) in terms of its frequency (denoted by λ\lambda later). The factor (log|ξ2|)2(\log|{\xi_{2}}|)^{-2} is somewhat arbitrary but fixed for simplicity.

Many natural choices of γ\gamma satisfy the above assumptions, including γ(ξ)=ξβ\gamma(\xi)=\langle{\xi}\rangle^{\beta} for any β>0\beta>0, γ(ξ)=exp(βlogα(10+|ξ|))\gamma(\xi)=\exp(\beta\log^{\alpha}(10+|{\xi}|)) for any β>0\beta>0 and 0<α<10<\alpha<1, γ(ξ)=logβ(10+|ξ|)\gamma(\xi)=\log^{\beta}(10+|{\xi}|) for any β>0\beta>0, γ(ξ)=logβ(10+logα(10+|ξ|))\gamma(\xi)=\log^{\beta}(10+\log^{\alpha}(10+|{\xi}|)) for any α,β>0\alpha,\beta>0 etc.

By Assumption 1, there exists β0>0\beta_{0}>0 such that

supξ:|ξ||ξ|2|ξ|γ(ξ)2β0γ(ξ) for |ξ|Ξ0.\sup_{\xi^{\prime}:|{\xi}|\leq|{\xi^{\prime}}|\leq 2|{\xi}|}\gamma(\xi^{\prime})\leq 2^{\beta_{0}}\gamma(\xi)\quad\hbox{ for }|{\xi}|\geq\Xi_{0}. (1.11)

For the remainder of this paper, we fix one such β0>0\beta_{0}>0 and let other constants depend on it. Iterating this bound, it follows that

γ(ξ)|ξ|β0|Ξ0|β0γ(Ξ0) for |ξ|Ξ0.\gamma(\xi)\lesssim|{\xi}|^{\beta_{0}}|{\Xi_{0}}|^{-\beta_{0}}\gamma(\Xi_{0})\quad\hbox{ for }|{\xi}|\geq\Xi_{0}.

The infimum of possible β0\beta_{0}’s (among all real numbers) is called the order of γ\gamma. The justification for this terminology comes from the property that if γ=|ξ|β0\gamma=|{\xi}|^{\beta_{0}}, then β0\beta_{0} is its order in the usual sense.

Assumptions on Υ\Upsilon. For the dissipative operator Υ\Upsilon, we simply assume that its symbol υ\upsilon is a smooth even positive symbol that is elliptic and slowly varying, in the sense that

|ξIυ(ξ)|I|ξ||I|υ(ξ).|{\partial_{\xi}^{{I}}\upsilon(\xi)}|\lesssim_{{I}}|{\xi}|^{-|{{I}}|}\upsilon(\xi). (1.12)

As in the case of γ\gamma, there exists α0>0\alpha_{0}>0 such that

supξ:|ξ||ξ|2|ξ|υ(ξ)2α0υ(ξ) for |ξ|Ξ0.\sup_{\xi^{\prime}:|{\xi}|\leq|{\xi^{\prime}}|\leq 2|{\xi}|}\upsilon(\xi^{\prime})\leq 2^{\alpha_{0}}\upsilon(\xi)\quad\hbox{ for }|{\xi}|\geq\Xi_{0}. (1.13)

For the remainder of this paper, we fix one such α0>0\alpha_{0}>0 and let other constants depend on it.

Assumptions on θ̊\mathring{\theta}. Next, we specify the class of shear states that will be proved to be unstable, in linear/nonlinear settings and with/without dissipation, in high-regularity Sobolev spaces.

In the absence of dissipation, recall that any shear steady state θ=f(x2)\theta=f(x_{2}) (with reasonable regularity assumptions) solves (1.1). We shall assume that ff is smooth, bounded and has a quadratic degeneracy in the following sense:

Definition 1.5.

We say that a shear steady state θ̊=f(x2)\mathring{\theta}=f(x_{2}) for (1.1) is quadratically degenerate at x̊2(𝕋,)\mathring{x}_{2}\in(\mathbb{T},\mathbb{R}) if

f(x̊2)=0,f′′(x̊2)0.f^{\prime}(\mathring{x}_{2})=0,\qquad f^{\prime\prime}(\mathring{x}_{2})\neq 0.
Remark 1.6.

Our method easily extends to the case when the order of vanishing of f(x2)f^{\prime}(x_{2}) at x2=x̊2x_{2}=\mathring{x}_{2} is generalized to any positive real number. The general heuristic principle is that the slower the vanishing (i.e., the lower the order), the faster the frequency growth. The quadratically degenerate case considered in Definition 1.5 is distinguished by the fact that it is the generic order for a smooth f(x2)f^{\prime}(x_{2}). We also note that the boundedness assumption can be readily generalized to a polynomial growth condition at infinity.

In the presence of dissipation, equation (1.8) for a shear state θ̊=f(t,x2)\mathring{\theta}=f(t,x_{2}) reduces to

(t+κΥ)f(t,x2)=0,(\partial_{t}+\kappa\Upsilon)f(t,x_{2})=0, (1.14)

where Υ\Upsilon obeys the assumptions made above. By Fourier analysis, (1.14) is clearly well-posed forward in time in Hs((𝕋,)x2)H^{s}((\mathbb{T},\mathbb{R})_{x_{2}}) for any ss\in\mathbb{R}. Moreover, for any well-posed solution ff, if f(0,x2)f(0,x_{2}) is even then so is f(t,x2)f(t,x_{2}) for each t>0t>0. In what follows, we shall take as our background shear state a smooth bounded solution θ̊=f(t,x2)\mathring{\theta}=f(t,x_{2}) such that f0(x2)=f(0,x2)f_{0}(x_{2})=f(0,x_{2}) is even and f0′′(0)0{f_{0}^{\prime\prime}(0)}\neq 0.

Remark 1.7.

The evenness assumption brings a technical simplification in our argument, as the degenerate point x2=0x_{2}=0 is then fixed in tt. Our methods may be extended to the case when this assumption is removed, in which case the degenerate point may move in time, but at the price of additional technical constraints on the length of the time interval.

Linear results. We begin by stating our main results concerning the linearization of (1.1) and (1.8) around θ̊\mathring{\theta} introduced above. The direct linearization of (1.1) around θ̊\mathring{\theta} is given by Lθ̊ϕ=0L_{\mathring{\theta}}\phi=0, where

Lθ̊ϕ=tϕθ̊Γϕ+Γθ̊ϕ,L_{\mathring{\theta}}\phi=\partial_{t}\phi-\nabla^{\perp}\mathring{\theta}\cdot\nabla\Gamma\phi+\nabla^{\perp}\Gamma\mathring{\theta}\cdot\nabla\phi, (1.15)

where as for (1.8), the linearization around θ̊\mathring{\theta} takes the form Lθ̊(κ)ϕ=0L^{(\kappa)}_{\mathring{\theta}}\phi=0, where

Lθ̊(κ)ϕ=tϕθ̊Γϕ+Γθ̊ϕ+κΥϕ.L^{(\kappa)}_{\mathring{\theta}}\phi=\partial_{t}\phi-\nabla^{\perp}\mathring{\theta}\cdot\nabla\Gamma\phi+\nabla^{\perp}\Gamma\mathring{\theta}\cdot\nabla\phi+\kappa\Upsilon\phi. (1.16)

To formulate a linear illposedness result of the desired generality and precision, it is convenient to introduce the following set of parameters. Given λ0>1\lambda_{0}>1, called the initial frequency parameter and M>1M>1, called the frequency growth factor, define the corresponding (normalized) frequency growth time τM\tau_{M} to be333The definition of τM\tau_{M}, (1.17), is motivated by the consideration of bicharacteristic curves associated with the principal symbol of (2.1).

τM:=λ0Mλ01γ(λ0,λ)dλλ0.\tau_{M}:=\int_{\lambda_{0}}^{M\lambda_{0}}\frac{1}{\gamma(\lambda_{0},\lambda)}\frac{\mathrm{d}\lambda}{\lambda_{0}}. (1.17)

Justification of the formula (1.17) shall be given later in Section 2. In addition to these parameters, we also fix an arbitrarily small parameter 0<δ0<11000<\delta_{0}<\frac{1}{100} and a nonnegative parameter 0σ012δ00\leq\sigma_{0}\leq\frac{1}{2}-\delta_{0}, which shall be used in the conditions that λ0,M\lambda_{0},M need to satisfy.

Non-dissipative case. Roughly speaking, our main linear result in the non-dissipative case, Theorem A, states that given λ0\lambda_{0} and MM obeying a suitable condition (see (1.18)–(1.20) below), there exists an initial data set ϕ0\phi_{0} for the linearized equation Lθ̊ϕ=0L_{\mathring{\theta}}\phi=0 with frequency O(λ0)O(\lambda_{0}) such that any corresponding solution exhibits frequency growth by factor MM in time Of(τM)O_{f}(\tau_{M}).

Theorem A (Linear illposedness, non-dissipative case).

Let θ̊=f(x2)\mathring{\theta}=f(x_{2}) be a smooth bounded shear steady state that is quadratically degenerate at x2=x̊2x_{2}=\mathring{x}_{2} and fix a small parameter 0<δ0<11000<\delta_{0}<\frac{1}{100} and 0σ013(12δ0)0\leq\sigma_{0}\leq\frac{1}{3}(1-2\delta_{0}). Then there exist Λ0=Λ0(f,γ,δ0,σ0)>1\Lambda_{0}=\Lambda_{0}(f,\gamma,\delta_{0},\sigma_{0})>1 and T0=T0(f,γ,δ0,σ0)>0T_{0}=T_{0}(f,\gamma,\delta_{0},\sigma_{0})>0 such that the following holds. For each λ0\lambda_{0}\in\mathbb{N} such that λ0Λ0\lambda_{0}\geq\Lambda_{0}, τMT0\tau_{M}\leq T_{0} and M>1M>1 satisfying the growth conditions

supM[1,M]γ(λ0,Mλ0)MτM\displaystyle\sup_{M^{\prime}\in[1,M]}\frac{\gamma(\lambda_{0},M^{\prime}\lambda_{0})}{{M^{\prime}}}\tau_{M^{\prime}} min{γ(λ0,λ0)1δ0,λ0σ0},\displaystyle\leq{\min\{\gamma(\lambda_{0},\lambda_{0})^{1-\delta_{0}},\lambda_{0}^{\sigma_{0}}\}}, (1.18)
τM\displaystyle\tau_{M} min{λ01δ03σ0γ(λ0,λ0),1},\displaystyle\leq\min\left\{\frac{\lambda_{0}^{1-\delta_{0}{-3\sigma_{0}}}}{\gamma(\lambda_{0},\lambda_{0})},1\right\}, (1.19)
M\displaystyle M λ01δ0,\displaystyle\leq\lambda_{0}^{\frac{1}{\delta_{0}}}, (1.20)

there exists a smooth function ϕ0\phi_{0} such that

ϕ0HsC0;s,sλ0ssϕ0Hs for any ss,\|{{\phi_{0}}}\|_{H^{s^{\prime}}}\leq{C_{0;s^{\prime},s}}\lambda_{0}^{s^{\prime}-s}\|{{\phi_{0}}}\|_{H^{s}}\quad\hbox{ for any }s^{\prime}\geq s, (1.21)

yet any Γ12L2\Gamma^{-\frac{1}{2}}L^{2}-solution ϕ\phi to Lθ̊ϕ=0L_{\mathring{\theta}}\phi=0 on [0,100991|f′′(x̊2)|τM]\left[0,\tfrac{100}{99}\tfrac{1}{|{f^{\prime\prime}(\mathring{x}_{2})}|}\tau_{M}\right] with ϕ|t=0=ϕ0\left.\phi\right|_{t=0}=\phi_{0} obeys

supt[0,100991|f′′(x̊2)|τM]ϕ(t,)HsCs,sγ(λ0,λ0)12γ(λ0,Mλ0)12Msλ0ssϕ0Hs for any s,s>0.\sup_{t\in\left[0,\tfrac{100}{99}\tfrac{1}{|{f^{\prime\prime}(\mathring{x}_{2})}|}\tau_{M}\right]}\|{{\phi}(t,\cdot)}\|_{H^{s^{\prime}}}\geq C_{s^{\prime},s}{\frac{\gamma(\lambda_{0},\lambda_{0})^{\frac{1}{2}}}{\gamma(\lambda_{0},M\lambda_{0})^{\frac{1}{2}}}}M^{s^{\prime}}\lambda_{0}^{s^{\prime}-s}\|{{\phi_{0}}}\|_{H^{s}}\quad\hbox{ for any }s,s^{\prime}>0. (1.22)

A Γ12L2\Gamma^{-\frac{1}{2}}L^{2} solution to Lθ̊ϕL_{\mathring{\theta}}\phi is a natural notion of a weak solution in view of the energy structure of Lθ̊L_{\mathring{\theta}}; we postpone its precise definition to Section 2 below.

Remark 1.8.

Condition (1.18) is used to control the evolution of the frequency of the wave packet. The condition τMγ(λ0,λ0)1λ01δ03σ0\tau_{M}\leq\gamma(\lambda_{0},\lambda_{0})^{-1}\lambda_{0}^{1-\delta_{0}-3\sigma_{0}} in (1.19) arises naturally from the error estimate (see Section 6.3), while τM1\tau_{M}\leq 1 is natural in view of the local-in-time energy argument employed in the proof (see Section 7). Condition (1.20) is a mild technical condition that is assumed to simplify the control of some non-main error terms; we expect that it can be removed for specific γ\gamma’s (see, for instance, [20] for electron and Hall MHD, where the frequency growth factor could be of size ecλ0e^{c\lambda_{0}}).

Remark 1.9.

We note that the LHS of (1.18) is uniformly bounded when the order β0\beta_{0} of γ\gamma is less than 11, in which case we only need (1.19) with σ0=0\sigma_{0}=0. In fact, the parameter σ0\sigma_{0} is simply a device introduced to handle (in an non-optimal way) the case β01\beta_{0}\geq 1, which is “supercritical” in many ways; see the discussion following Corollary 1.13. If we assume that f′′′f^{\prime\prime\prime} vanishes to a high order at x̊2\mathring{x}_{2}, then the factor 33 in front of σ0\sigma_{0} in (1.19) can be lowered to any number greater than 22 , and the condition σ013(12δ0)\sigma_{0}\leq\frac{1}{3}(1-2\delta_{0}) would be relaxed accordingly; see Remark 4.2.

Remark 1.10.

In the proof, it will be shown that indeed the growth (1.22) is generic; it occurs for an open set of initial data that have large frequency in both x1,x2x_{1},x_{2} and supported in a neighborhood of the degeneracy of ff.

We now discuss the implications of Theorem A. We first demonstrate that under our assumptions on γ\gamma, the linearized equation around any smooth shear steady state with a quadratic degeneracy is always ill-posed in HsH^{s} for any s>β02s>{\frac{\beta_{0}}{2}}. As λγ(λ0,λ)\lambda\mapsto\gamma(\lambda_{0},\lambda) is increasing on [λ0,)[\lambda_{0},\infty), we have the obvious bound

τMMγ(λ0,λ0) for λ0 sufficiently large.\tau_{M}\leq\frac{M}{\gamma(\lambda_{0},\lambda_{0})}\quad\hbox{ for $\lambda_{0}$ sufficiently large.}

As a result, if we set

M(λ0)=min{γ(λ0,λ0)1δ0,λ0δ0,γ(λ0,λ0)12δ0β0,λ015δ03β0},M(\lambda_{0})={\min\left\{\gamma(\lambda_{0},\lambda_{0})^{1-\delta_{0}},\lambda_{0}^{\delta_{0}},\,\gamma(\lambda_{0},\lambda_{0})^{\frac{1-2\delta_{0}}{\beta_{0}}},\lambda_{0}^{\frac{1-5\delta_{0}}{3\beta_{0}}}\right\}},

then (1.19) (with σ0=13(12δ0)\sigma_{0}=\frac{1}{3}(1-2\delta_{0})) and (1.20) are satisfied for sufficiently large λ0\lambda_{0}. To check (1.18), we estimate

γ(λ0,Mλ0)Mmin{γ(λ0,λ0)1δ0,λ013(12δ0)}τM\displaystyle\frac{\gamma(\lambda_{0},M^{\prime}\lambda_{0})}{M^{\prime}\min\{\gamma(\lambda_{0},\lambda_{0})^{1-\delta_{0}},\lambda_{0}^{\frac{1}{3}(1-2\delta_{0})}\}}\tau_{M^{\prime}} C(M)β0γ(λ0,λ0)min{γ(λ0,λ0)1δ0,λ013(12δ0)}1γ(λ0,λ0)\displaystyle\leq C(M^{\prime})^{\beta_{0}}\frac{\gamma(\lambda_{0},\lambda_{0})}{\min\{\gamma(\lambda_{0},\lambda_{0})^{1-\delta_{0}},\lambda_{0}^{\frac{1}{3}(1-2\delta_{0})}\}}\frac{1}{\gamma(\lambda_{0},\lambda_{0})}
Cmin{γ(λ0,λ0)δ0,λ0δ0}1,\displaystyle\leq\frac{C}{\min\{\gamma(\lambda_{0},\lambda_{0})^{\delta_{0}},\lambda_{0}^{\delta_{0}}\}}\leq 1,

using the above bound for τM\tau_{M^{\prime}} and M(λ0)γ(λ0,λ0)12δ0β0,λ015δ03β0M(\lambda_{0})\leq\gamma(\lambda_{0},\lambda_{0})^{\frac{1-2\delta_{0}}{\beta_{0}}},\lambda_{0}^{\frac{1-5\delta_{0}}{3\beta_{0}}}. Furthermore, observe that M(λ0)M(\lambda_{0})\to\infty and τM(λ0)0\tau_{M(\lambda_{0})}\to 0 as λ0\lambda_{0}\to\infty. Finally, by (1.11), note that

γ(λ0,λ0)12γ(λ0,Mλ0)12Mβ02,\frac{\gamma(\lambda_{0},\lambda_{0})^{\frac{1}{2}}}{\gamma(\lambda_{0},M\lambda_{0})^{\frac{1}{2}}}\gtrsim M^{-\frac{\beta_{0}}{2}}, (1.23)

so the RHS of (1.22) is increasing in MM for s>β02s^{\prime}>\frac{\beta_{0}}{2}. We have therefore proved:

Corollary 1.11 (Linear HsH^{s} illposedness for any s>β02s>{\frac{\beta_{0}}{2}}).

Let θ̊=f(x2)\mathring{\theta}=f(x_{2}) be a smooth bounded shear steady state that is quadratically degenerate at x2=x̊2x_{2}=\mathring{x}_{2}. For any s>β02s>{\frac{\beta_{0}}{2}}, there exists a sequence ϕ(n)0{\phi_{(n)0}} of initial data sets and times T(n)T_{(n)} such that

ϕ(n)0Hs1,0<T(n)1,T(n)0,\|{{\phi_{(n)0}}}\|_{H^{s}}\leq 1,\qquad 0<T_{(n)}\leq 1,\qquad T_{(n)}\searrow 0,

yet for any sequence ϕ(n){\phi_{(n)}} of Γ12L2\Gamma^{-\frac{1}{2}}L^{2}-solutions to Lθ̊ϕ=0L_{\mathring{\theta}}\phi=0 on [0,T(n)][0,T_{(n)}] with ϕ(n)|t=0=ϕ(n)0\left.\phi_{(n)}\right|_{t=0}=\phi_{(n)0}, we have

supt[0,T(n)]ϕ(n)(t,)Hsϕ(n)0Hs as n.\frac{\sup_{t\in[0,T_{(n)}]}\|{{\phi_{(n)}}(t,\cdot)}\|_{H^{s}}}{\|{{\phi_{(n)0}}}\|_{H^{s}}}\to\infty\quad\hbox{ as }n\to\infty.

On the other hand, an inspection of the size of the optimal frequency growth factor MM for various model cases (see Corollary 1.13) shows that the instability mechanism in hand is stronger for a multiplier γ\gamma with faster growth. One way to make quantify this idea is to introduce the following notion:

Definition 1.12.

We say that the shear steady state θ̊\mathring{\theta} with a quadratic degeneracy is linearly HsH^{s}-HsH^{s^{\prime}} unstable by degenerate dispersion according to Theorem A if there exist a sequence (M(n),λ0(n))(M_{(n)},\lambda_{0(n)}) satisfying (1.18)–(1.20) and λ0(n)Λ0\lambda_{0(n)}\geq\Lambda_{0} such that

γ(λ0(n),λ0(n))12γ(λ0(n),M(n)λ0(n))12M(n)sλ0(n)ss,τM(n)0.{\frac{\gamma(\lambda_{0(n)},\lambda_{0(n)})^{\frac{1}{2}}}{\gamma(\lambda_{0(n)},M_{(n)}\lambda_{0(n)})^{\frac{1}{2}}}}M_{(n)}^{s^{\prime}}\lambda_{0(n)}^{s^{\prime}-s}\to\infty,\qquad\tau_{M_{(n)}}\to 0.

In what follows, we shall often drop the proviso according to Theorem A. Given a shear steady state θ̊\mathring{\theta} that is linearly HsH^{s}-HsH^{s^{\prime}} unstable by degenerate dispersion, no mapping from HsH^{s} to a set of Γ12L2\Gamma^{-\frac{1}{2}}L^{2}-solutions to Lθ̊ϕ=0L_{\mathring{\theta}}\phi=0 in Lt([0,δ],Hs)L^{\infty}_{t}([0,\delta],H^{s^{\prime}}) can be bounded thanks to (1.22) in Theorem A. Note that if Definition 1.12 holds for one shear steady state θ̊\mathring{\theta} with a quadratic degeneracy, then it holds for any other such shear steady states; hence, Definition 1.12 is a property of the system (1.1) (more precisely, of γ\gamma).

Next, we specialize γ\gamma to several model cases and compute the (essentially) optimal growth factor given by Theorem A.

Corollary 1.13.

Let θ̊=f(x2)\mathring{\theta}=f(x_{2}) be a smooth shear steady state that is quadratically degenerate at x2=x̊2x_{2}=\mathring{x}_{2}. In each of the following cases, Theorem A applies with the specified choice of MM for any 0σ<10\leq\sigma<1, provided that δ0\delta_{0} and σ0\sigma_{0} are chosen appropriately depending on σ\sigma and β\beta; moreover, in each case, τM0\tau_{M}\to 0 as λ0\lambda_{0}\to\infty. As a result, (1.1) is linearly HsH^{s}-HsH^{s^{\prime}} unstable by degenerate dispersion for the specified values of s,ss,s^{\prime}.

Multiplier Γ\Gamma Freq. growth MM Lin. HsH^{s}-HsH^{s^{\prime}} inst.
γ=ξβ\gamma=\langle{\xi}\rangle^{\beta}, β>1\beta>1 M=λ0σ13(β1)M=\lambda_{0}^{\sigma\frac{1}{3(\beta-1)}} s,s>β2s^{\prime},s>\frac{\beta}{2}, (1+13(β1))s>s+β6(β1){(1+\frac{1}{3(\beta-1)})s^{\prime}>s+\frac{\beta}{6(\beta-1)}}
γ=ξ\gamma=\langle{\xi}\rangle M=λ011σM=\lambda_{0}^{\frac{1}{1-\sigma}} s,s>12s^{\prime},s>\frac{1}{2}
γ=ξβ\gamma=\langle{\xi}\rangle^{\beta}, β<1\beta<1 M=λ0σβ1βM=\lambda_{0}^{\sigma\frac{\beta}{1-\beta}} s,s>β2s^{\prime},s>\frac{\beta}{2}, 11βs>s+β22(1β){\frac{1}{1-\beta}s^{\prime}>s+\frac{\beta^{2}}{2(1-\beta)}}
γ=logβ(10+|ξ|)\gamma=\log^{\beta}(10+|{\xi}|), β>0\beta>0 M=logσβλ0M=\log^{\sigma\beta}\lambda_{0} s=s>0s^{\prime}=s>0
γ=logβ(10+log(10+|ξ|))\gamma=\log^{\beta}(10+\log(10+|{\xi}|)), β>0\beta>0 M=logσβlogλ0M=\log^{\sigma\beta}\log\lambda_{0} s=s>0s^{\prime}=s>0

Corollary 1.13 makes quantitative the expectation that the faster growth of γ\gamma, the stronger the instability by degenerate dispersion.444The case ξβ\langle{\xi}\rangle^{\beta} for β>1\beta>1 is an exception, but this seems to be due to the inefficiency of our method in this case.

Dissipative case. We now state the main linear result in the dissipative case. Let λ0\lambda_{0}, MM and τM\tau_{M} be defined as before. We introduce two small constants 0<δ1δ0<11000<\delta_{1}\leq\delta_{0}<\frac{1}{100}. In place of |f′′(0)|1τ|{f^{\prime\prime}(0)}|^{-1}\tau that arose in Theorem A, we define the function tf(τ)t_{f}(\tau) by the relation

0tf(τ)|f′′(t,0)|dt=τ.\int_{0}^{t_{f}(\tau)}|{f^{\prime\prime}(t^{\prime},0)}|\,\mathrm{d}t^{\prime}=\tau.

Note that f′′f^{\prime\prime} is time-independent, then tf(τ)=|f′′(0)|1τt_{f}(\tau)=|{f^{\prime\prime}(0)}|^{-1}\tau.

Theorem B (Linear illposedness, dissipative case).

Let θ̊0=f0(x2)\mathring{\theta}_{0}=f_{0}(x_{2}) a smooth even function with f0′′(0)0f_{0}^{\prime\prime}(0)\neq 0, let θ̊=f(t,x2)\mathring{\theta}=f(t,x_{2}) be the smooth solution to (1.14) with f(0,x2)=f0(x2)f(0,x_{2})=f_{0}(x_{2}) and fix small parameters 0<δ1δ0<11000<\delta_{1}\leq\delta_{0}<\frac{1}{100} and a parameter 0σ013(12δ0)0\leq\sigma_{0}\leq\frac{1}{3}(1-2\delta_{0}). Then there exist Λ1=Λ1(f,γ,υ,δ0,δ1,σ0)\Lambda_{1}=\Lambda_{1}({f,\gamma,\upsilon,\delta_{0},\delta_{1},\sigma_{0}}) and T1=T1(f,γ,υ,δ0,δ1,σ0)>0T_{1}=T_{1}({f,\gamma,\upsilon,\delta_{0},\delta_{1},\sigma_{0}})>0 such that the following holds. For each λ0\lambda_{0}\in\mathbb{N} such that λ0Λ1\lambda_{0}\geq\Lambda_{1}, τMT1\tau_{M}\leq T_{1} and M>1M>1 satisfying the nondissipative growth conditions (1.18)–(1.20), as well as the conditions

λ0Mλ0υ(λ0,λ)γ(λ0,λ)dλλ0=oλ0(1),\displaystyle\int_{\lambda_{0}}^{M\lambda_{0}}\frac{\upsilon(\lambda_{0},\lambda)}{\gamma(\lambda_{0},\lambda)}\frac{\mathrm{d}\lambda}{\lambda_{0}}=o_{\lambda_{0}}(1), (1.24)
supM[1,M]λ0Mλ0(ξ2γ(λ0,λ)ξ2γ(λ0,Mλ0))1δ1γ(λ0,λ0)γ(λ0,λ)2dλλ01,\displaystyle\sup_{M^{\prime}\in[1,M]}\int_{\lambda_{0}}^{M^{\prime}\lambda_{0}}\left(\frac{\partial_{\xi_{2}}\gamma(\lambda_{0},\lambda)}{\partial_{\xi_{2}}\gamma(\lambda_{0},M^{\prime}\lambda_{0})}\right)^{1-\delta_{1}}\frac{\gamma(\lambda_{0},\lambda_{0})}{\gamma(\lambda_{0},\lambda)^{2}}\frac{\mathrm{d}\lambda}{\lambda_{0}}\leq 1, (1.25)

there exists a smooth function ϕ0{\phi_{0}} such that

ϕ0HsCs,sλ0ssϕ0Hs for any ss,\|{{\phi_{0}}}\|_{H^{s^{\prime}}}\leq{C_{s,s^{\prime}}}\lambda_{0}^{s^{\prime}-s}\|{{\phi_{0}}}\|_{H^{s}}\quad\hbox{ for any }s^{\prime}\geq s, (1.26)

yet any Γ12L2\Gamma^{-\frac{1}{2}}L^{2}-solution ϕ\phi to Lθ̊(κ)ϕ=0L^{(\kappa)}_{\mathring{\theta}}\phi=0 on [0,10099tf(τM)]\left[0,\tfrac{100}{99}t_{f}(\tau_{M})\right] with ϕ(t=0)=ϕ0\phi(t=0)=\phi_{0} obeys

supt[0,10099tf(τM)]ϕ(t,)HsCs,sγ(λ0,λ0)12γ(λ0,Mλ0)12Msλ0ssϕ0Hs for any s>0.\sup_{t\in\left[0,\tfrac{100}{99}t_{f}(\tau_{M})\right]}\|{{\phi}(t,\cdot)}\|_{H^{s^{\prime}}}\geq C_{s^{\prime},s}{\frac{\gamma(\lambda_{0},\lambda_{0})^{\frac{1}{2}}}{\gamma(\lambda_{0},M\lambda_{0})^{\frac{1}{2}}}}M^{s^{\prime}}\lambda_{0}^{s^{\prime}-s}\|{{\phi_{0}}}\|_{H^{s}}\quad\hbox{ for any }s^{\prime}>0. (1.27)
Remark 1.14.

Condition (1.24) arises naturally from the contribution of the dissipative term in the error estimate in our degenerating wave packet construction. Condition (1.25) is a technical condition arising due to the time dependence of θ̊=f(t,x2)\mathring{\theta}=f(t,x_{2}). We note that when δ1=0\delta_{1}=0 and τM1\tau_{M}\leq 1, (1.25) already holds up to a logarithmic power of λ0\lambda_{0} (from Assumption 4 for γ\gamma); see (4.53) and (4.56).

We introduce the following analogue of Definition 1.12:

Definition 1.15.

We say that the shear state θ̊\mathring{\theta} satisfying the assumptions of Theorem B is linearly HsH^{s}-HsH^{s^{\prime}} unstable by degenerate dispersion according to Theorem B if there exist a sequence (M(n),λ0(n))(M_{(n)},\lambda_{0(n)}) satisfying (1.18)–(1.20), (1.24)–(1.25) and λ0(n)Λ1\lambda_{0(n)}\geq\Lambda_{1} such that

γ(λ0(n),λ0(n))12γ(λ0(n),M(n)λ0(n))12M(n)sλ0(n)ss,τM(n)0.{\frac{\gamma(\lambda_{0(n)},\lambda_{0(n)})^{\frac{1}{2}}}{\gamma(\lambda_{0(n)},M_{(n)}\lambda_{0(n)})^{\frac{1}{2}}}}M_{(n)}^{s^{\prime}}\lambda_{0(n)}^{s^{\prime}-s}\to\infty,\qquad\tau_{M_{(n)}}\to 0.

In what follows, we shall often drop the proviso according to Theorem B. As before, Theorem B implies that given a shear state θ̊\mathring{\theta} that is linearly HsH^{s}-HsH^{s^{\prime}} unstable by degenerate dispersion, no mapping from HsH^{s} to a set of Γ12L2\Gamma^{-\frac{1}{2}}L^{2}-solutions to Lθ̊ϕ=0L_{\mathring{\theta}}\phi=0 in Lt([0,δ],Hs)L^{\infty}_{t}([0,\delta],H^{s^{\prime}}) can be bounded. Moreover, Definition 1.15 is a property of the system (1.8) (more precisely, of γ\gamma, κυ\kappa\upsilon) in the same sense as before.

We now specialize γ\gamma and υ\upsilon to several model cases and exhibit instances of illposedness in the dissipative case given by Theorem B.

Corollary 1.16.

Fix κ>0\kappa>0, let θ̊0=f0(x2)\mathring{\theta}_{0}=f_{0}(x_{2}) a smooth odd function with f0(0)0f_{0}^{\prime}(0)\neq 0, and let θ̊=f(t,x2)\mathring{\theta}=f(t,x_{2}) be the smooth solution to (1.14). In each of the following cases, Theorem B applies with the specified choice of MM for any 0σ<10\leq\sigma<1, provided that λ0\lambda_{0} is sufficiently large depending on σ\sigma; moreover, in each case, τM0\tau_{M}\to 0 as λ0\lambda_{0}\to\infty. As a result, (1.8) is linearly HsH^{s}-HsH^{s^{\prime}} unstable by degenerate dispersion for the specified values of ss^{\prime}. In the table below, the restrictions on MM and (s,s)(s,s^{\prime}) are in addition to those from the nondissipative case.

Multiplier Γ\Gamma Dissipation Υ\Upsilon Freq. growth MM Lin. HsH^{s}-HsH^{s^{\prime}} inst.
γ=ξβ\gamma=\langle{\xi}\rangle^{\beta}, β>1\beta>1 υ=ξα\upsilon=\langle{\xi}\rangle^{\alpha}, α<β\alpha<\beta M=λ0σβα1(βα)M=\lambda_{0}^{\sigma\frac{\beta-\alpha}{1-(\beta-\alpha)}} s1(βα)>s+β(βα)2(1(βα))\frac{s^{\prime}}{1-(\beta-\alpha)}>s+\frac{\beta(\beta-\alpha)}{2(1-(\beta-\alpha))}
γ=ξ\gamma=\langle{\xi}\rangle υ=ξα\upsilon=\langle{\xi}\rangle^{\alpha}, α<1\alpha<1 M=λ0σ1ααM=\lambda_{0}^{\sigma\frac{1-\alpha}{\alpha}} sα>s+1α2α\frac{s^{\prime}}{\alpha}>s+\frac{1-\alpha}{2\alpha}
γ=ξβ\gamma=\langle{\xi}\rangle^{\beta}, β<1\beta<1 υ=ξα\upsilon=\langle{\xi}\rangle^{\alpha}, α<β\alpha<\beta M=λ0σβα1(βα)M=\lambda_{0}^{\sigma\frac{\beta-\alpha}{1-(\beta-\alpha)}} s1(βα)>s+β(βα)2(1(βα))\frac{s^{\prime}}{1-(\beta-\alpha)}>s+\frac{\beta(\beta-\alpha)}{2(1-(\beta-\alpha))}
γ=logβ(10+|ξ|)\gamma=\log^{\beta}(10+|{\xi}|), β>0\beta>0 υ=logα(10+|ξ|)\upsilon=\log^{\alpha}(10+|{\xi}|), α<β\alpha<\beta M=logσ(βα)λ0M=\log^{\sigma(\beta-\alpha)}\lambda_{0} s=ss^{\prime}=s

The illposedness results in Corollary 1.16 are essentially sharp, in that it is not difficult to prove using standard energy estimates that in each case, Lθ̊(κ)ϕ=0L^{(\kappa)}_{\mathring{\theta}}\phi=0 is locally well-posed in HsH^{s} for any s0s\geq 0 if the dissipative exponent β\beta is greater than α\alpha.

Nonlinear illposedness results. Remarkably, the linear norm growth results (Theorems A and B) may be extended to corresponding norm inflation properties of the nonlinear Cauchy problem.

Theorem C (Nonlinear illposedness).

Assume that a shear steady state θ̊\mathring{\theta} with a quadratic degeneracy (resp. a shear state θ̊\mathring{\theta} satisfying the assumptions of Theorem B) is linearly HsH^{s}-HsH^{s^{\prime}} unstable by degenerate dispersion according to Theorem A (resp. Theorem B) with s>32β0+3s^{\prime}>{\frac{3}{2}}\beta_{0}+3. Then θ̊\mathring{\theta} is nonlinearly HsHsH^{s}\to H^{s^{\prime}} ill-posed with respect to (1.1) (resp. (1.8)) in the following sense: For any ϵ,δ,A>0\epsilon,\delta,A>0, there exists initial data θ0Cc(Ω)\theta_{0}\in C^{\infty}_{c}(\Omega) with θ0Hs<ϵ{\|{\theta_{0}}\|_{H^{s}}}<\epsilon such that either

  • there exists no solution θθ̊+L([0,δ];Hs)\theta\in\mathring{\theta}+L^{\infty}([0,\delta];{H^{s^{\prime}}}) to (1.1) (resp. (1.8)) with θ|t=0=θ̊|t=0+θ0\left.\theta\right|_{t=0}=\left.\mathring{\theta}\right|_{t=0}+\theta_{0}, or

  • any solution θ\theta belonging to θ̊+L([0,δ];Hs)\mathring{\theta}+L^{\infty}([0,\delta];{H^{s^{\prime}}}) satisfy the growth

    supt[0,δ](θθ̊)(t,)Hs>A.\begin{split}\sup_{t\in[0,\delta]}\|{{(\theta-\mathring{\theta})}(t,\cdot)}\|_{H^{s^{\prime}}}>A.\end{split}

Since θ̊(t,)\mathring{\theta}(t,\cdot) can be chosen to have an arbitrarily small HsHsH^{s^{\prime}}\cap H^{s} norm on a sufficiently short time interval in both the dissipative and non-dissipative cases, we immediately obtain the following illposedness statement with θ̊=0\mathring{\theta}=0:

Corollary 1.17.

Assume that (1.1) (resp. (1.8)) is linearly HsH^{s}-HsH^{s^{\prime}} unstable by degenerate dispersion with s>3+32β0s^{\prime}>3+{\frac{3}{2}}\beta_{0}. Then (1.1) (resp. (1.8)) is nonlinearly HsHsH^{s}\to H^{s^{\prime}} ill-posed in the following sense: For any ϵ,δ,A>0\epsilon,\delta,A>0, there exists initial data θ0Cc(Ω)\theta_{0}\in C^{\infty}_{c}(\Omega) with θ0Hs<ϵ{\|{\theta_{0}}\|_{H^{s}}}<\epsilon such that either

  • there exists no solution θL([0,δ];Hs)\theta\in L^{\infty}([0,\delta];{H^{s^{\prime}}}) to (1.1) (resp. (1.8)) with θ|t=0=θ0\left.\theta\right|_{t=0}=\theta_{0}, or

  • any solution θ\theta belonging to L([0,δ];Hs)L^{\infty}([0,\delta];{H^{s^{\prime}}}) satisfy the growth

    supt[0,δ]θ(t,)Hs>A.\begin{split}\sup_{t\in[0,\delta]}\|{{\theta}(t,\cdot)}\|_{H^{s^{\prime}}}>A.\end{split}

While we shall refrain from giving details, we note that the assumption s>3+32β0s^{\prime}>3+{\frac{3}{2}}\beta_{0} can be lowered (up to s>2s^{\prime}>2) when Γ\Gamma becomes less singular. By Corollary 1.11, it follows that any (1.1) satisfying our assumptions for γ\gamma is nonlinearly HsHsH^{s}\to H^{s^{\prime}} ill-posed for s>3+32β0s^{\prime}>3+{\frac{3}{2}}\beta_{0}. Moreover, Corollaries 1.13 and 1.16 provide ranges of (s,s)(s,s^{\prime}) for which (1.1) and (1.8), respectively, are nonlinearly HsHsH^{s}\to H^{s^{\prime}} ill-posed for some model cases of γ\gamma and κυ\kappa\upsilon.

When Ω=𝕋×\Omega=\mathbb{T}\times\mathbb{R} and s=ss=s^{\prime} sufficiently large, we may furthermore exhibit an Hs(Ω)H^{s}(\Omega) initial data corresponding to which no LHsL^{\infty}H^{s} solution to (1.1) exists on any time interval.

Theorem D (Nonexistence).

Let Ω=𝕋×\Omega=\mathbb{T}\times\mathbb{R}. For any s>32β0+3s>\frac{3}{2}\beta_{0}+3 and ϵ>0\epsilon>0, there exists θ0Hs(Ω)\theta_{0}\in H^{s}(\Omega) satisfying θ0Hs<ϵ\|{\theta_{0}}\|_{H^{s}}<\epsilon such that for any δ>0\delta>0, there is no solution to (1.1) belonging to L([0,δ];Hs(Ω))L^{\infty}([0,\delta];H^{s}(\Omega)) with initial data θ0\theta_{0}.

Result form [10]: Nonlinear well(!)posedness of the logarithmically singular case with loss of regularity. Through Corollary 1.13 and Corollary 1.17, we concluded nonlinear HsH^{s}-HsH^{s^{\prime}} illposedness of (1.1) for ss^{\prime} strictly smaller than ss when, say, γ(ξ)=ξβ\gamma(\xi)=\langle{\xi}\rangle^{\beta} for β>0\beta>0. However, this conclusion did not apply to multipliers with slower growth, e.g., γ(ξ)=log(10+|ξ|)\gamma(\xi)=\log(10+|{\xi}|). This difference is no shortcoming of our approach. In the paper [10], we actually obtain local well(!)posedness of (1.1) with γ(ξ)=log(10+|ξ|)\gamma(\xi)=\log(10+|{\xi}|) in Sobolev spaces with exponents that decrease in time. As a result, this system is not nonlinearly HsH^{s}-HsH^{s^{\prime}} ill-posed for any s<ss^{\prime}<s; in the opposite direction, Corollaries 1.13 and 1.17 demonstrate that a decrease of the Sobolev exponent in time is inevitable.

Theorem E (Wellposedness in the logarithmically singular system [10]).

Consider the logarithmically singular SQG equation, possibly with dissipation:

tθ+uθ+κΥ(θ)=0,u=log(10+Λ)θ.\begin{split}\partial_{t}\theta+u\cdot\nabla\theta+\kappa\Upsilon(\theta)=0,\\ u=\nabla^{\perp}\log(10+\Lambda)\theta.\end{split} (1.28)

In the inviscid case (κ=0)(\kappa=0), for any s0>4s_{0}>4 and θ0Hs0\theta_{0}\in H^{s_{0}}, there exists some T=T(s0,θ0Hs0)>0T=T(s_{0},\|{\theta_{0}}\|_{H^{s_{0}}})>0 such that there is a solution θC([0,T];H4)\theta\in C([0,T];H^{4}) to (1.28) with initial data θ0\theta_{0} satisfying

θ(t,)Hs(t)Cθ0Hs0\begin{split}\|{\theta(t,\cdot)}\|_{H^{s(t)}}\leq C\|{\theta_{0}}\|_{H^{s_{0}}}\end{split}

for some continuous function s(t)>4s(t)>4 of tt with s(0)=s0s(0)=s_{0} in t[0,T]t\in[0,T]. The solution is unique in the class C([0,T];H4)C([0,T];H^{4}).

Furthermore, the dissipative system (κ>0)(\kappa>0) is locally well-posed in C([0,T];Hs)C([0,T];H^{s}) for any s>4s>4, as long as there exists some Ξ0>0\Xi_{0}>0 such that

υ(|ξ|)log(10+|ξ|)Csκθ0Hs,|ξ|>Ξ0\begin{split}\frac{\upsilon(|\xi|)}{\log(10+|\xi|)}\geq\frac{C_{s}}{\kappa}\|{\theta_{0}}\|_{H^{s}},\quad|\xi|>\Xi_{0}\end{split}

for some Cs>0C_{s}>0 depending only on ss.

Remark 1.18.

A similar wellposedness result can be proved for Γ=logβ(10+Λ)\Gamma=\log^{\beta}(10+\Lambda) with any β1\beta\leq 1.

Remark 1.19.

The nonexistence theorem (Theorem D), when combined with Theorem E, shows that for logarithmic Γ\Gamma, the local solution with θ0Hs\theta_{0}\in H^{s} for some s>4s>4 in general instantaneously escapes HsH^{s} for t>0t>0.

1.6 Explicitly solvable toy model and discussion of difficulties

Let us present a toy model (1.30) which clearly demonstrates degeneration of linear solutions. To arrive at the toy model, we may start from the nonlinear equation

tθ+Γ(θ)θ=0,\begin{split}\partial_{t}\theta+\nabla^{\perp}\Gamma(\theta)\cdot\nabla\theta=0,\end{split}

and consider the linearization around the steady state θ̊(x1,x2)=x222\mathring{\theta}(x_{1},x_{2})=-\frac{x_{2}^{2}}{2}, under the formal assumption that Γ(x22)0\nabla^{\perp}\Gamma(x_{2}^{2})\equiv 0. The resulting equation for the perturbation, which is again denoted by θ\theta, is simply

tθ+x2x1Γ(θ)=0.\begin{split}\partial_{t}\theta+x_{2}\partial_{x_{1}}\Gamma(\theta)=0.\end{split} (1.29)

We may separate x1x_{1}-dependence under the ansatz

θ(λ0)(t,x1,x2)=eiλ0x1φ(t,x2)\begin{split}\theta^{(\lambda_{0})}(t,x_{1},x_{2})=e^{i\lambda_{0}x_{1}}\varphi(t,x_{2})\end{split}

for some λ0\lambda_{0}. Assume for simplicity that the multiplier γ\gamma for Γ\Gamma is radial. Denoting the dual variable of x2x_{2} by ξ\xi and taking the Fourier transform, we have with Λ(ξ):=λ02+ξ2\Lambda(\xi):=\sqrt{\lambda_{0}^{2}+\xi^{2}} that

λ01tφ^(ξ)+γ(Λ)ξφ^(ξ)=(ξγ)(Λ)ξΛφ^(ξ).\begin{split}\lambda_{0}^{-1}\partial_{t}\widehat{\varphi}(\xi)+\gamma(\Lambda)\partial_{\xi}\widehat{\varphi}(\xi)=-(\partial_{\xi}\gamma)(\Lambda)\frac{\xi}{\Lambda}\widehat{\varphi}(\xi).\end{split} (1.30)

This is simply a transport equation in ξ\xi, which can be explicitly solved along the characteristics: we may define the trajectories ξ(t;ξ0)\xi(t;\xi_{0}) by

{ξ˙(t;ξ0)=λ0γ(Λ(ξ(t;ξ0))),ξ(0;ξ0)=ξ0.\left\{\begin{aligned} \dot{\xi}(t;\xi_{0})&=\lambda_{0}\gamma(\Lambda(\xi(t;\xi_{0}))),\\ \xi(0;\xi_{0})&=\xi_{0}.\end{aligned}\right. (1.31)

The solution can be written by

φ^(t,ξ(t;ξ0))=γ(Λ(ξ0))γ(Λ(ξ(t;ξ0)))φ^0(ξ0).\begin{split}\widehat{\varphi}(t,\xi(t;\xi_{0}))=\frac{\gamma(\Lambda(\xi_{0}))}{\gamma(\Lambda(\xi(t;\xi_{0})))}\widehat{\varphi}_{0}(\xi_{0}).\end{split} (1.32)

The Jacobian of the flow map ξ0ξ(t;ξ0)\xi_{0}\mapsto\xi(t;\xi_{0}) is given simply by

γ(Λ(ξ(t;ξ0)))γ(Λ0).\begin{split}\frac{\gamma(\Lambda(\xi(t;\xi_{0})))}{\gamma(\Lambda_{0})}.\end{split}

In the following, we shall take initial data φ^0\widehat{\varphi}_{0} which is sharply concentrated near ξ0λ0\xi_{0}\simeq\lambda_{0} and introduce simplifying notation Λ(t):=Λ(ξ(t;λ0))\Lambda(t):=\Lambda(\xi(t;\lambda_{0})), Λ0:=2λ0\Lambda_{0}:=\sqrt{2}\lambda_{0}. Then, we have that

φ^(t)Lξ2(γ(Λ(t))γ(Λ0))12φ^0Lξ2.\begin{split}\|{\widehat{\varphi}(t)}\|_{L^{2}_{\xi}}\simeq\left(\frac{\gamma(\Lambda(t))}{\gamma(\Lambda_{0})}\right)^{-\frac{1}{2}}\|{\widehat{\varphi}_{0}}\|_{L^{2}_{\xi}}.\end{split}

This is consistent with propagation of γ12φ^\gamma^{\frac{1}{2}}\widehat{\varphi} in L2L^{2}. On the other hand,

|ξ|sφ^(t)Lξ2|ξ(t;λ0)|s|λ0|s(γ(Λ(t))γ(Λ0))12|ξ|sφ^0Lξ2.\begin{split}\|{|\xi|^{s}\widehat{\varphi}(t)}\|_{L^{2}_{\xi}}\simeq\frac{|\xi(t;\lambda_{0})|^{s}}{|\lambda_{0}|^{s}}\left(\frac{\gamma(\Lambda(t))}{\gamma(\Lambda_{0})}\right)^{-\frac{1}{2}}\|{|\xi|^{s}\widehat{\varphi}_{0}}\|_{L^{2}_{\xi}}.\end{split}

Hence, for (1.29) to be illposed in HsH^{s}, it suffices to have for |λ0|1|\lambda_{0}|\gg 1 and T=T(λ0)1T=T(\lambda_{0})\ll 1 that

|ξ(T;λ0)|s|λ0|s(γ(Λ(T;λ0))γ(Λ0(λ0)))121\begin{split}\frac{|\xi(T;\lambda_{0})|^{s}}{|\lambda_{0}|^{s}}\left(\frac{\gamma(\Lambda(T;\lambda_{0}))}{\gamma(\Lambda_{0}(\lambda_{0}))}\right)^{-\frac{1}{2}}\gg 1\end{split} (1.33)

or simply

|ξ(T)|2sγ(ξ(T))|λ0|2sγ(λ0).\begin{split}\frac{|\xi(T)|^{2s}}{\gamma(\xi(T))}\gg\frac{|\lambda_{0}|^{2s}}{\gamma(\lambda_{0})}.\end{split}

Assuming that γ\gamma is increasing, we have

ξ(t;λ0)λ0+tλ0γ(λ0).\begin{split}\xi(t;\lambda_{0})\geq\lambda_{0}+t\lambda_{0}\gamma(\lambda_{0}).\end{split}

(This is expected to be sharp for small timescales.) Therefore, for some T=T(λ0)T=T(\lambda_{0}) satisfying 1γ(λ0)T1\frac{1}{\gamma(\lambda_{0})}\ll T\ll 1, we have ξ(T)λ0\xi(T)\gg\lambda_{0}, and this will guarantee (1.33) for s>0s>0 large.

Given some concrete symbol γ\gamma (e.g. γ(Λ)=Λβ\gamma(\Lambda)=\langle{\Lambda}\rangle^{\beta} for some β>0\beta>0), one can see the range of s,ss,s^{\prime} where the toy model (1.29) is HsH^{s}HsH^{s^{\prime}} unstable in the sense of Definition 1.12, using the above formula for the solution in the Fourier variable. We leave the details of this computation for the interested reader. Let us demonstrate that in the logarithmic case (where γ(ξ)log(ξ)\gamma(\xi)\lesssim\log(\xi) for large |ξ||\xi|), s=ss=s^{\prime} is forced. Indeed, for s(T)=sMTs(T)=s-MT with some M>0M>0 independent of λ0\lambda_{0}, we have

|ξ(T)|2s(T)γ(ξ(T))|λ0|2sγ(λ0),\begin{split}\frac{|\xi(T)|^{2s(T)}}{\gamma(\xi(T))}\lesssim\frac{|\lambda_{0}|^{2s}}{\gamma(\lambda_{0})},\end{split}

which shows a losing estimate in the scale of time-dependent Sobolev spaces Hs(T)H^{s(T)}. It suggests that there could be a similar estimate in the nonlinear case; this is precisely the content of [10].

One may use the above model equation to understand the dissipative case as well. Then, (1.29) and (1.30) (after separating x1x_{1}-dependence) are simply replaced with

tθ+x2x1Γ(θ)=κΥ(θ)\begin{split}\partial_{t}\theta+x_{2}\partial_{x_{1}}\Gamma(\theta)=-\kappa\Upsilon(\theta)\end{split} (1.34)

and

λ01tφ^(ξ)+γ(Λ)ξφ^(ξ)=(ξγ)(Λ)ξΛφ^(ξ)κυ(Λ)φ^(ξ).\begin{split}\lambda_{0}^{-1}\partial_{t}\widehat{\varphi}(\xi)+\gamma(\Lambda)\partial_{\xi}\widehat{\varphi}(\xi)=-(\partial_{\xi}\gamma)(\Lambda)\frac{\xi}{\Lambda}\widehat{\varphi}(\xi)-\kappa\upsilon(\Lambda)\widehat{\varphi}(\xi).\end{split} (1.35)

(We are assuming that the symbol υ\upsilon is radial.) Again, this equation is solvable along characteristics. Using the solution, one may see that when υγ\upsilon\gg\gamma, there is well-posedness in HsH^{s}-spaces, while υγ\upsilon\ll\gamma still gives HsH^{s} illposedness.

Discussion of difficulties. We are now in a good position to explain the main difficulties in establishing Sobolev illposedness for the actual linear homogeneous equations associated with (2.1) and (2.2), as well as the nonlinear equations (1.1) and (1.8). Comparing the inviscid linear case (2.1) with (1.29), there are two differences: (1) first, the principal coefficient is not exactly linear, and (2) second, a lower order term is present in (2.1). With these two differences combined, it becomes a challenging problem to just construct an approximate solution to (2.1) which exhibits the same illposedness behavior with an explicit solution to (1.29). Meanwhile, once a good approximate construction has been constructed, it can be upgraded to illlposedness results for (2.1), and even the nonlinear equation (1.1), using the robust testing (or duality) method introduced in [20].

Henceforth, we focus on approximate solution construction. Regarding difference (1), it has been known that having an exactly linear principal coefficient significantly simplifies the analysis (Craig–Goodman [14] is a good example). We have seen in the above that in such a case, taking the Fourier transform results in a transport term. Moreover, one can perform WKB analysis with a linear phase. In view of this, when the principal term is given by a differential operator with a linearly degenerate coefficient, it is natural to first apply a coordinate transform which makes the coefficient exactly linear near the degeneracy; this was the approach in our previous work [20]. Such a coordinate change is not available in the case of pseudo-differential principal term. Next, regarding difference (2), note that the lower order term in (2.1) is again given by a pseudo-differential operator, and its (generalized) order could become very close to that for the principal operator when Γ\Gamma is only slightly singular (e.g. when γ(ξ)=log(10+|ξ|),log(10+log(10+|ξ|))\gamma(\xi)=\log(10+|\xi|),\log(10+\log(10+|\xi|))). In such cases, it is basically impossible to distinguish the lower order term from the principal term, and in the proof we indeed incorporate a part of the lower order term into the principal part.

In the dissipative linear case (2.2), the main idea is to use the same approximate solution to (2.1) and treat the dissipative term as an error. Nevertheless, there is yet another serious difficulty: in general, there is no nontrivial steady solutions to (1.14) and we have to work with time-dependent shear flows. In general, the strength of the degeneracy of the principal coefficient Γθ̊\Gamma\nabla^{\perp}\mathring{\theta} change with time555In principle, the location of the degeneracy would move in time as well; in the current work, we prohibit such behavior by imposing even symmetry both in the shear profile θ̊\mathring{\theta} and the dissipative operator Υ\Upsilon., whose effect cannot be regarded as a perturbation from the initial data Γθ̊0\Gamma\nabla^{\perp}\mathring{\theta}_{0}. Again, with respect to this difficulty, the most problematic case is when Γ\Gamma is slightly singular, because then the frequency growth is slower (hence most sensitive to time dependence of the background shear).

To overcome these difficulties, we develop a fairly general framework for directly constructing degenerating wave packets for a linear pseudo-differential equation of the form tϕ+ip(t,x,D)ϕ=0\partial_{t}\phi+ip(t,x,D)\phi=0, where p(t,x,D)p(t,x,D) takes into account not only the principal term but also key lower order terms in (2.1), and may possibly be time-dependent. While the formal derivation of the ansatz for the approximate solution is straight forward (see Section 3), the difficulty is to rigorously control the ansatz and the error in sufficiently long time scales within which significant degeneration occurs. Among others, two key ideas in this work that allow us to resolve this difficulty are: observations concerning the Hamilton–Jacobi equation for t+ip\partial_{t}+ip and the associated transport equations that allow for controlling the ansatz in long enough time scales (see Section 4) and sharp estimates for oscillatory integrals appearing in the symbol for the error term (see Section 5). We refer the reader to Section 2.2 for a more detailed discussion of the key ideas.

1.7 Organization of the paper

The rest of the paper is divided into six sections, which we briefly describe below. The proofs of the main results in each section is largely independent of those from other sections.

  • Section 2 contain preliminary computations on the linearized operators and a preview of the proofs of main results.

  • Section 3 begins with algebraic preliminaries regarding pseudo-differential calculus. The equations satisfied by the phase and amplitude functions are fixed in this section. Then, we provide expansion formulas for the pseudo-differential operators appearing after the conjugation by the phase function. In particular, an explicit representation formula for the remainder operator is derived.

  • Section 4 deals with the equations for the phase and amplitude functions chosen in Section 3. In this section, we fix the choice of initial data for the phase and amplitude functions and derive sharp high order estimates for them.

  • In Section 5, we obtain operator bounds for symbols appeared in Section 3.

  • In Section 6, key estimates for the degenerating wave packet solutions are obtained, by applying sharp estimates from Section 4 to operator bounds from Section 5.

  • All the main theorems are proved in Section 7, combining all the ingredients.

Acknowledgments

D. Chae was supported by NRF grant No. 2021R1A2C1003234. I.-J. Jeong was supported by the Samsung Science and Technology Foundation under Project Number SSTF-BA2002-04. S.-J. Oh was supported in part by the Samsung Science and Technology Foundation under Project Number SSTF-BA1702-02, a Sloan Research Fellowship and a National Science Foundation CAREER Grant under NSF-DMS-1945615.

2 Preliminaries and ideas of the proof

In Section 2.1, we derive the linearized operators and discuss its energy structure. We then proceed to define the notion of an L2L^{2} solution, and derive the generalized energy identity. An outline of the proof together with some key ideas are given in Section 2.2.

2.1 Energy structure of the linearized operator

2.1.1 Conjugated linearized operators

Note that θ=f(x2)\theta=f(x_{2}) is a (formal) solution for any (regular, decaying) profile ff. Indeed,

uθ=x2ψ(x2)x1f(x2)x1ψ(x2)x2f(x2)=0.u\cdot\nabla\theta=\partial_{x_{2}}\psi(x_{2})\partial_{x_{1}}f(x_{2})-\partial_{x_{1}}\psi(x_{2})\partial_{x_{2}}f(x_{2})=0.

The direct linearization around θ̊\mathring{\theta} is given as follows:

Lθ̊ϕ=tϕθ̊Γϕ+Γθ̊ϕ=0.\displaystyle L_{\mathring{\theta}}\phi=\partial_{t}\phi-\nabla^{\perp}\mathring{\theta}\cdot\nabla\Gamma\phi+\nabla^{\perp}\Gamma\mathring{\theta}\cdot\nabla\phi=0.

Indeed, writing θ=θ̊+ϕ\theta=\mathring{\theta}+\phi,

u\displaystyle u =Γ(θ̊+ϕ),\displaystyle=\nabla^{\perp}\Gamma(\mathring{\theta}+\phi),
t(f+ϕ)+u(θ̊+ϕ)\displaystyle\partial_{t}(f+\phi)+u\cdot\nabla(\mathring{\theta}+\phi) =tϕ+Γϕθ̊+Γθ̊ϕ+Γϕϕ\displaystyle=\partial_{t}\phi+\nabla^{\perp}\Gamma\phi\cdot\nabla\mathring{\theta}+\nabla^{\perp}\Gamma\mathring{\theta}\cdot\nabla\phi+\nabla^{\perp}\Gamma\phi\cdot\nabla\phi
=tϕΓϕθ̊+Γθ̊ϕ+Γϕϕ.\displaystyle=\partial_{t}\phi-\nabla\Gamma\phi\cdot\nabla^{\perp}\mathring{\theta}+\nabla^{\perp}\Gamma\mathring{\theta}\cdot\nabla\phi+\nabla^{\perp}\Gamma\phi\cdot\nabla\phi.

Already we may observe that

~θ̊ϕ=tϕ+P~θ̊ϕ,\widetilde{\mathcal{L}}_{\mathring{\theta}}\phi=\partial_{t}\phi+\widetilde{P}_{\mathring{\theta}}\phi,

where the principal symbol of P~θ̊\widetilde{P}_{\mathring{\theta}} is

iθ̊(x1,x2)ξΓ(ξ).-i\nabla^{\perp}\mathring{\theta}(x_{1},x_{2})\cdot\xi\Gamma(\xi).

where ξ=(ξ1,ξ2)\xi=(\xi_{1},\xi_{2}), and we assumed that γ(ξ)\gamma(\xi)\to\infty as |ξ||{\xi}|\to\infty. Note that pθ̊p_{\mathring{\theta}} is purely imaginary.

However, as we will see soon, Lθ̊L_{\mathring{\theta}} does not have a good energy structure; there is a problem with the sub-principal terms. For this reason, we will have to conjugate Lθ̊L_{\mathring{\theta}} by Γ12\Gamma^{\frac{1}{2}}.

As a motivation, let us first discuss the energy structure of Lθ̊L_{\mathring{\theta}}. Our computation will be formal.

We first write

Lθ̊ϕϕdxdy\displaystyle\int L_{\mathring{\theta}}\phi\phi\,\mathrm{d}x\mathrm{d}y =12ddtϕ2dxdy(θ̊Γϕ)ϕdxdy\displaystyle=\frac{1}{2}\frac{\mathrm{d}}{\mathrm{d}t}\int\phi^{2}\,\mathrm{d}x\mathrm{d}y-\int\left(\nabla^{\perp}\mathring{\theta}\cdot\nabla\Gamma\phi\right)\phi\,\mathrm{d}x\mathrm{d}y
+12Γθ̊ϕ2dxdy\displaystyle\mathrel{\phantom{=}}+\frac{1}{2}\int\nabla^{\perp}\Gamma\mathring{\theta}\nabla\phi^{2}\,\mathrm{d}x\mathrm{d}y

Integrating \nabla by parts, the last term vanishes since =0\nabla\cdot\nabla^{\perp}=0. For the second term, we have the following chain of identities:

(θ̊Γϕ)ϕdxdy\displaystyle-\int\left(\nabla^{\perp}\mathring{\theta}\cdot\nabla\Gamma\phi\right)\phi\,\mathrm{d}x\mathrm{d}y =12(θ̊Γϕ)ϕdxdy+12ϕΓ(θ̊ϕ)dxdy\displaystyle=-\frac{1}{2}\int\left(\nabla^{\perp}\mathring{\theta}\cdot\nabla\Gamma\phi\right)\phi\,\mathrm{d}x\mathrm{d}y+\frac{1}{2}\int\phi\Gamma\left(\nabla^{\perp}\mathring{\theta}\cdot\nabla\phi\right)\,\mathrm{d}x\mathrm{d}y
=12ϕ[θ̊,Γ]ϕdxdy.\displaystyle=-\frac{1}{2}\int\phi[\nabla^{\perp}\mathring{\theta}\cdot\nabla,\Gamma]\phi\,\mathrm{d}x\mathrm{d}y.

However, [θ̊,Γ][\nabla^{\perp}\mathring{\theta}\cdot\nabla,\Gamma] is symmetric and is not bounded in L2L^{2}, which is problematic.

We wish to remove [θ̊,Γ][\nabla^{\perp}\mathring{\theta}\cdot\nabla,\Gamma], we need to conjugate the equation. Motivated by the desire to make the principal term exactly anti-symmetric (alternatively, via a formal, symbolic-calculus computation), we work with the variable φ\varphi given by

ϕ=Γ12φ.\phi=\Gamma^{-\frac{1}{2}}\varphi.

Accordingly, we define

θ̊φ=Γ12Lθ̊(Γ12φ).\mathcal{L}_{\mathring{\theta}}\varphi=\Gamma^{\frac{1}{2}}L_{\mathring{\theta}}(\Gamma^{-\frac{1}{2}}\varphi).

We compute

θ̊φ=tφΓ12θ̊Γ12φ+Γ12Γθ̊Γ12φ.\displaystyle\mathcal{L}_{\mathring{\theta}}\varphi=\partial_{t}\varphi-\Gamma^{\frac{1}{2}}\nabla^{\perp}\mathring{\theta}\cdot\nabla\Gamma^{\frac{1}{2}}\varphi+\Gamma^{\frac{1}{2}}\nabla^{\perp}\Gamma\mathring{\theta}\cdot\nabla\Gamma^{-\frac{1}{2}}\varphi. (2.1)

We now investigate the energy structure of θ̊\mathcal{L}_{\mathring{\theta}}; as before, we present only a formal computation. We begin with

θ̊φφdxdy\displaystyle\int\mathcal{L}_{\mathring{\theta}}\varphi\varphi\,\mathrm{d}x\mathrm{d}y =12ddtφ2dxdy(Γ12θ̊Γ12φ)φdxdy\displaystyle=\frac{1}{2}\frac{\mathrm{d}}{\mathrm{d}t}\int\varphi^{2}\,\mathrm{d}x\mathrm{d}y-\int\left(\Gamma^{\frac{1}{2}}\nabla^{\perp}\mathring{\theta}\cdot\nabla\Gamma^{\frac{1}{2}}\varphi\right)\varphi\,\mathrm{d}x\mathrm{d}y
+(Γ12Γθ̊Γ12φ)φdxdy.\displaystyle\mathrel{\phantom{=}}+\int\left(\Gamma^{\frac{1}{2}}\nabla^{\perp}\Gamma\mathring{\theta}\cdot\nabla\Gamma^{-\frac{1}{2}}\varphi\right)\varphi\,\mathrm{d}x\mathrm{d}y.

By the symmetry of Γ12\Gamma^{\frac{1}{2}} and anti-symmetry of θ̊\nabla^{\perp}\mathring{\theta}\cdot\nabla, the second term vanishes. However, the last term does not vanish (cf. the computation for Lθ̊ϕL_{\mathring{\theta}}\phi). For this term, we compute

(Γ12Γθ̊Γ12φ)φdxdy\displaystyle\int\left(\Gamma^{\frac{1}{2}}\nabla^{\perp}\Gamma\mathring{\theta}\cdot\nabla\Gamma^{-\frac{1}{2}}\varphi\right)\varphi\,\mathrm{d}x\mathrm{d}y
=12(Γ12Γθ̊Γ12φ)φdxdy12φΓ12Γθ̊Γ12φdxdy\displaystyle=\frac{1}{2}\int\left(\Gamma^{\frac{1}{2}}\nabla^{\perp}\Gamma\mathring{\theta}\cdot\nabla\Gamma^{-\frac{1}{2}}\varphi\right)\varphi\,\mathrm{d}x\mathrm{d}y-\frac{1}{2}\int\varphi\Gamma^{-\frac{1}{2}}\nabla^{\perp}\Gamma\mathring{\theta}\cdot\nabla\Gamma^{\frac{1}{2}}\varphi\,\mathrm{d}x\mathrm{d}y
=12(Γθ̊φ)φdxdy+12([Γ12,Γθ̊]Γ12φ)φdxdy\displaystyle=\frac{1}{2}\int\left(\nabla^{\perp}\Gamma\mathring{\theta}\cdot\nabla\varphi\right)\varphi\,\mathrm{d}x\mathrm{d}y+\frac{1}{2}\int\left([\Gamma^{\frac{1}{2}},\nabla^{\perp}\Gamma\mathring{\theta}\cdot\nabla]\Gamma^{-\frac{1}{2}}\varphi\right)\varphi\,\mathrm{d}x\mathrm{d}y
12φΓθ̊φdxdy12φΓ12[Γθ̊,Γ12]φdxdy\displaystyle\mathrel{\phantom{=}}-\frac{1}{2}\int\varphi\nabla^{\perp}\Gamma\mathring{\theta}\cdot\nabla\varphi\,\mathrm{d}x\mathrm{d}y-\frac{1}{2}\int\varphi\Gamma^{-\frac{1}{2}}[\nabla^{\perp}\Gamma\mathring{\theta}\cdot\nabla,\Gamma^{\frac{1}{2}}]\varphi\,\mathrm{d}x\mathrm{d}y
=12φ([Γ12,Γθ̊]Γ12+Γ12[Γ12,Γθ̊])φdxdy.\displaystyle=\frac{1}{2}\int\varphi\left([\Gamma^{\frac{1}{2}},\nabla^{\perp}\Gamma\mathring{\theta}\cdot\nabla]\Gamma^{-\frac{1}{2}}+\Gamma^{-\frac{1}{2}}[\Gamma^{\frac{1}{2}},\nabla^{\perp}\Gamma\mathring{\theta}\cdot\nabla]\right)\varphi\,\mathrm{d}x\mathrm{d}y.

Since Γ\Gamma is a classical symbol (more precisely, see Assumption 1), the operator

[Γ12,Γθ̊]Γ12+Γ12[Γ12,Γθ̊][\Gamma^{\frac{1}{2}},\nabla^{\perp}\Gamma\mathring{\theta}\cdot\nabla]\Gamma^{-\frac{1}{2}}+\Gamma^{-\frac{1}{2}}[\Gamma^{\frac{1}{2}},\nabla^{\perp}\Gamma\mathring{\theta}\cdot\nabla]

is bounded on L2L^{2}; hence this term is acceptable.

Finally, we state the linearized operator for (1.8):

θ̊(κ)φ=tφΓ12θ̊Γ12φ+Γ12Γθ̊Γ12φ+κΥφ.\mathcal{L}^{(\kappa)}_{\mathring{\theta}}\varphi=\partial_{t}\varphi-\Gamma^{\frac{1}{2}}\nabla^{\perp}\mathring{\theta}\cdot\nabla\Gamma^{\frac{1}{2}}\varphi+\Gamma^{\frac{1}{2}}\nabla^{\perp}\Gamma\mathring{\theta}\cdot\nabla\Gamma^{-\frac{1}{2}}\varphi+\kappa\Upsilon\varphi. (2.2)

Note that (κ)θ̊Γ12ϕ=Γ12L(κ)θ̊ϕ\mathcal{L}^{(\kappa)}_{\mathring{\theta}}\Gamma^{\frac{1}{2}}\phi=\Gamma^{\frac{1}{2}}L^{(\kappa)}_{\mathring{\theta}}\phi.

2.1.2 Notion of an L2L^{2} solution

In the following, for simplicity, we consider (2.1) as a special case of (2.2) obtained by taking κ=0\kappa=0. Recall from above that a sufficiently smooth and decaying solution to (2.2) satisfies

12ddtφL22+κΥ12φL22+12φ,([Γ12,Γθ̊]Γ12+Γ12[Γ12,Γθ̊])φ=0.\begin{split}\frac{1}{2}\frac{\mathrm{d}}{\mathrm{d}t}\|{\varphi}\|_{L^{2}}^{2}+\kappa\|{\Upsilon^{\frac{1}{2}}\varphi}\|_{L^{2}}^{2}+\frac{1}{2}\langle{\varphi,\left([\Gamma^{\frac{1}{2}},\nabla^{\perp}\Gamma\mathring{\theta}\cdot\nabla]\Gamma^{-\frac{1}{2}}+\Gamma^{-\frac{1}{2}}[\Gamma^{\frac{1}{2}},\nabla^{\perp}\Gamma\mathring{\theta}\cdot\nabla]\right)\varphi}\rangle=0.\end{split}

Since

([Γ12,Γθ̊]Γ12+Γ12[Γ12,Γθ̊])φL2Γ,θ̊φL2,\begin{split}\|{\left([\Gamma^{\frac{1}{2}},\nabla^{\perp}\Gamma\mathring{\theta}\cdot\nabla]\Gamma^{-\frac{1}{2}}+\Gamma^{-\frac{1}{2}}[\Gamma^{\frac{1}{2}},\nabla^{\perp}\Gamma\mathring{\theta}\cdot\nabla]\right)\varphi}\|_{L^{2}}\lesssim_{\Gamma,\mathring{\theta}}\|{\varphi}\|_{L^{2}},\end{split}

we obtain from Grönwall’s inequality that

φ(t)L2φ0L2exp(C(Γ,θ̊)t).\begin{split}\|{\varphi(t)}\|_{L^{2}}\leq\|{\varphi_{0}}\|_{L^{2}}\exp\left(C(\Gamma,\mathring{\theta})t\right).\end{split} (2.3)

This motivates the following definition of an L2L^{2}-solution:

Definition 2.1.

Given some interval I=[0,τ]I=[0,\tau], we say that φ\varphi is an L2L^{2}-solution of (κ)θ̊φ=0\mathcal{L}^{(\kappa)}_{\mathring{\theta}}\varphi=0 if

  • φCw(I;L2)L2t(I;κ12Υ12L2)\varphi\in C_{w}(I;L^{2})\cap L^{2}_{t}(I;\kappa^{-\frac{1}{2}}\Upsilon^{-\frac{1}{2}}L^{2});

  • φ\varphi satisfies (κ)θ̊φ=0\mathcal{L}^{(\kappa)}_{\mathring{\theta}}\varphi=0 in the sense of distributions;

  • φ\varphi satisfies (2.3).

We say that ϕ\phi is an Γ12L2\Gamma^{-\frac{1}{2}}L^{2}-solution of L(κ)θ̊ϕ=0L^{(\kappa)}_{\mathring{\theta}}\phi=0 if φ=Γ12ϕ\varphi=\Gamma^{\frac{1}{2}}\phi is an L2L^{2}-solution of (κ)θ̊φ=0\mathcal{L}^{(\kappa)}_{\mathring{\theta}}\varphi=0. Moreover, we simply drop the requirement φL2t(I;κ12Υ12L2)\varphi\in L^{2}_{t}(I;\kappa^{-\frac{1}{2}}\Upsilon^{-\frac{1}{2}}L^{2}) in the inviscid case κ=0\kappa=0.

Here, Cw(I;L2)C_{w}(I;L^{2}) is a subspace of Lt(I;L2)L^{\infty}_{t}(I;L^{2}) containing functions weakly continuous in time with values in L2L^{2}. The space L2t(I;κ12Υ12L2)L^{2}_{t}(I;\kappa^{-\frac{1}{2}}\Upsilon^{-\frac{1}{2}}L^{2}) is defined by the norm Iκ12Υ12φ(t)L22dt\int_{I}\|{\kappa^{\frac{1}{2}}\Upsilon^{\frac{1}{2}}\varphi(t)}\|_{L^{2}}^{2}\,\mathrm{d}t.

We have the following existence result:

Proposition 2.2.

Given any φ0L2\varphi_{0}\in L^{2}, there is at least one L2L^{2} solution to (κ)θ̊φ=0\mathcal{L}^{(\kappa)}_{\mathring{\theta}}\varphi=0 satisfying Definition 2.1 for any κ0\kappa\geq 0.

We omit the proof, which is a simple application of the Aubin–Lions lemma; see [20, Appendix A] for details.

2.1.3 Generalized energy identity

We now present the generalized energy identity, which is one of the main tools in the proof of linear and nonlinear illposedness. For the moment assume that φ\varphi and ψ\psi are sufficiently smooth, decaying fast at infinity, and solve (κ)θ̊φ=0\mathcal{L}^{(\kappa)}_{\mathring{\theta}}\varphi=0 with errors ϵφ\boldsymbol{\epsilon}_{\varphi} and ϵψ\boldsymbol{\epsilon}_{\psi}; that is,

(κ)θ̊φ=ϵφ,(κ)θ̊ψ=ϵψ.\begin{split}\mathcal{L}^{(\kappa)}_{\mathring{\theta}}\varphi=\boldsymbol{\epsilon}_{\varphi},\qquad\mathcal{L}^{(\kappa)}_{\mathring{\theta}}\psi=\boldsymbol{\epsilon}_{\psi}.\end{split}

Then, we compute

ddtφ,ψ+2κΥ12φ,Υ12ψ\displaystyle\frac{\mathrm{d}}{\mathrm{d}t}\langle{\varphi,\psi}\rangle+2\kappa\langle{\Upsilon^{\frac{1}{2}}\varphi,\Upsilon^{\frac{1}{2}}\psi}\rangle =ϵφ,ψ+φ,ϵψ+Γ12θ̊Γ12φ,ψ+φ,Γ12θ̊Γ12ψ\displaystyle=\langle{\boldsymbol{\epsilon}_{\varphi},\psi}\rangle+\langle{\varphi,\boldsymbol{\epsilon}_{\psi}}\rangle+\langle{\Gamma^{\frac{1}{2}}\nabla^{\perp}\mathring{\theta}\cdot\nabla\Gamma^{\frac{1}{2}}\varphi,\psi}\rangle+\langle{\varphi,\Gamma^{\frac{1}{2}}\nabla^{\perp}\mathring{\theta}\cdot\nabla\Gamma^{\frac{1}{2}}\psi}\rangle
Γ12Γθ̊Γ12φ,ψφ,Γ12Γθ̊Γ12ψ.\displaystyle\mathrel{\phantom{=}}-\langle{\Gamma^{\frac{1}{2}}\nabla^{\perp}\Gamma\mathring{\theta}\cdot\nabla\Gamma^{-\frac{1}{2}}\varphi,\psi}\rangle-\langle{\varphi,\Gamma^{\frac{1}{2}}\nabla^{\perp}\Gamma\mathring{\theta}\cdot\nabla\Gamma^{-\frac{1}{2}}\psi}\rangle.

Similarly as in the derivation of the energy identity, we have that the third and fourth terms on the right hand side cancel each other, using the anti-symmetry of θ̊\nabla^{\perp}\mathring{\theta}\cdot\nabla. Next, following the proof of the energy identity, the last two terms can be combined as follows:

Γ12Γθ̊Γ12φ,ψ+φ,Γ12Γθ̊Γ12ψ=φ,([Γ12,Γθ̊]Γ12+Γ12[Γ12,Γθ̊])ψ.\begin{split}\langle{\Gamma^{\frac{1}{2}}\nabla^{\perp}\Gamma\mathring{\theta}\cdot\nabla\Gamma^{-\frac{1}{2}}\varphi,\psi}\rangle+\langle{\varphi,\Gamma^{\frac{1}{2}}\nabla^{\perp}\Gamma\mathring{\theta}\cdot\nabla\Gamma^{-\frac{1}{2}}\psi}\rangle=\langle{\varphi,\left([\Gamma^{\frac{1}{2}},\nabla^{\perp}\Gamma\mathring{\theta}\cdot\nabla]\Gamma^{-\frac{1}{2}}+\Gamma^{-\frac{1}{2}}[\Gamma^{\frac{1}{2}},\nabla^{\perp}\Gamma\mathring{\theta}\cdot\nabla]\right)\psi}\rangle.\end{split}

Assuming for simplicity that ϵφ=0\boldsymbol{\epsilon}_{\varphi}=0, we have the following generalized energy identity:

Proposition 2.3.

Let φ\varphi be an L2L^{2} solution to (κ)θ̊φ=0\mathcal{L}^{(\kappa)}_{\mathring{\theta}}\varphi=0 on II in the sense of Definition 2.1, and assume that ψ\psi satisfy (κ)θ̊ψ=ϵψ\mathcal{L}^{(\kappa)}_{\mathring{\theta}}\psi=\boldsymbol{\epsilon}_{\psi} on II with regularity

ψCt(I;L2)L2t(I;κ12Υ12L2)L2t(I;H1),ϵψL1t(I;L2).\begin{split}\psi\in C_{t}(I;L^{2})\cap L^{2}_{t}(I;\kappa^{-\frac{1}{2}}\Upsilon^{-\frac{1}{2}}L^{2})\cap L^{2}_{t}(I;H^{1}),\qquad\boldsymbol{\epsilon}_{\psi}\in L^{1}_{t}(I;L^{2}).\end{split}

Then, we have

ddtφ,ψ+2κΥ12φ,Υ12ψ=φ,ϵψ+φ,([Γ12,Γθ̊]Γ12+Γ12[Γ12,Γθ̊])ψ\begin{split}\frac{\mathrm{d}}{\mathrm{d}t}\langle{\varphi,\psi}\rangle+2\kappa\langle{\Upsilon^{\frac{1}{2}}\varphi,\Upsilon^{\frac{1}{2}}\psi}\rangle=\langle{\varphi,\boldsymbol{\epsilon}_{\psi}}\rangle+\langle{\varphi,\left([\Gamma^{\frac{1}{2}},\nabla^{\perp}\Gamma\mathring{\theta}\cdot\nabla]\Gamma^{-\frac{1}{2}}+\Gamma^{-\frac{1}{2}}[\Gamma^{\frac{1}{2}},\nabla^{\perp}\Gamma\mathring{\theta}\cdot\nabla]\right)\psi}\rangle\end{split} (2.4)

on tIt\in I.

The proof follows from first mollifying φ,ψ\varphi,\psi and repeating the computations above, which gives a generalized energy identity with some error terms arising from the mollification. Then it is not difficult to observe that the mollification errors vanish as the mollification parameter goes to zero, using the assumed regularity of φ,ψ,ϵψ\varphi,\psi,\boldsymbol{\epsilon}_{\psi}. We omit the straightforward details (cf. [20, Proposition 2.3]). Moreover, one may generalize the above proposition to the case when φ\varphi is an L2L^{2} solution with an L1t(I;L2)L^{1}_{t}(I;L^{2}) error, denoted by ϵφ\boldsymbol{\epsilon}_{\varphi}. Then, we again have (2.4) with ϵφ,ψ\langle{\boldsymbol{\epsilon}_{\varphi},\psi}\rangle added to the right hand side.

2.2 Key ideas

Outline of the proof. We give an overall picture of the illposedness proofs. Given (2.1), most of the work goes into construction of degenerating wave packet solutions φ~\widetilde{\varphi} to (2.1). Basically, we would like to construct some approximate solutions which behave similarly with the solution of the toy model (1.32), given a quadratically degenerate shear steady state θ̊\mathring{\theta}. Some key properties that are required for the degenerating wave packets solutions can be summarized as follows:

  • Frequency localization: the initial data is sharply concentrated near some frequency λ01\lambda_{0}\gg 1; φ~0Hsλ0s\|{\widetilde{\varphi}_{0}}\|_{H^{s}}\sim\lambda_{0}^{s} for s0s\geq 0.

  • Error bound: for some interval I=[0,t0]I=[0,{t_{0}}] with t0>0{t_{0}}>0, θ̊[φ~]L1t(I;L2)1\|{\mathcal{L}_{\mathring{\theta}}[\widetilde{\varphi}]}\|_{L^{1}_{t}(I;L^{2})}\ll 1. That is, φ~\widetilde{\varphi} is an approximate solution to (2.1). From the energy identity, it also follows that φ~Lt(I;L2)φ~0L21\|{\widetilde{\varphi}}\|_{L^{\infty}_{t}(I;L^{2})}\lesssim\|{\widetilde{\varphi}_{0}}\|_{L^{2}}\lesssim 1.

  • Decay of negative Sobolev norms (degeneration estimate): we have a decomposition φ~=φ~main+φ~small\widetilde{\varphi}=\widetilde{\varphi}^{main}+\widetilde{\varphi}^{small} such that φ~smallLt(I;L2)1\|{\widetilde{\varphi}^{small}}\|_{L^{\infty}_{t}(I;L^{2})}\ll 1 and λ0sφ~main(τ)Hs1\lambda_{0}^{s}\|{\widetilde{\varphi}^{main}(\tau)}\|_{H^{-s}}\ll 1 for some s>0s>0.666The decomposition φ~=φ~main+φ~small\widetilde{\varphi}=\widetilde{\varphi}^{main}+\widetilde{\varphi}^{small} is necessary since we would like to take advantage of the amplitude function compactly supported in space. Therefore, unlike the toy model solution (which is compactly supported in the frequency side), there is some low-frequency part in φ~\widetilde{\varphi} which does not degenerate. (The large parameter is λ01\lambda_{0}\gg 1.)

Even in the case of dissipative linear equation (κ)θ̊φ=0\mathcal{L}^{(\kappa)}_{\mathring{\theta}}\varphi=0, the above requirements remain unchanged, except that the profile θ̊\mathring{\theta} becomes time-dependent (where tf+κΥf=0\partial_{t}f+\kappa\Upsilon f=0): we shall always incorporate the dissipation term κΥφ~\kappa\Upsilon\widetilde{\varphi} as a part of the error term. Assuming for a moment that we are given φ~\widetilde{\varphi} satisfying the above, the rest of the illposedness proof follows a duality or generalized energy argument that originated from [20] for Hall- and electron-MHD systems, which we now explain. For the moment we consider the linear illposedness result. To begin, simply take the initial data φ0=φ~0\varphi_{0}=\widetilde{\varphi}_{0} and apply the generalized energy identity (2.4) with ψ=φ~\psi=\widetilde{\varphi} and φ\varphi some L2L^{2}-solution to θ̊φ=0\mathcal{L}_{\mathring{\theta}}\varphi=0 on I=[0,t0]I=[0,{t_{0}}] associated with φ0\varphi_{0}. We then obtain

|ddtφ,φ~|φ(t)L2(θ̊[φ~](t)L2+φ~(t)L2).\begin{split}\left|\frac{\mathrm{d}}{\mathrm{d}t}\langle{\varphi,\widetilde{\varphi}}\rangle\right|\lesssim\|{\varphi(t)}\|_{L^{2}}(\|{\mathcal{L}_{\mathring{\theta}}[\widetilde{\varphi}](t)}\|_{L^{2}}+\|{\widetilde{\varphi}(t)}\|_{L^{2}}).\end{split}

Using

φ(t)L2φ0L2,φ~(t)L2φ~0L2,\begin{split}\|{\varphi(t)}\|_{L^{2}}\lesssim\|{\varphi_{0}}\|_{L^{2}},\qquad\|{\widetilde{\varphi}(t)}\|_{L^{2}}\lesssim\|{\widetilde{\varphi}_{0}}\|_{L^{2}},\end{split}

which is valid for 0tt00\leq t\leq{t_{0}} with t0=Oθ̊(1){t_{0}}=O_{\mathring{\theta}}(1), and then integrating in time,

φ,φ~(t)φ0,φ~0Cφ0L2(θ̊[φ~]L1([0,t];L2)+tφ~0L2).\begin{split}\langle{\varphi,\widetilde{\varphi}}\rangle(t)\geq\langle{\varphi_{0},\widetilde{\varphi}_{0}}\rangle-C\|{\varphi_{0}}\|_{L^{2}}(\|{\mathcal{L}_{\mathring{\theta}}[\widetilde{\varphi}]}\|_{L^{1}([0,t];L^{2})}+t\|{\widetilde{\varphi}_{0}}\|_{L^{2}}).\end{split}

Using the error bound, we can deduce for t0>0t_{0}>0 small that φ,φ~(t0)>910φ0,φ~0\langle{\varphi,\widetilde{\varphi}}\rangle(t_{0})>\frac{9}{10}\langle{\varphi_{0},\widetilde{\varphi}_{0}}\rangle (say). Then, combining this estimate with

φ,φ~(t0)φ(t0)L2φ~small(t0)L2+φ(t0)Hsφ~main(t0)Hs\begin{split}\langle{\varphi,\widetilde{\varphi}}\rangle({t_{0}})\leq\|{\varphi({t_{0}})}\|_{L^{2}}\|{\widetilde{\varphi}^{small}({t_{0}})}\|_{L^{2}}+\|{\varphi({t_{0}})}\|_{H^{s}}\|{\widetilde{\varphi}^{main}({t_{0}})}\|_{H^{-s}}\end{split}

and φ(t0)L2φ~small(t0)L2<110φ0,φ~0\|{\varphi({t_{0}})}\|_{L^{2}}\|{\widetilde{\varphi}^{small}({t_{0}})}\|_{L^{2}}<\frac{1}{10}\langle{\varphi_{0},\widetilde{\varphi}_{0}}\rangle (say), we deduce that λ0sφ~(t0)Hs1\lambda_{0}^{-s}\|{\widetilde{\varphi}({t_{0}})}\|_{H^{s}}\gg 1. Here, the initial data φ0\varphi_{0} only needs to satisfy that φ0,φ~0\langle{\varphi_{0},\widetilde{\varphi}_{0}}\rangle is comparable with φ0L2φ~0L2\|{\varphi_{0}}\|_{L^{2}}\|{\widetilde{\varphi}_{0}}\|_{L^{2}}.

To prove the nonlinear illposedness results, we assume towards a contradiction that a sufficiently smooth solution to the nonlinear equation exists, with initial data which is given by a small and smooth perturbation of some quadratically degenerate shear steady state θ̊\mathring{\theta}. Here, a key observation is that then the perturbation, after applying Γ12\Gamma^{\frac{1}{2}}, is a solution of the linear operator with a nonlinear error, and that the nonlinear error in L2L^{2} can be bounded in terms of the L2L^{2}-norm of the perturbation, under the assumption that the perturbation remains sufficiently smooth. This is responsible for the restriction on ss^{\prime} in the nonlinear statements.

In what follows, we explain a few key ideas that are involved in the construction of degenerating wave packets. In [20, Sections 1–2], a detailed introduction is given regarding the construction of degenerating wave packets for the case of Hall- and electron-MHD systems. For this reason, here we emphasize on how the additional difficulties (over the Hall- and electron-MHD cases) are handled.

Degenerating wave packet construction. There is a general recipe for construction of wave packet solutions, sometimes referred to as the WKB ansatz; one prepares the ansatz

φ~=Re(exp(iΦ)a),\begin{split}\widetilde{\varphi}={\operatorname{Re}\left(\exp(i\Phi)a\right)},\end{split}

and derive the equations that Φ\Phi and aa should satisfy, based on the given linear operator. The functions Φ\Phi and aa will be referred to as the phase and the amplitude, respectively. In our problem, it is not only nontrivial to choose the correct evolution equations for the phase and amplitude but also to choose the initial data whose associated solution is well-behaved (in particular satisfying the required properties in the above) for a sufficiently long period of time.777When the linear operator is given by a differential operator, this process of extracting the equations for the phase and amplitude is rather straightforward. See Mizohata [25] for the case of linear Schrödinger operator with lower order terms. Furthermore, one may work with a renormalized independent variable xx^{\prime} which makes the coefficient of the principal term into a constant. In turn, this allows us to propagate Φ(t,x)ξ0x\Phi(t,x^{\prime})\sim\xi_{0}\cdot x^{\prime}, see below for the bicharacteristic ODE as well as the equation for Φ\Phi.

Before we proceed to explain how such choices are made, let us give some heuristics for the evolution of wave packets. Roughly, we would like to regard φ~\widetilde{\varphi} as a function which is well-localized both in the physical and Fourier variables, centered at some point (X(t),Ξ(t))(X(t),\Xi(t)) in the phase space. Writing our linearized operator as θ̊=t+ipθ̊\mathcal{L}_{\mathring{\theta}}=\partial_{t}+ip_{\mathring{\theta}} modulo lower order terms and observing that pθ̊p_{\mathring{\theta}} is purely real, we expect that if φ~\widetilde{\varphi} at the initial time corresponds to (X0,Ξ0)(X_{0},\Xi_{0}), then for t>0t>0, φ~(t)\widetilde{\varphi}(t) corresponds to (X(t),Ξ(t))(X(t),\Xi(t)) which is given by the solution to the bicharacteristic ODE

{X˙=ξpθ̊(t,X,Ξ),Ξ˙=xpθ̊(t,X,Ξ)\left\{\begin{aligned} \dot{X}&=\nabla_{\xi}p_{\mathring{\theta}}(t,X,\Xi),\\ \dot{\Xi}&=-\nabla_{x}p_{\mathring{\theta}}(t,X,\Xi)\end{aligned}\right. (2.5)

with initial data (X0,Ξ0)(X_{0},\Xi_{0}). We shall take |Ξ0|1|\Xi_{0}|\gg 1 and X0X_{0} sufficiently close to the degeneracy of θ̊\mathring{\theta}, and choose the sign of Ξ0\Xi_{0} in a way that X(t)X(t) moves towards the degeneracy for t>0t>0.

To actually construct wave packets following the above ODE trajectories, given θ̊\mathcal{L}_{\mathring{\theta}}, we begin with writing

θ̊=t+ipθ̊+sθ̊+rθ̊,\begin{split}\mathcal{L}_{\mathring{\theta}}=\partial_{t}+ip_{\mathring{\theta}}+s_{\mathring{\theta}}+r_{\mathring{\theta}},\end{split} (2.6)

where sθ̊s_{\mathring{\theta}} and rθ̊r_{\mathring{\theta}} are smoother than ipθ̊ip_{\mathring{\theta}} by order at least one and two, respectively. Here, we take

pθ̊:=θ̊ξγ(ξ)+Γθ̊ξ\begin{split}p_{\mathring{\theta}}:=-\nabla^{\perp}\mathring{\theta}\cdot\xi\gamma(\xi)+\nabla^{\perp}\Gamma\mathring{\theta}\cdot\xi\end{split}

and define the equation for Φ\Phi by

tΦ+pθ̊(t,x,Φ)=0.\begin{split}\partial_{t}\Phi+p_{\mathring{\theta}}(t,x,\nabla\Phi)=0.\end{split}

Observe that we have taken the main part of the lower order operator Γ12θ̊Γ12\Gamma^{\frac{1}{2}}\nabla^{\perp}\mathring{\theta}\cdot\nabla\Gamma^{-\frac{1}{2}} into pθ̊p_{\mathring{\theta}}. This is inevitable, as the two terms in pθ̊p_{\mathring{\theta}} are barely distinguishable when γ\gamma is only slightly singular, and in general we do not have any control over the coefficient Γθ̊\nabla^{\perp}\Gamma\mathring{\theta}. Hence, we are forced to design the initial data Φ0\Phi_{0} which takes this term into account. Again, the goal is to propagate that Φ(t,x)Ξ(t)\nabla\Phi(t,x)\simeq\Xi(t) for xX(t)x\simeq X(t) for a sufficiently long interval of time, within which degeneration occurs. Returning to the expression (2.6), note that the difference θ̊(t+ipθ̊)\mathcal{L}_{\mathring{\theta}}-(\partial_{t}+ip_{\mathring{\theta}}) consists of a few commutators. The operator sθ̊s_{\mathring{\theta}} is defined by taking the principal terms of the commutators which are given by Poisson brackets, and then we simply write rθ̊r_{\mathring{\theta}} for the remainder. The equation for aa is then given by

ta+ξpθ̊(x,Φ)a+(122ξpθ̊(x,Φ):2Φ+sθ̊(x,Φ))a=0.\begin{split}\partial_{t}a+\nabla_{\xi}p_{\mathring{\theta}}(x,\nabla\Phi)\cdot\nabla a+\left(\frac{1}{2}\nabla^{2}_{\xi}p_{\mathring{\theta}}(x,\nabla\Phi):\nabla^{2}\Phi+s_{\mathring{\theta}}(x,\nabla\Phi)\right)a=0.\end{split}

The initial data a0a_{0} is taken to be a smooth bump function supported in a small neighborhood of X0X_{0}. The choice of the zeroth order coefficient will become clear from the following discussion.

After solving the equations for Φ\Phi and aa to find φ~=Re(ei𝚽a)\widetilde{\varphi}={\operatorname{Re}(e^{i\mathbf{\Phi}}a)}, we then need to estimate Sobolev norms of both φ~\widetilde{\varphi} and the error θ̊(φ~)\mathcal{L}_{\mathring{\theta}}(\widetilde{\varphi}). To begin with, compared with pθ̊p_{\mathring{\theta}}, the operator rθ̊r_{\mathring{\theta}} is smoother by order 2, and therefore we expect rθ̊(φ~)r_{\mathring{\theta}}(\widetilde{\varphi}) to be small. However, in the error there is also a contribution from ipθ̊ip_{\mathring{\theta}} and sθ̊s_{\mathring{\theta}}, which occurs since eiΦe^{i\Phi} is not completely localized in the frequency space. To see this contribution, we introduce the conjugation operator (Φ)pθ̊(x,D){}^{(\Phi)}p_{\mathring{\theta}}(x,D) defined by

eiΦpθ̊(x,D)(eiΦa)=:(Φ)pθ̊(x,D)a.\begin{split}e^{-i\Phi}p_{\mathring{\theta}}(x,D)(e^{i\Phi}a)=:{}^{(\Phi)}p_{\mathring{\theta}}(x,D)a.\end{split}

We can then formally expand

(Φ)rp:=(Φ)pθ̊(x,ξ)(pθ̊(x,Φ)+ξpθ̊(x,Φ)i2ξ2pθ̊(x,Φ)2Φ).\begin{split}{}^{(\Phi)}r_{p}:={}^{(\Phi)}p_{\mathring{\theta}}(x,\xi)-\left(p_{\mathring{\theta}}(x,\nabla\Phi)+\nabla_{\xi}p_{\mathring{\theta}}(x,\nabla\Phi)-\frac{i}{2}\nabla_{\xi}^{2}p_{\mathring{\theta}}(x,\nabla\Phi)\nabla^{2}\Phi\right).\end{split}

Note that the last two terms are taken into account in the equation for aa above. The symbol for (Φ)rp{}^{(\Phi)}r_{p} can be expressed explicitly by an oscillatory integral, and we need a sharp bound for its operator norm into L2L^{2}. It turns out that both for the oscillatory integral bound and the degeneration estimate, the key step is to find the inverse wave packet scale μ(t)\mu(t) such that μ(t)λ(t)\mu(t)\ll\lambda(t), |k+1Φ(t)|μk(t)λ(t)|\partial^{k+1}\Phi(t)|\lesssim\mu^{k}(t)\lambda(t) (on the support of a(t)a(t)), and |ka(t)|μk(t)|\partial^{k}a(t)|\lesssim\mu^{k}(t), where λ(t)|Φ(t)|(|Ξ(t)|)\lambda(t)\simeq|\partial\Phi(t)|(\simeq|{\Xi(t)}|) is the overall wave packet frequency. Below, let us briefly explain how it is done.

Estimate for the derivatives of the phase and amplitude. A basic difficulty in controlling the high derivatives of Φ\partial\Phi is that 2Φ\partial^{2}\Phi obeys a Ricatti-type ODE along characteristics, which tends to blow up in finite time (indeed, such a blow up corresponds to a focal point along a bicharacteristic). Therefore, some care is needed to control 2Φ\partial^{2}\Phi on a sufficiently long time interval, so as to see the effect of degeneration.

When the shear profile θ̊=f(x2)\mathring{\theta}=f(x_{2}) is time-independent and degenerate at x2=0x_{2}=0, it turns out that the following solution given by separation of variables avoids the basic difficulty and achieves the required bounds:

𝚽(t,x)=Eλ0+λ0x1+x21(γλ01(E+Γff)),\begin{split}\mathbf{\Phi}(t,x)=E\lambda_{0}+\lambda_{0}x_{1}+\partial_{x_{2}}^{-1}\left(\gamma_{\lambda_{0}}^{-1}\left(\frac{E+\Gamma f^{\prime}}{-f^{\prime}}\right)\right),\end{split}

where

E=f(x¯)γλ0((112ϵ)λ0)Γf(x¯)\begin{split}E=-f^{\prime}(\overline{x})\gamma_{\lambda_{0}}((1-\tfrac{1}{2}\epsilon)\lambda_{0})-\Gamma f^{\prime}(\overline{x})\end{split}

with some x¯>0\overline{x}>0 and ϵ>0\epsilon>0 sufficiently small, γλ0():=γ(λ0,)\gamma_{\lambda_{0}}(\cdot):=\gamma(\lambda_{0},\cdot), and γλ01\gamma_{\lambda_{0}}^{-1} is the inverse of γλ0\gamma_{\lambda_{0}}. As in the solution of the toy model above, we have separated out the dependence in x1x_{1}, which does not change in time due to the x1x_{1}-independence of the linear operator. While the formula looks somewhat complicated, when pθ̊p_{\mathring{\theta}} is given by a differential operator, this form corresponds to the linear ansatz in the renormalized coordinates in which the coefficient of the principal symbol of pθ̊p_{\mathring{\theta}} is non-degenerate, which seems to be a natural choice. This choice of initial data is especially important in the case of time-dependent background θ̊=f(t,x2)\mathring{\theta}=f(t,x_{2}). As discussed before, already in the case k=1k=1, the equation for x22Φ\partial_{x_{2}}^{2}\Phi becomes a Riccati-type ODE along characteristics, which could in principle grow much faster than |xΦ|2λ(t)2|\partial_{x}\Phi|^{2}\simeq\lambda(t)^{2}. On the other hand, one can check that our choice of Φ(t=0)\Phi(t=0) makes the variable

h(t,X2(t)):=ξ2pθ̊(t,X2(t),Ξ2(t))x2pθ̊(t,X2(t),Ξ2(t))x22Φ(t,X2(t))+1\begin{split}h(t,X_{2}(t)):=\frac{\partial_{\xi_{2}}p_{\mathring{\theta}}(t,X_{2}(t),\Xi_{2}(t))}{\partial_{x_{2}}p_{\mathring{\theta}}(t,X_{2}(t),\Xi_{2}(t))}\partial_{x_{2}}^{2}\Phi(t,X_{2}(t))+1\end{split}

vanish at t=0t=0. (Here we are neglecting the dependence on x1x_{1} and ξ1\xi_{1}.) Moreover, one can check that the quantity hh satisfies a remarkably simple equation, which allows to propagate |h(t)|1|h(t)|\ll 1 for a suitably long time and in turn gives a sharp bound for x22Φ(t,X2(t))\partial_{x_{2}}^{2}\Phi(t,X_{2}(t)). Moreover, we obtain a sharp control on the instantaneous expansion of characteristics

x2(ξ2p(t,x2,x2Φ(t,x2)))|t,x2=X2(t)=(γλ0(Ξ2(t))γλ0(Ξ2(t))γλ0(Ξ2(t))γλ0(Ξ2(t)))Ξ˙2+=:A(t)+.\left.\partial_{x_{2}}(\partial_{\xi_{2}}p(t,x_{2},\partial_{x_{2}}\Phi(t,x_{2})))\right|_{t,x_{2}=X_{2}(t)}=-\left(\frac{\gamma_{\lambda_{0}}^{\prime}(\Xi_{2}(t))}{\gamma_{\lambda_{0}}(\Xi_{2}(t))}-\frac{\gamma_{\lambda_{0}}^{\prime\prime}(\Xi_{2}(t))}{\gamma_{\lambda_{0}}^{\prime}(\Xi_{2}(t))}\right)\dot{\Xi}_{2}+\cdots=:-A(t)+\cdots.

The wave packet scale μ1\mu^{-1} is then nothing but the product of the initial scale Δx0\Delta x_{0} and the integrated expansion factor, i.e.,

μ1Δx0exp(0tA(t)dt)Δx0γλ0(λ0)γλ0(λ(t))γλ0(λ(t))γλ0(λ0)(up to a small power of λ0).\mu^{-1}\simeq\Delta x_{0}\exp\left(-\int_{0}^{t}A(t^{\prime})\mathrm{d}t^{\prime}\right)\simeq\Delta x_{0}\frac{\gamma_{\lambda_{0}}(\lambda_{0})}{\gamma_{\lambda_{0}}(\lambda(t))}\frac{\gamma_{\lambda_{0}}^{\prime}(\lambda(t))}{\gamma_{\lambda_{0}}^{\prime}(\lambda_{0})}\qquad\hbox{(up to a small power of $\lambda_{0}$).}

The actual proof of the bound x2k+1Φμ(t)kλ(t)\partial_{x_{2}}^{k+1}\Phi\lesssim\mu(t)^{k}\lambda(t) proceeds by first propagating sharp bounds for x2k1h(t)\partial_{x_{2}}^{k-1}h(t), and then converting it to that for x2k+1Φ\partial_{x_{2}}^{k+1}\Phi. It turns out to be essential to take advantage of the cancellations embedded in the transformation from x22Φ\partial_{x_{2}}^{2}\Phi to hh, which are very difficult to see when one works directly with the equations for x2k+1Φ\partial_{x_{2}}^{k+1}\Phi. Given the bounds for x2k+1Φ(t)\partial_{x_{2}}^{k+1}\Phi(t), it is relatively straightforward to obtain the bounds for x2ka(t)\partial_{x_{2}}^{k}a(t).

3 Algebraic preliminaries for degenerate wave packet construction

Our goal in Sections 36 is to construct an approximate solution with initial frequency O(λ0)O(\lambda_{0}) to the equation θ̊φ=0\mathcal{L}_{\mathring{\theta}}\varphi=0 that is valid for times [0,11ϵ(λ0)tf(τM)][0,\frac{1}{1-\epsilon(\lambda_{0})}t_{f}(\tau_{M})], where θ̊\mathring{\theta}, MM and τM\tau_{M} obey the hypotheses of either Theorem A (steady state case) or Theorem B (ff is time-dependent but even) and ϵ(λ0)\epsilon(\lambda_{0}) is a small parameter that will be fixed in the construction (see Section 4.1). We look for an approximate solution of the form

φ~=Re(a(t,x)ei𝚽(t,x)),\widetilde{\varphi}=\operatorname{Re}\left(a(t,x)e^{i\mathbf{\Phi}(t,x)}\right),

where the amplitude a(t,x)a(t,x) is “slowly varying” compared to the real-valued phase 𝚽\mathbf{\Phi}, such that θ̊φ~\mathcal{L}_{\mathring{\theta}}\widetilde{\varphi} is sufficiently small. Anticipating the spatial degeneration property of φ~\widetilde{\varphi} (which comes with the frequency growth property that we want), we will refer to such an object φ~\widetilde{\varphi} as a degenerating wave packet adapted to θ̊\mathcal{L}_{\mathring{\theta}}.

In order to construct a degenerating wave packet φ~\widetilde{\varphi} and, more importantly, to bound the error θ̊φ~\mathcal{L}_{\mathring{\theta}}\widetilde{\varphi}, we need to develop tools for computing the action of the multiplier Γ\Gamma on aei𝚽ae^{i\mathbf{\Phi}}. Section 3.1 is devoted to accomplishing this task. Then in Section 3.2, we use the algebraic identities derived in Section 3.1 to specify the construction of the phase 𝚽\mathbf{\Phi} and the amplitude aa of a degenerate wave packet. The tasks of verifying the degeneration property of φ~\widetilde{\varphi} and bounding the error θ̊φ~\mathcal{L}_{\mathring{\theta}}\widetilde{\varphi} are deferred to Section 6, after further necessary tools are developed (see also the end of Section 3.2).

3.1 Some symbolic calculus

In this subsection, it will be convenient to generalize the set-up a bit and work with classical pseudo-differential operators with generalized order on d\mathbb{R}^{d}.

Let m=m(ξ)m=m(\xi) be a smooth, slowly varying even positive symbol on d\mathbb{R}^{d} (as introduced in Section 1.5, γ(ξ)\gamma(\xi) is an example of such a symbol on 2\mathbb{R}^{2}). As in the case of Fourier multipliers, we say that a symbol p=p(x,ξ)p=p(x,\xi) belongs to the class S(m)S(m) if

|xJξIp(x,ξ)|I,Jξ|I|m(ξ) for every pair of multi-indices IJ.|{\partial_{x}^{J}\partial_{\xi}^{I}p(x,\xi)}|\lesssim_{I,J}\langle{\xi}\rangle^{-|{I}|}m(\xi)\quad\hbox{ for every pair of multi-indices $I$, $J$.} (3.1)

When m=ξsm=\langle{\xi}\rangle^{s}, S(m)S(m) coincides with the class of classical symbols of order ss. Nonstandard examples of mm include m(ξ)=logs(10+|ξ|)m(\xi)=\log^{s}(10+|{\xi}|), m(ξ)=logs(10+log(10+|ξ|))m(\xi)=\log^{s}(10+\log(10+|{\xi}|)) etc.

Given p(x,ξ)S(m)p(x,\xi)\in S(m), we define its left quantization to be the operator

p(x,D)u=p(x,ξ)u(y)eiξ(xy)dydξ(2π)d.p(x,D)u=\int p(x,\xi)u(y)e^{i\xi\cdot(x-y)}\,\mathrm{d}y\frac{\mathrm{d}\xi}{(2\pi)^{d}}. (3.2)

We denote by Op(S(m))Op(S(m)) the space of all left quantizations of symbols in S(m)S(m).

We begin by stating the basic symbolic calculus for composition of Op(S(m))Op(S(m))-operators.

Lemma 3.1 (Symbolic calculus).

Let pS(mp)p\in S(m_{p}) and qS(mq)q\in S(m_{q}). Then the composition p(x,D)q(x,D)p(x,D)q(x,D) belongs to Op(S(mpmq))Op(S(m_{p}m_{q})), and for every NN\in\mathbb{N}, its symbol pq(x,ξ)p\circ q(x,\xi) obeys

pq(x,ξ)α:|α|N11α!i|α|ξαp(x,ξ)xαq(x,ξ)S(ξNmpmq).p\circ q(x,\xi)-\sum_{\alpha:|{\alpha}|\leq N-1}\frac{1}{\alpha!i^{|{\alpha}|}}\partial_{\xi}^{\alpha}p(x,\xi)\partial_{x}^{\alpha}q(x,\xi)\in S(\langle{\xi}\rangle^{-N}m_{p}m_{q}). (3.3)

In particular,

pq(x,ξ)qp(x,ξ)i1{p,q}(x,ξ)S(ξ2mpmq),p\circ q(x,\xi)-q\circ p(x,\xi)-i^{-1}\{p,q\}(x,\xi)\in S(\langle{\xi}\rangle^{-2}m_{p}m_{q}), (3.4)

where {p,q}=j(ξjpxjqxjpξjq)\{p,q\}=\sum_{j}\left(\partial_{\xi_{j}}p\partial_{x_{j}}q-\partial_{x_{j}}p\partial_{\xi_{j}}q\right) is the Poisson bracket.

Proof.

The proof of this lemma is analogous to the standard case mp(ξ)=ξspm_{p}(\xi)=\langle{\xi}\rangle^{s_{p}}, mq(ξ)=ξsqm_{q}(\xi)=\langle{\xi}\rangle^{s_{q}}, so we shall only sketch the main points. Formally, the symbol pq(x,ξ)p\circ q(x,\xi) for the composition p(x,D)q(x,D)p(x,D)q(x,D) is given by

pq(x,ξ)=p(x,η)q(y,ξ)ei(xy)(ξη)dη(2π)ddy.p\circ q(x,\xi)=\iint p(x,\eta)q(y,\xi)e^{-i(x-y)\cdot(\xi-\eta)}\frac{\mathrm{d}\eta}{(2\pi)^{d}}\mathrm{d}y.

Then pqS(mpmq)p\circ q\in S(m_{p}m_{q}) and the expansion (3.3) then follows, as usual, by Taylor expansion of p(x,η)p(x,\eta) around η=ξ\eta=\xi, integration by parts in yy and (3.1). Moreover, (3.4) follows from (3.3). ∎

The following conjugation result will be a starting point for our degenerating wave packet construction.

Lemma 3.2 (Conjugation by ei𝚽e^{i\mathbf{\Phi}}).

Let pS(m)p\in S(m) and 𝚽C\mathbf{\Phi}\in C^{\infty}. Then we have

ei𝚽p(x,D)(aei𝚽)=(𝚽)p(x,D)a,\displaystyle e^{-i\mathbf{\Phi}}p(x,D)(ae^{i\mathbf{\Phi}})={}^{(\mathbf{\Phi})}p(x,D)a,

where

(𝚽)p(x,ξ)=p(x,η)ei(𝚽(y)𝚽(x)+(ηξ)(xy))dη(2π)ddy.{}^{(\mathbf{\Phi})}p(x,\xi)=\iint p(x,\eta)e^{i(\mathbf{\Phi}(y)-\mathbf{\Phi}(x)+(\eta-\xi)\cdot(x-y))}\,\frac{\mathrm{d}\eta}{(2\pi)^{d}}\,\mathrm{d}y. (3.5)

Observe carefully that we have not placed (𝚽)p(x,ξ){}^{(\mathbf{\Phi})}p(x,\xi) in any standard symbol class. In Propositions 3.3 and 5.1, we will obtain an expansion and a bound for the operator (𝚽)p(x,ξ){}^{(\mathbf{\Phi})}p(x,\xi) that is adapted to the specific scenario we are interested in.

Proof.

Indeed,

ei𝚽p(x,D)(aei𝚽)\displaystyle e^{-i\mathbf{\Phi}}p(x,D)(ae^{i\mathbf{\Phi}}) =p(x,ξ)a(y)ei(𝚽(y)𝚽(x)+ξ(xy))dξ(2π)ddy\displaystyle=\iint p(x,\xi)a(y)e^{i(\mathbf{\Phi}(y)-\mathbf{\Phi}(x)+\xi\cdot(x-y))}\,\frac{\mathrm{d}\xi}{(2\pi)^{d}}\mathrm{d}y
=p(x,ξ)a^(η)ei(𝚽(y)𝚽(x)+ξ(xy)+ηy)dη(2π)ddξ(2π)ddy\displaystyle=\iiint p(x,\xi)\hat{a}(\eta)e^{i(\mathbf{\Phi}(y)-\mathbf{\Phi}(x)+\xi\cdot(x-y)+\eta\cdot y)}\,\frac{\mathrm{d}\eta}{(2\pi)^{d}}\frac{\mathrm{d}\xi}{(2\pi)^{d}}\mathrm{d}y
=(p(x,ξ)ei(𝚽(y)𝚽(x)+(ξη)(xy)dξ(2π)ddy)a^(η)eixηdη(2π)d,\displaystyle=\int\left(\iint p(x,\xi)e^{i(\mathbf{\Phi}(y)-\mathbf{\Phi}(x)+(\xi-\eta)\cdot(x-y)}\frac{\mathrm{d}\xi}{(2\pi)^{d}}\mathrm{d}y\right)\hat{a}(\eta)e^{ix\cdot\eta}\,\frac{\mathrm{d}\eta}{(2\pi)^{d}},

so switching the variables ξ\xi and η\eta, we obtain the desired claim. ∎

For our construction, we would like to expand (𝚽)p{}^{(\mathbf{\Phi})}p under the assumption that ξ\xi and the characteristic frequency of x𝚽\partial_{x}\mathbf{\Phi} (bounded by μ\mu in what follows) are smaller than the typical magnitude of x𝚽\partial_{x}\mathbf{\Phi} (denoted by λ\lambda in what follows). To begin with, the stationary set for the phase of the oscillatory integral on the RHS of (3.5) is

y𝚽(y)(ηξ)=0,xy=0,\displaystyle\partial_{y}\mathbf{\Phi}(y)-(\eta-\xi)=0,\quad x-y=0,

or equivalently, η=ξ+y𝚽(y)\eta=\xi+\partial_{y}\mathbf{\Phi}(y) and y=xy=x. Our assumption leads us to also expand in ξ\xi about ξ=0\xi=0. Following such a route, we are led to the following formulae for the expansion and the remainder.

Proposition 3.3 (Formal expansion of (𝚽)p{}^{(\mathbf{\Phi})}p).

Let pS(m)p\in S(m) and 𝚽C\mathbf{\Phi}\in C^{\infty}. The symbol (𝚽)p(x,ξ){}^{(\mathbf{\Phi})}p(x,\xi) admits expansions of the form

(𝚽)p(x,ξ)\displaystyle{}^{(\mathbf{\Phi})}p(x,\xi) =p(x,x𝚽)+(𝚽)rp,1(x,ξ)\displaystyle=p(x,\partial_{x}\mathbf{\Phi})+{}^{(\mathbf{\Phi})}r_{p,-1}(x,\xi) (3.6)
=p(x,x𝚽)+jξjξjp(x,x𝚽(x))i2j,kξjξkp(x,x𝚽(x))jk𝚽(x)\displaystyle=p(x,\partial_{x}\mathbf{\Phi})+\sum_{j}\xi_{j}\partial_{\xi_{j}}p(x,\partial_{x}\mathbf{\Phi}(x))-\frac{i}{2}\sum_{j,k}\partial_{\xi_{j}}\partial_{\xi_{k}}p(x,\partial_{x}\mathbf{\Phi}(x))\partial_{j}\partial_{k}\mathbf{\Phi}(x)
+(𝚽)rp,2(x,ξ),\displaystyle\mathrel{\phantom{=}}+{}^{(\mathbf{\Phi})}r_{p,-2}(x,\xi),

where

(𝚽)rp,1(x,ξ)\displaystyle{}^{(\mathbf{\Phi})}r_{p,-1}(x,\xi) (3.7)
=(01ξjp(x,ση+(1σ)y𝚽(y))ei(𝚽(y)𝚽(x)+(ηξ)(xy))dη(2π)ddydσ)ξj\displaystyle=\left(\int_{0}^{1}\iint\partial_{\xi_{j}}p(x,\sigma\eta+(1-\sigma)\partial_{y}\mathbf{\Phi}(y))e^{i(\mathbf{\Phi}(y)-\mathbf{\Phi}(x)+(\eta-\xi)\cdot(x-y))}\,\frac{\mathrm{d}\eta}{(2\pi)^{d}}\,\mathrm{d}y\mathrm{d}\sigma\right)\xi_{j}
i01(1σ)ξjξkp(x,ση+(1σ)y𝚽(y))jk𝚽(y)\displaystyle\mathrel{\phantom{=}}-i\int_{0}^{1}(1-\sigma)\iint\partial_{\xi_{j}}\partial_{\xi_{k}}p(x,\sigma\eta+(1-\sigma)\partial_{y}\mathbf{\Phi}(y))\partial_{j}\partial_{k}\mathbf{\Phi}(y)
×ei(𝚽(y)𝚽(x)+(ηξ)(xy))dη(2π)ddydσ,\displaystyle\phantom{\mathrel{\phantom{=}}-i\int_{0}^{1}(1-\sigma)\iint}\times e^{i(\mathbf{\Phi}(y)-\mathbf{\Phi}(x)+(\eta-\xi)\cdot(x-y))}\,\frac{\mathrm{d}\eta}{(2\pi)^{d}}\,\mathrm{d}y\mathrm{d}\sigma,

and

(𝚽)rp,2(x,ξ)\displaystyle{}^{(\mathbf{\Phi})}r_{p,-2}(x,\xi) (3.8)
=(01(1σ)ξjξkp(x,ση+(1σ)y𝚽(y))ei(𝚽(y)𝚽(x)+(ηξ)(xy))dη(2π)ddydσ)ξjξk\displaystyle=\left(\int_{0}^{1}(1-\sigma)\iint\partial_{\xi_{j}}\partial_{\xi_{k}}p(x,\sigma\eta+(1-\sigma)\partial_{y}\mathbf{\Phi}(y))e^{i(\mathbf{\Phi}(y)-\mathbf{\Phi}(x)+(\eta-\xi)\cdot(x-y))}\,\frac{\mathrm{d}\eta}{(2\pi)^{d}}\,\mathrm{d}y\mathrm{d}\sigma\right)\xi_{j}\xi_{k}
i(01(1σ)(32σ)ξjξkξp(x,ση+(1σ)y𝚽(y))k𝚽(y)\displaystyle\mathrel{\phantom{=}}-i\left(\int_{0}^{1}(1-\sigma)\left(\frac{3}{2}-\sigma\right)\iint\partial_{\xi_{j}}\partial_{\xi_{k}}\partial_{\xi_{\ell}}p(x,\sigma\eta+(1-\sigma)\partial_{y}\mathbf{\Phi}(y))\partial_{k}\partial_{\ell}\mathbf{\Phi}(y)\right.
×ei(𝚽(y)𝚽(x)+(ηξ)(xy))dη(2π)ddydσ)ξj\displaystyle\left.\phantom{\mathrel{\phantom{=}}-i\int_{0}^{1}(1-\sigma)(\frac{3}{2}-\sigma)\iint}\times e^{i(\mathbf{\Phi}(y)-\mathbf{\Phi}(x)+(\eta-\xi)\cdot(x-y))}\,\frac{\mathrm{d}\eta}{(2\pi)^{d}}\,\mathrm{d}y\mathrm{d}\sigma\right)\xi_{j}
1201(1σ)2ξjξkξp(x,ση+(1σ)y𝚽(y))jk𝚽(y)\displaystyle\mathrel{\phantom{=}}-\frac{1}{2}\int_{0}^{1}(1-\sigma)^{2}\iint\partial_{\xi_{j}}\partial_{\xi_{k}}\partial_{\xi_{\ell}}p(x,\sigma\eta+(1-\sigma)\partial_{y}\mathbf{\Phi}(y))\partial_{j}\partial_{k}\partial_{\ell}\mathbf{\Phi}(y)
×ei(𝚽(y)𝚽(x)+(ηξ)(xy))dη(2π)ddydσ\displaystyle\phantom{\mathrel{\phantom{=}}-\frac{1}{2}\int_{0}^{1}(1-\sigma)^{2}\iint}\times e^{i(\mathbf{\Phi}(y)-\mathbf{\Phi}(x)+(\eta-\xi)\cdot(x-y))}\,\frac{\mathrm{d}\eta}{(2\pi)^{d}}\,\mathrm{d}y\mathrm{d}\sigma
1201(1σ)2ξjξkξξmp(x,ση+(1σ)y𝚽(y))jm𝚽(y)k𝚽(y)\displaystyle\mathrel{\phantom{=}}-\frac{1}{2}\int_{0}^{1}(1-\sigma)^{2}\iint\partial_{\xi_{j}}\partial_{\xi_{k}}\partial_{\xi_{\ell}}\partial_{\xi_{m}}p(x,\sigma\eta+(1-\sigma)\partial_{y}\mathbf{\Phi}(y))\partial_{j}\partial_{m}\mathbf{\Phi}(y)\partial_{k}\partial_{\ell}\mathbf{\Phi}(y)
×ei(𝚽(y)𝚽(x)+(ηξ)(xy))dη(2π)ddydσ,\displaystyle\phantom{\mathrel{\phantom{=}}-\frac{1}{2}\int_{0}^{1}(1-\sigma)^{2}\iint}\times e^{i(\mathbf{\Phi}(y)-\mathbf{\Phi}(x)+(\eta-\xi)\cdot(x-y))}\,\frac{\mathrm{d}\eta}{(2\pi)^{d}}\,\mathrm{d}y\mathrm{d}\sigma,

where the repeated indices in (3.7) and (3.8) are implicitly summed.

The formal expansion of (𝚽)p{}^{(\mathbf{\Phi})}p is standard; see, for instance, [35, Chapter VII]. The point of Proposition 3.3 is the explicit formulae for the remainder symbols (𝚽)rp,j{}^{(\mathbf{\Phi})}r_{p,-j}, which we shall analyze in Section 5 with the specific estimates on xkx𝚽\partial_{x}^{k}\partial_{x}\mathbf{\Phi} in our problem, to be proved in Section 4.

Proof.

To ease the notation, in what follows, we work with the convention that the repeated indices are summed. Moreover, we introduce the nn-th order remainder symbol

r(n)j1jn[p(x,)](ξ1,ξ0)=01ξj1ξjnp(x,σξ1+(1σ)ξ0)n(1σ)n1dσ.r^{(n)j_{1}\ldots j_{n}}[p(x,\cdot)](\xi_{1},\xi_{0})=\int_{0}^{1}\partial_{\xi_{j_{1}}}\cdots\partial_{\xi_{j_{n}}}p(x,\sigma\xi_{1}+(1-\sigma)\xi_{0})n(1-\sigma)^{n-1}\,\mathrm{d}\sigma. (3.9)

Then the (n1)(n-1)-th Taylor expansion of p(x,ξ)p(x,\xi) in ξ\xi about ξ0\xi_{0} takes the form

p(x,ξ1)=m=0n1m!ξj1ξjmp(x,ξ0)(ξ1ξ0)j1(ξ1ξ0)jm+1n!r(n)j1jn[p(x,)](ξ1,ξ0)(ξ1ξ0)j1(ξ1ξ0)jn.\begin{split}p(x,\xi_{1})&=\sum_{m=0}^{n}\frac{1}{m!}\partial_{\xi_{j_{1}}}\cdots\partial_{\xi_{j_{m}}}p(x,\xi_{0})(\xi_{1}-\xi_{0})^{j_{1}}\cdots(\xi_{1}-\xi_{0})^{j_{m}}\\ &\qquad+\frac{1}{n!}r^{(n)j_{1}\ldots j_{n}}[p(x,\cdot)](\xi_{1},\xi_{0})(\xi_{1}-\xi_{0})^{j_{1}}\cdots(\xi_{1}-\xi_{0})^{j_{n}}.\end{split} (3.10)

Also note the useful recursive identity (which is simply an integration by parts in σ\sigma)

r(n)j1jn[p(x,)](ξ1,ξ0)=ξj1ξjnp(x,ξ0)+jn+11n+1r(n+1)j1jnjn+1[p(x,)](ξ1,ξ0)(ξ1ξ0)jn+1.r^{(n)j_{1}\ldots j_{n}}[p(x,\cdot)](\xi_{1},\xi_{0})=\partial_{\xi_{j_{1}}}\cdots\partial_{\xi_{j_{n}}}p(x,\xi_{0})+\sum_{j_{n+1}}\frac{1}{n+1}r^{(n+1)j_{1}\ldots j_{n}j_{n+1}}[p(x,\cdot)](\xi_{1},\xi_{0})(\xi_{1}-\xi_{0})^{j_{n+1}}. (3.11)

In view of the stationary set of the phase and the expectation that |ξ||x𝚽||{\xi}|\ll|{\partial_{x}\mathbf{\Phi}}|, we wish to expand p(x,η)p(x,\eta) in η\eta about η=ξ+y𝚽(y)\eta=\xi+\partial_{y}\mathbf{\Phi}(y):

(𝚽)p(x,ξ)\displaystyle{}^{(\mathbf{\Phi})}p(x,\xi) =p(x,y𝚽(y))ei(𝚽(y)𝚽(x)+(ηξ)(xy))dη(2π)ddy\displaystyle=\iint p(x,\partial_{y}\mathbf{\Phi}(y))e^{i(\mathbf{\Phi}(y)-\mathbf{\Phi}(x)+(\eta-\xi)\cdot(x-y))}\,\frac{\mathrm{d}\eta}{(2\pi)^{d}}\,\mathrm{d}y
+r(1)j[p(x,)](η,y𝚽(y))(ηjj𝚽(y))ei(𝚽(y)𝚽(x)+(ηξ)(xy))dη(2π)ddy.\displaystyle\mathrel{\phantom{=}}+\iint r^{(1)j}[p(x,\cdot)](\eta,\partial_{y}\mathbf{\Phi}(y))(\eta_{j}-\partial_{j}\mathbf{\Phi}(y))e^{i(\mathbf{\Phi}(y)-\mathbf{\Phi}(x)+(\eta-\xi)\cdot(x-y))}\,\frac{\mathrm{d}\eta}{(2\pi)^{d}}\,\mathrm{d}y.

Note the identity

(ηjj𝚽(y))ei(𝚽(y)𝚽(x)+(ηξ)(xy))\displaystyle(\eta_{j}-\partial_{j}\mathbf{\Phi}(y))e^{i(\mathbf{\Phi}(y)-\mathbf{\Phi}(x)+(\eta-\xi)\cdot(x-y))}
=(ξj+ηjξjj𝚽(y))ei(𝚽(y)𝚽(x)+(ηξ)(xy))\displaystyle=(\xi_{j}+\eta_{j}-\xi_{j}-\partial_{j}\mathbf{\Phi}(y))e^{i(\mathbf{\Phi}(y)-\mathbf{\Phi}(x)+(\eta-\xi)\cdot(x-y))}
=ξjei(𝚽(y)𝚽(x)+(ηξ)(xy))+iyjei(𝚽(y)𝚽(x)+(ηξ)(xy)).\displaystyle=\xi_{j}e^{i(\mathbf{\Phi}(y)-\mathbf{\Phi}(x)+(\eta-\xi)\cdot(x-y))}+i\partial_{y^{j}}e^{i(\mathbf{\Phi}(y)-\mathbf{\Phi}(x)+(\eta-\xi)\cdot(x-y))}. (3.12)

After integrating yj\partial_{y^{j}} by parts, we obtain

(𝚽)p(x,ξ)\displaystyle{}^{(\mathbf{\Phi})}p(x,\xi) =p(x,x𝚽)+ξjr(1)j[p(x,)](η,y𝚽(y))ei(𝚽(y)𝚽(x)+(ηξ)(xy))dη(2π)ddy\displaystyle=p(x,\partial_{x}\mathbf{\Phi})+\xi_{j}\iint r^{(1)j}[p(x,\cdot)](\eta,\partial_{y}\mathbf{\Phi}(y))e^{i(\mathbf{\Phi}(y)-\mathbf{\Phi}(x)+(\eta-\xi)\cdot(x-y))}\,\frac{\mathrm{d}\eta}{(2\pi)^{d}}\,\mathrm{d}y
+r(1)j[p(x,)](η,y𝚽(y))(ηjξjj𝚽(y))ei(𝚽(y)𝚽(x)+(ηξ)(xy))dη(2π)ddy\displaystyle\mathrel{\phantom{=}}+\iint r^{(1)j}[p(x,\cdot)](\eta,\partial_{y}\mathbf{\Phi}(y))(\eta_{j}-\xi_{j}-\partial_{j}\mathbf{\Phi}(y))e^{i(\mathbf{\Phi}(y)-\mathbf{\Phi}(x)+(\eta-\xi)\cdot(x-y))}\,\frac{\mathrm{d}\eta}{(2\pi)^{d}}\,\mathrm{d}y
=p(x,x𝚽)+ξjr(1)j[p(x,)](η,y𝚽(y))ei(𝚽(y)𝚽(x)+(ηξ)(xy))dη(2π)ddy\displaystyle=p(x,\partial_{x}\mathbf{\Phi})+\xi_{j}\iint r^{(1)j}[p(x,\cdot)](\eta,\partial_{y}\mathbf{\Phi}(y))e^{i(\mathbf{\Phi}(y)-\mathbf{\Phi}(x)+(\eta-\xi)\cdot(x-y))}\,\frac{\mathrm{d}\eta}{(2\pi)^{d}}\,\mathrm{d}y
+r(1)j[p(x,)](η,y𝚽(y))iyjei(𝚽(y)𝚽(x)+(ηξ)(xy))dη(2π)ddy\displaystyle\mathrel{\phantom{=}}+\iint r^{(1)j}[p(x,\cdot)](\eta,\partial_{y}\mathbf{\Phi}(y))i\partial_{y^{j}}e^{i(\mathbf{\Phi}(y)-\mathbf{\Phi}(x)+(\eta-\xi)\cdot(x-y))}\,\frac{\mathrm{d}\eta}{(2\pi)^{d}}\,\mathrm{d}y
=p(x,x𝚽)+ξjr(1)j[p(x,)](η,y𝚽(y))ei(𝚽(y)𝚽(x)+(ηξ)(xy))dη(2π)ddy\displaystyle=p(x,\partial_{x}\mathbf{\Phi})+\xi_{j}\iint r^{(1)j}[p(x,\cdot)](\eta,\partial_{y}\mathbf{\Phi}(y))e^{i(\mathbf{\Phi}(y)-\mathbf{\Phi}(x)+(\eta-\xi)\cdot(x-y))}\,\frac{\mathrm{d}\eta}{(2\pi)^{d}}\,\mathrm{d}y
i(yj(r(1)j[p(x,)](η,y𝚽(y))))ei(𝚽(y)𝚽(x)+(ηξ)(xy))dη(2π)ddy,\displaystyle\mathrel{\phantom{=}}-i\iint\left(\partial_{y^{j}}\left(r^{(1)j}[p(x,\cdot)](\eta,\partial_{y}\mathbf{\Phi}(y))\right)\right)e^{i(\mathbf{\Phi}(y)-\mathbf{\Phi}(x)+(\eta-\xi)\cdot(x-y))}\,\frac{\mathrm{d}\eta}{(2\pi)^{d}}\,\mathrm{d}y,

which already proves (3.7).

We are interested in the next order terms, so we expand

r(1)j[p(x,)](η,y𝚽(y))=ξjp(x,y𝚽(y))+12r(2)jk[p(x,)](η,y𝚽(y))(ηkk𝚽(y)).r^{(1)j}[p(x,\cdot)](\eta,\partial_{y}\mathbf{\Phi}(y))=\partial_{\xi_{j}}p(x,\partial_{y}\mathbf{\Phi}(y))+\frac{1}{2}r^{(2)jk}[p(x,\cdot)](\eta,\partial_{y}\mathbf{\Phi}(y))(\eta_{k}-\partial_{k}\mathbf{\Phi}(y)).

Thus,

(𝚽)p(x,ξ)\displaystyle{}^{(\mathbf{\Phi})}p(x,\xi) =p(x,x𝚽)+ξj(ξjp)(x,x𝚽(x))iyjξjp(x,y𝚽(y))|y=x\displaystyle=p(x,\partial_{x}\mathbf{\Phi})+\xi_{j}(\partial_{\xi_{j}}p)(x,\partial_{x}\mathbf{\Phi}(x))-\left.i\partial_{y^{j}}\partial_{\xi_{j}}p(x,\partial_{y}\mathbf{\Phi}(y))\right|_{y=x}
+12ξjr(2)jk[p(x,)](η,y𝚽(y))(ηkk𝚽(y))ei(𝚽(y)𝚽(x)+(ηξ)(xy))dη(2π)ddy\displaystyle\mathrel{\phantom{=}}+\frac{1}{2}\xi_{j}\iint r^{(2)jk}[p(x,\cdot)](\eta,\partial_{y}\mathbf{\Phi}(y))(\eta_{k}-\partial_{k}\mathbf{\Phi}(y))e^{i(\mathbf{\Phi}(y)-\mathbf{\Phi}(x)+(\eta-\xi)\cdot(x-y))}\,\frac{\mathrm{d}\eta}{(2\pi)^{d}}\,\mathrm{d}y
i2(yj(r(2)jk[p(x,)](η,y𝚽(y))(ηkk𝚽(y))))ei(𝚽(y)𝚽(x)+(ηξ)(xy))dη(2π)ddy\displaystyle\mathrel{\phantom{=}}-\frac{i}{2}\iint\left(\partial_{y^{j}}\left(r^{(2)jk}[p(x,\cdot)](\eta,\partial_{y}\mathbf{\Phi}(y))(\eta_{k}-\partial_{k}\mathbf{\Phi}(y))\right)\right)e^{i(\mathbf{\Phi}(y)-\mathbf{\Phi}(x)+(\eta-\xi)\cdot(x-y))}\,\frac{\mathrm{d}\eta}{(2\pi)^{d}}\,\mathrm{d}y
=p(x,x𝚽)+ξj(ξjp)(x,x𝚽(x))iξjξkp(x,x𝚽(x))jk𝚽(x)\displaystyle=p(x,\partial_{x}\mathbf{\Phi})+\xi_{j}(\partial_{\xi_{j}}p)(x,\partial_{x}\mathbf{\Phi}(x))-i\partial_{\xi_{j}}\partial_{\xi_{k}}p(x,\partial_{x}\mathbf{\Phi}(x))\partial_{j}\partial_{k}\mathbf{\Phi}(x)
+12ξjr(2)jk[p(x,)](η,y𝚽(y))(ηkk𝚽(y))ei(𝚽(y)𝚽(x)+(ηξ)(xy))dη(2π)ddy\displaystyle\mathrel{\phantom{=}}+\frac{1}{2}\xi_{j}\iint r^{(2)jk}[p(x,\cdot)](\eta,\partial_{y}\mathbf{\Phi}(y))(\eta_{k}-\partial_{k}\mathbf{\Phi}(y))e^{i(\mathbf{\Phi}(y)-\mathbf{\Phi}(x)+(\eta-\xi)\cdot(x-y))}\,\frac{\mathrm{d}\eta}{(2\pi)^{d}}\,\mathrm{d}y
+i2r(2)jk[p(x,)](η,y𝚽(y))jk𝚽(y)ei(𝚽(y)𝚽(x)+(ηξ)(xy))dη(2π)ddy\displaystyle\mathrel{\phantom{=}}+\frac{i}{2}\iint r^{(2)jk}[p(x,\cdot)](\eta,\partial_{y}\mathbf{\Phi}(y))\partial_{j}\partial_{k}\mathbf{\Phi}(y)e^{i(\mathbf{\Phi}(y)-\mathbf{\Phi}(x)+(\eta-\xi)\cdot(x-y))}\,\frac{\mathrm{d}\eta}{(2\pi)^{d}}\,\mathrm{d}y
i2(yj(r(2)jk[p(x,)](η,y𝚽(y))))(ηkk𝚽(y))ei(𝚽(y)𝚽(x)+(ηξ)(xy))dη(2π)ddy.\displaystyle\mathrel{\phantom{=}}-\frac{i}{2}\iint\left(\partial_{y^{j}}\left(r^{(2)jk}[p(x,\cdot)](\eta,\partial_{y}\mathbf{\Phi}(y))\right)\right)(\eta_{k}-\partial_{k}\mathbf{\Phi}(y))e^{i(\mathbf{\Phi}(y)-\mathbf{\Phi}(x)+(\eta-\xi)\cdot(x-y))}\,\frac{\mathrm{d}\eta}{(2\pi)^{d}}\,\mathrm{d}y.

For the second to last term, we also expand

r(2)jk[p(x,)](η,y𝚽(y))=ξjξkp(x,y𝚽(y))+13r(3)jk[p(x,)](η,y𝚽(y))(η𝚽(y)),r^{(2)jk}[p(x,\cdot)](\eta,\partial_{y}\mathbf{\Phi}(y))=\partial_{\xi_{j}}\partial_{\xi_{k}}p(x,\partial_{y}\mathbf{\Phi}(y))+\frac{1}{3}r^{(3)jk\ell}[p(x,\cdot)](\eta,\partial_{y}\mathbf{\Phi}(y))(\eta_{\ell}-\partial_{\ell}\mathbf{\Phi}(y)),

so that

(𝚽)p(x,ξ)\displaystyle{}^{(\mathbf{\Phi})}p(x,\xi) =p(x,x𝚽)+ξj(ξjp)(x,x𝚽(x))i2ξjξkp(x,x𝚽(x))jk𝚽(x)\displaystyle=p(x,\partial_{x}\mathbf{\Phi})+\xi_{j}(\partial_{\xi_{j}}p)(x,\partial_{x}\mathbf{\Phi}(x))-\frac{i}{2}\partial_{\xi_{j}}\partial_{\xi_{k}}p(x,\partial_{x}\mathbf{\Phi}(x))\partial_{j}\partial_{k}\mathbf{\Phi}(x)
+12ξjr(2)jk[p(x,)](η,y𝚽(y))(ηkk𝚽(y))ei(𝚽(y)𝚽(x)+(ηξ)(xy))dη(2π)ddy\displaystyle\mathrel{\phantom{=}}+\frac{1}{2}\xi_{j}\iint r^{(2)jk}[p(x,\cdot)](\eta,\partial_{y}\mathbf{\Phi}(y))(\eta_{k}-\partial_{k}\mathbf{\Phi}(y))e^{i(\mathbf{\Phi}(y)-\mathbf{\Phi}(x)+(\eta-\xi)\cdot(x-y))}\,\frac{\mathrm{d}\eta}{(2\pi)^{d}}\,\mathrm{d}y
+i6r(3)jk[p(x,)](η,y𝚽(y))jk𝚽(y)(η𝚽(y))ei(𝚽(y)𝚽(x)+(ηξ)(xy))dη(2π)ddy\displaystyle\mathrel{\phantom{=}}+\frac{i}{6}\iint r^{(3)jk\ell}[p(x,\cdot)](\eta,\partial_{y}\mathbf{\Phi}(y))\partial_{j}\partial_{k}\mathbf{\Phi}(y)(\eta_{\ell}-\partial_{\ell}\mathbf{\Phi}(y))e^{i(\mathbf{\Phi}(y)-\mathbf{\Phi}(x)+(\eta-\xi)\cdot(x-y))}\,\frac{\mathrm{d}\eta}{(2\pi)^{d}}\,\mathrm{d}y
i2(yj(r(2)jk[p(x,)](η,y𝚽(y))))(ηkk𝚽(y))ei(𝚽(y)𝚽(x)+(ηξ)(xy))dη(2π)ddy.\displaystyle\mathrel{\phantom{=}}-\frac{i}{2}\iint\left(\partial_{y^{j}}\left(r^{(2)jk}[p(x,\cdot)](\eta,\partial_{y}\mathbf{\Phi}(y))\right)\right)(\eta_{k}-\partial_{k}\mathbf{\Phi}(y))e^{i(\mathbf{\Phi}(y)-\mathbf{\Phi}(x)+(\eta-\xi)\cdot(x-y))}\,\frac{\mathrm{d}\eta}{(2\pi)^{d}}\,\mathrm{d}y.

Using (3.12) and integrating by parts, we get

(𝚽)p(x,ξ)\displaystyle{}^{(\mathbf{\Phi})}p(x,\xi) =p(x,x𝚽)+ξj(ξjp)(x,x𝚽(x))i2ξjξkp(x,x𝚽(x))jk𝚽(x)\displaystyle=p(x,\partial_{x}\mathbf{\Phi})+\xi_{j}(\partial_{\xi_{j}}p)(x,\partial_{x}\mathbf{\Phi}(x))-\frac{i}{2}\partial_{\xi_{j}}\partial_{\xi_{k}}p(x,\partial_{x}\mathbf{\Phi}(x))\partial_{j}\partial_{k}\mathbf{\Phi}(x)
+12ξjξkr(2)jk[p(x,)](η,y𝚽(y))ei(𝚽(y)𝚽(x)+(ηξ)(xy))dη(2π)ddy\displaystyle\mathrel{\phantom{=}}+\frac{1}{2}\xi_{j}\xi_{k}\iint r^{(2)jk}[p(x,\cdot)](\eta,\partial_{y}\mathbf{\Phi}(y))e^{i(\mathbf{\Phi}(y)-\mathbf{\Phi}(x)+(\eta-\xi)\cdot(x-y))}\,\frac{\mathrm{d}\eta}{(2\pi)^{d}}\,\mathrm{d}y
i2ξj(yk(r(2)jk[p(x,)](η,y𝚽(y))))ei(𝚽(y)𝚽(x)+(ηξ)(xy))dη(2π)ddy\displaystyle\mathrel{\phantom{=}}-\frac{i}{2}\xi_{j}\iint\left(\partial_{y^{k}}\left(r^{(2)jk}[p(x,\cdot)](\eta,\partial_{y}\mathbf{\Phi}(y))\right)\right)e^{i(\mathbf{\Phi}(y)-\mathbf{\Phi}(x)+(\eta-\xi)\cdot(x-y))}\,\frac{\mathrm{d}\eta}{(2\pi)^{d}}\,\mathrm{d}y
+i6ξr(3)jk[p(x,)](η,y𝚽(y))jk𝚽(y)ei(𝚽(y)𝚽(x)+(ηξ)(xy))dη(2π)ddy\displaystyle\mathrel{\phantom{=}}+\frac{i}{6}\xi_{\ell}\iint r^{(3)jk\ell}[p(x,\cdot)](\eta,\partial_{y}\mathbf{\Phi}(y))\partial_{j}\partial_{k}\mathbf{\Phi}(y)e^{i(\mathbf{\Phi}(y)-\mathbf{\Phi}(x)+(\eta-\xi)\cdot(x-y))}\,\frac{\mathrm{d}\eta}{(2\pi)^{d}}\,\mathrm{d}y
+16(y(r(3)jk[p(x,)](η,y𝚽(y))jk𝚽(y)))ei(𝚽(y)𝚽(x)+(ηξ)(xy))dη(2π)ddy\displaystyle\mathrel{\phantom{=}}+\frac{1}{6}\iint\left(\partial_{y^{\ell}}\left(r^{(3)jk\ell}[p(x,\cdot)](\eta,\partial_{y}\mathbf{\Phi}(y))\partial_{j}\partial_{k}\mathbf{\Phi}(y)\right)\right)e^{i(\mathbf{\Phi}(y)-\mathbf{\Phi}(x)+(\eta-\xi)\cdot(x-y))}\,\frac{\mathrm{d}\eta}{(2\pi)^{d}}\,\mathrm{d}y
i2ξk(yj(r(2)jk[p(x,)](η,y𝚽(y))))ei(𝚽(y)𝚽(x)+(ηξ)(xy))dη(2π)ddy\displaystyle\mathrel{\phantom{=}}-\frac{i}{2}\xi_{k}\iint\left(\partial_{y^{j}}\left(r^{(2)jk}[p(x,\cdot)](\eta,\partial_{y}\mathbf{\Phi}(y))\right)\right)e^{i(\mathbf{\Phi}(y)-\mathbf{\Phi}(x)+(\eta-\xi)\cdot(x-y))}\,\frac{\mathrm{d}\eta}{(2\pi)^{d}}\,\mathrm{d}y
12(yjyk(r(2)jk[p(x,)](η,y𝚽(y))))ei(𝚽(y)𝚽(x)+(ηξ)(xy))dη(2π)ddy.\displaystyle\mathrel{\phantom{=}}-\frac{1}{2}\iint\left(\partial_{y^{j}}\partial_{y^{k}}\left(r^{(2)jk}[p(x,\cdot)](\eta,\partial_{y}\mathbf{\Phi}(y))\right)\right)e^{i(\mathbf{\Phi}(y)-\mathbf{\Phi}(x)+(\eta-\xi)\cdot(x-y))}\,\frac{\mathrm{d}\eta}{(2\pi)^{d}}\,\mathrm{d}y.

Hence, we obtain (3.6) with (𝚽)rp(x,ξ){}^{(\mathbf{\Phi})}r_{p}(x,\xi) consisting of all terms on the RHS except for the first three terms. Recalling (3.9), (3.8) then follows after a straightforward computation. ∎

3.2 Specialization to θ̊\mathcal{L}_{\mathring{\theta}} and equations for the phase and the amplitude

We now return to the problem at hand. In view of the x1x_{1}-independence of θ̊\mathring{\theta}, we shall choose 𝚽\mathbf{\Phi} and aa in the separated form

𝚽(t,x)=λ0x1+Φ(t,x2),a=a(t,x2),\mathbf{\Phi}(t,x)=\lambda_{0}x_{1}+\Phi(t,x_{2}),\quad a=a(t,x_{2}),

where λ0\lambda_{0} is an integer. Using standard symbolic calculus (Lemma 3.1), we first rewrite θ̊\mathcal{L}_{\mathring{\theta}} in a form that is more convenient to apply Lemma 3.2 and Proposition 3.3.

Proposition 3.4.

Let θ̊(t,x)=f(t,x2)\mathring{\theta}(t,x)=f(t,x_{2}) and λ0\lambda_{0}\in\mathbb{Z}. We have

θ̊(eiλ0x1ψ)=eiλ0x1(tψ+ipθ̊,λ0(x2,D2)ψ+sθ̊,λ0(x2,D2)ψ+rθ̊,λ0(x2,D2)ψ),\mathcal{L}_{\mathring{\theta}}(e^{i\lambda_{0}x_{1}}\psi)=e^{i\lambda_{0}x_{1}}\left(\partial_{t}\psi+ip_{\mathring{\theta},\lambda_{0}}(x_{2},D_{2})\psi+s_{\mathring{\theta},\lambda_{0}}(x_{2},D_{2})\psi+r_{\mathring{\theta},\lambda_{0}}(x_{2},D_{2})\psi\right), (3.13)

where

pθ̊,λ0(x2,ξ2)\displaystyle p_{\mathring{\theta},\lambda_{0}}(x_{2},\xi_{2}) =x2f(t,x2)λ0γλ0(ξ2)+(x2Γf)(t,x2)λ0,\displaystyle=\partial_{x_{2}}f(t,x_{2})\lambda_{0}\gamma_{\lambda_{0}}(\xi_{2})+(\partial_{x_{2}}\Gamma f)(t,x_{2})\lambda_{0},
sθ̊,λ0(x2,ξ2)\displaystyle s_{\mathring{\theta},\lambda_{0}}(x_{2},\xi_{2}) =12x22f(t,x2)ξ2γλ0(ξ2)λ012(x22Γf)(t,x2)γλ01(ξ2)ξ2γλ0(ξ2)λ0,\displaystyle=-\frac{1}{2}\partial_{x_{2}}^{2}f(t,x_{2})\partial_{\xi_{2}}\gamma_{\lambda_{0}}(\xi_{2})\lambda_{0}-\frac{1}{2}(\partial_{x_{2}}^{2}\Gamma f)(t,x_{2})\gamma_{\lambda_{0}}^{-1}(\xi_{2})\partial_{\xi_{2}}\gamma_{\lambda_{0}}(\xi_{2})\lambda_{0},
rθ̊,λ0(x2,ξ2)\displaystyle r_{\mathring{\theta},\lambda_{0}}(x_{2},\xi_{2}) S(λ01+λ02+ξ22γ(λ0,ξ2)).\displaystyle\in S\left({\tfrac{\lambda_{0}}{1+\lambda_{0}^{2}+\xi_{2}^{2}}\gamma(\lambda_{0},\xi_{2})}\right).

Observe that both pθ̊,λ0(x2,ξ2)p_{\mathring{\theta},\lambda_{0}}(x_{2},\xi_{2}) and sθ̊,λ0(x2,ξ2)s_{\mathring{\theta},\lambda_{0}}(x_{2},\xi_{2}) are real-valued and even.

Proof.

We begin by computing

θ̊φ\displaystyle\mathcal{L}_{\mathring{\theta}}\varphi =tφΓ12θ̊Γ12φ+Γ12Γθ̊Γ12φ\displaystyle=\partial_{t}\varphi-\Gamma^{\frac{1}{2}}\nabla^{\perp}\mathring{\theta}\cdot\nabla\Gamma^{\frac{1}{2}}\varphi+\Gamma^{\frac{1}{2}}\nabla^{\perp}\Gamma\mathring{\theta}\cdot\nabla\Gamma^{-\frac{1}{2}}\varphi
=tφ+Γ12x2f(t,x2)x1Γ12φΓ12x2Γf(t,x2)x1Γ12φ\displaystyle=\partial_{t}\varphi+\Gamma^{\frac{1}{2}}\partial_{x_{2}}f(t,x_{2})\partial_{x_{1}}\Gamma^{\frac{1}{2}}\varphi-\Gamma^{\frac{1}{2}}\partial_{x_{2}}\Gamma f(t,x_{2})\partial_{x_{1}}\Gamma^{-\frac{1}{2}}\varphi
=tφ+x2fx1Γφx2Γfx1φ\displaystyle=\partial_{t}\varphi+\partial_{x_{2}}f\partial_{x_{1}}\Gamma\varphi-\partial_{x_{2}}\Gamma f\partial_{x_{1}}\varphi
+[Γ12,x2f]x1Γ12φ[Γ12,x2Γf]x1Γ12φ.\displaystyle\mathrel{\phantom{=}}+[\Gamma^{\frac{1}{2}},\partial_{x_{2}}f]\partial_{x_{1}}\Gamma^{\frac{1}{2}}\varphi-[\Gamma^{\frac{1}{2}},\partial_{x_{2}}\Gamma f]\partial_{x_{1}}\Gamma^{-\frac{1}{2}}\varphi.

For the last line, we factor out x1\partial_{x_{1}} (whose symbol is iξ1i\xi_{1}) and use Lemma 3.1 to write

[Γ12,x2f]Γ12(i1{γ12,x2f}γ12(ξ))(x,D)Op(S(ξ2m)),\displaystyle[\Gamma^{\frac{1}{2}},\partial_{x_{2}}f]\Gamma^{\frac{1}{2}}-\left(i^{-1}\{\gamma^{\frac{1}{2}},\partial_{x_{2}}f\}\gamma^{\frac{1}{2}}(\xi)\right)(x,D)\in Op(S(\langle{\xi}\rangle^{-2}m)),
[Γ12,x2Γf]Γ12(i1{γ12,x2f}γ12(ξ))(x,D)Op(S(ξ2)).\displaystyle[\Gamma^{\frac{1}{2}},\partial_{x_{2}}\Gamma f]\Gamma^{-\frac{1}{2}}-\left(i^{-1}\{\gamma^{\frac{1}{2}},\partial_{x_{2}}f\}\gamma^{-\frac{1}{2}}(\xi)\right)(x,D)\in Op(S(\langle{\xi}\rangle^{-2})).

Then using the identity

{γ12,g}=12γ12ξ2γ(ξ)x2g\{\gamma^{\frac{1}{2}},g\}=\frac{1}{2}\gamma^{-\frac{1}{2}}\partial_{\xi_{2}}\gamma(\xi)\partial_{x_{2}}g

which holds for any function g=g(x2)g=g(x_{2}), it follows that

θ̊φ=tφ+ipθ̊(x2,D1,D2)φ+sθ̊(x2,D1,D2)φ+rθ̊(x2,D1,D2)φ,\mathcal{L}_{\mathring{\theta}}\varphi=\partial_{t}\varphi+ip_{\mathring{\theta}}(x_{2},D_{1},D_{2})\varphi+s_{\mathring{\theta}}(x_{2},D_{1},D_{2})\varphi+r_{\mathring{\theta}}(x_{2},D_{1},D_{2})\varphi, (3.14)

where

pθ̊(x2,ξ1,ξ2)\displaystyle p_{\mathring{\theta}}(x_{2},\xi_{1},\xi_{2}) =x2f(t,x2)ξ1γ(ξ1,ξ2)(x2Γf)(t,x2)ξ1,\displaystyle=\partial_{x_{2}}f(t,x_{2})\xi_{1}\gamma(\xi_{1},\xi_{2}){-}(\partial_{x_{2}}\Gamma f)(t,x_{2})\xi_{1},
sθ̊(x2,ξ1,ξ2)\displaystyle s_{\mathring{\theta}}(x_{2},\xi_{1},\xi_{2}) =+12x22f(t,x2)ξ2γ(ξ1,ξ2)ξ112(x22Γf)(t,x2)γ1(ξ)ξ2γ(ξ1,ξ2)ξ1,\displaystyle={+}\frac{1}{2}\partial_{x_{2}}^{2}f(t,x_{2})\partial_{\xi_{2}}\gamma(\xi_{1},\xi_{2})\xi_{1}-\frac{1}{2}(\partial_{x_{2}}^{2}\Gamma f)(t,x_{2})\gamma^{-1}(\xi)\partial_{\xi_{2}}\gamma(\xi_{1},\xi_{2})\xi_{1},
rθ̊(x2,ξ1ξ2)\displaystyle r_{\mathring{\theta}}(x_{2},\xi_{1}\xi_{2}) S(λ01+λ02+ξ22γ(λ0,ξ2)).\displaystyle\in S\left({\tfrac{\lambda_{0}}{1+\lambda_{0}^{2}+\xi_{2}^{2}}\gamma(\lambda_{0},\xi_{2})}\right).

Next, we conjugate (3.14) by eiλ0x1e^{i\lambda_{0}x_{1}}. For any symbol pp, note that the effect of conjugating by the linear phase λ0x1\lambda_{0}x_{1} is simply the translation ξ1ξ1+λ0\xi_{1}\mapsto\xi_{1}+\lambda_{0}; i.e., (λ0x1)p(x2,ξ1,ξ2)=p(x2,λ0+ξ1,ξ2){}^{(\lambda_{0}x_{1})}p(x_{2},\xi_{1},\xi_{2})=p(x_{2},\lambda_{0}+\xi_{1},\xi_{2}), where the LHS is given in Lemma 3.2. Moreover, it is clear from (3.2) that for any function ψ\psi that is independent of x1x_{1}, we have p(x2,λ0+D1,D2)ψ=p(x2,λ0,D2)ψp(x_{2},\lambda_{0}+D_{1},D_{2})\psi=p(x_{2},\lambda_{0},D_{2})\psi. Since pθ̊(x2,λ0,ξ2)=pθ̊,λ0(x2,ξ2)p_{\mathring{\theta}}(x_{2},\lambda_{0},\xi_{2})=p_{\mathring{\theta},\lambda_{0}}(x_{2},\xi_{2}), sθ̊(x2,λ0,ξ2)=sθ̊,λ0(x2,ξ2)s_{\mathring{\theta}}(x_{2},\lambda_{0},\xi_{2})=s_{\mathring{\theta},\lambda_{0}}(x_{2},\xi_{2}) and rθ̊(x2,λ0,ξ2)=rθ̊,λ0(x2,ξ2)r_{\mathring{\theta}}(x_{2},\lambda_{0},\xi_{2})=r_{\mathring{\theta},\lambda_{0}}(x_{2},\xi_{2}), the proof of Proposition 3.4 is complete. ∎

With Proposition 3.4 in hand, we are ready to specify the equations solved by the phase function 𝚽=λ0x1+Φ(t,x2)\mathbf{\Phi}=\lambda_{0}x_{1}+\Phi(t,x_{2}) and the amplitude function a=a(t,x2)a=a(t,x_{2}), which would constitute the desired degenerating wave packet φ~=Re(ei𝚽a)=Re(ei(λ0x1+Φ)a)\widetilde{\varphi}=\operatorname{Re}(e^{i\mathbf{\Phi}}a)=\operatorname{Re}(e^{i(\lambda_{0}x_{1}+\Phi)}a). The scheme itself is simply the standard WKB procedure for the nonlocal operator θ̊\mathcal{L}_{\mathring{\theta}}, but there are extra twists that arise in our scenario due to the degeneracy of θ̊\mathring{\theta}; see Remark 3.5 below.

Equation for the phase Φ\Phi. We choose the phase function Φ=Φ(t,x2)\Phi=\Phi(t,x_{2}) to be a solution to the Hamilton–Jacobi equation

tΦ+pθ̊,λ0(t,x2,x2Φ)=0,\partial_{t}\Phi+p_{\mathring{\theta},\lambda_{0}}(t,x_{2},\partial_{x_{2}}\Phi)=0, (3.15)

with initial data satisfying |x2Φ(0,x2)|λ0|{\partial_{x_{2}}\Phi(0,x_{2})}|\simeq\lambda_{0} (the precise choice of the initial data has to be well-adapted to our problem; see Section 4 below). We note that the (X,Ξ)(X,\Xi)-components of the characteristics for (3.15) solve the Hamiltonian ODE associated with pθ̊,λ0(t,x2,ξ2)p_{\mathring{\theta},\lambda_{0}}(t,x_{2},\xi_{2}),

{X˙2(t)=ξ2pθ̊,λ0(t,X2(t),Ξ2(t)),Ξ˙2(t)=x2pθ̊,λ0(t,X2(t),Ξ2(t)).\left\{\begin{aligned} \dot{X}_{2}(t)&=\partial_{\xi_{2}}p_{\mathring{\theta},\lambda_{0}}(t,X_{2}(t),\Xi_{2}(t)),\\ \dot{\Xi}_{2}(t)&=-\partial_{x_{2}}p_{\mathring{\theta},\lambda_{0}}(t,X_{2}(t),\Xi_{2}(t)).\end{aligned}\right.

Equation for the amplitude aa. We choose the amplitude function a=a(t,x2)a=a(t,x_{2}) to obey the transport equation

~ta:=ta+ξ2pθ̊,λ0(x2,x2Φ)x2a+(12ξ22pθ̊,λ0(x2,x2Φ)x22Φ+sθ̊,λ0(x2,x2Φ))a=0,\widetilde{\nabla}_{t}a:=\partial_{t}a+\partial_{\xi_{2}}p_{\mathring{\theta},\lambda_{0}}(x_{2},\partial_{x_{2}}\Phi)\partial_{x_{2}}a+\left(\frac{1}{2}\partial_{\xi_{2}}^{2}p_{\mathring{\theta},\lambda_{0}}(x_{2},\partial_{x_{2}}\Phi)\partial_{x_{2}}^{2}\Phi+s_{\mathring{\theta},\lambda_{0}}(x_{2},\partial_{x_{2}}\Phi)\right)a=0, (3.16)

with smooth and compactly supported initial data. At this point, we note that if one takes the inner product of (3.16) with aa and integrate by parts, then the specific structure of sθ̊,λ0s_{\mathring{\theta},\lambda_{0}} will show that the L2L^{2} norm of aa stays bounded on an O(1)O(1) timescale.

Key computation for the error term. The reason for the choices of the equations (3.15) and (3.16) is the following computation:

t(aeiΦ)+ipθ̊,λ0(aeiΦ)+sθ̊,λ0(aeiΦ)=(Φ)rpθ̊,λ0,2+(Φ)rsθ̊,λ0,1.\partial_{t}(ae^{i\Phi})+ip_{\mathring{\theta},\lambda_{0}}(ae^{i\Phi})+s_{\mathring{\theta},\lambda_{0}}(ae^{i\Phi})={}^{(\Phi)}r_{p_{\mathring{\theta},\lambda_{0}},-2}+{}^{(\Phi)}r_{s_{\mathring{\theta},\lambda_{0}},-1}. (3.17)

Indeed, by the equations for tΦ\partial_{t}\Phi and ta\partial_{t}a:

t(aeiΦ)+ipθ̊,λ0(aeiΦ)+sθ̊,λ0(aeiΦ)=itΦaeiΦ+(ta)eiΦ+ipθ̊,λ0(aeiΦ)+sθ̊,λ0(aeiΦ)=ipθ̊,λ0(aeiΦ)ipθ̊,λ0(x,xΦ)aeiΦξ2pθ̊,λ0(x2,x2Φ)x2aeiΦ12ξ22pθ̊,λ0(x2,x2Φ)x22ΦaeiΦ+sθ̊,λ0(aeiΦ)sθ̊,λ0(x,xΦ)aeiΦ.\begin{split}\partial_{t}(ae^{i\Phi})+ip_{\mathring{\theta},\lambda_{0}}(ae^{i\Phi})+s_{\mathring{\theta},\lambda_{0}}(ae^{i\Phi})&=i\partial_{t}\Phi ae^{i\Phi}+(\partial_{t}a)e^{i\Phi}+ip_{\mathring{\theta},\lambda_{0}}(ae^{i\Phi})+s_{\mathring{\theta},\lambda_{0}}(ae^{i\Phi})\\ &=ip_{\mathring{\theta},\lambda_{0}}(ae^{i\Phi})-ip_{\mathring{\theta},\lambda_{0}}(x,\partial_{x}\Phi)ae^{i\Phi}-\partial_{\xi_{2}}p_{\mathring{\theta},\lambda_{0}}(x_{2},\partial_{x_{2}}\Phi)\partial_{x_{2}}ae^{i\Phi}\\ &\quad-\frac{1}{2}\partial_{\xi_{2}}^{2}p_{\mathring{\theta},\lambda_{0}}(x_{2},\partial_{x_{2}}\Phi)\partial_{x_{2}}^{2}\Phi\,ae^{i\Phi}\\ &\quad+s_{\mathring{\theta},\lambda_{0}}(ae^{i\Phi})-s_{\mathring{\theta},\lambda_{0}}(x,\partial_{x}\Phi)ae^{i\Phi}.\end{split}

Then, (3.17) follows from Proposition 3.3.

In conclusion, to justify that φ~=Re(aei(λ0x1+Φ))\widetilde{\varphi}=\operatorname{Re}(ae^{i(\lambda_{0}x_{1}+\Phi)}) is a good approximate solution, we need to use (3.13) and (3.17) to show that θ̊φ~=Re(θ̊(aei(λ0x1+Φ)))\mathcal{L}_{\mathring{\theta}}\widetilde{\varphi}=\operatorname{Re}(\mathcal{L}_{\mathring{\theta}}(ae^{i(\lambda_{0}x_{1}+\Phi)})) is suitably small. As remarked at the beginning of this section, this goal shall be achieved in Section 6.

Remark 3.5.

In our case, due to the degeneracy of pθp_{\theta}, we expect x21+kΦ\partial_{x_{2}}^{1+k}\Phi and x2ka\partial_{x_{2}}^{k}a to grow rapidly in time for k1k\geq 1 (and indeed, this is precisely the behavior we wish to capture in order to show illposedness). For this reason, we cannot simply rely on standard pseudo-differential calculus to construct a wave packet that is valid up to a timescale where such a growth of higher derivatives (i.e., degeneration) occurs. Nevertheless, as we shall see in Section 4, the growth rate of x2kΦ\partial_{x_{2}}^{k}\Phi and x2ka\partial_{x_{2}}^{k}a will be smaller than |x2Φ|k|{\partial_{x_{2}}\Phi}|^{k}, which we may salvage via Proposition 5.1 in Section 5.

4 The Hamilton–Jacobi equation and the associated transport operator

This section is devoted to the choice and analysis of the phase and amplitude functions. The main results of this section are Proposition 4.10 and 4.12, which provide sharp estimates on the derivatives of the phase and amplitude, respectively.

Let us outline the structure of this section. After detailing the choice of parameters in Section 4.1, we consider the case of time-independent background state in Section 4.2. In this case, we take our phase function to be a solution of the Hamilton–Jacobi equation in separation of variables form. Thanks to the rather explicit form of the phase function, it is straightforward to obtain pointwise estimates for the derivatives of the phase function (Proposition 4.5), and the resulting bounds serve as a guide for the time-dependent background case which is handled in Section 4.3. In this latter case, we use the same initial data for the phase function, and estimate the solution by the method of characteristics and a bootstrapping argument. The key variable hh is introduced in this section in (4.33), which encodes some cancellations in the Hamilton–Jacobi equation and allows us to obtain sharp estimates on the second derivative of the phase function (Lemma 4.8 and Proposition 4.9). Using the variable hh, we prove Proposition 4.10 in Section 4.4, which gives the desired pointwise bounds on the derivatives of the phase function. Then, it is not very difficult to obtain estimates on the amplitude function, which is Proposition 4.12 of Section 4.5.

4.1 Initial reductions and conventions

Throughout this section, we study the Hamilton–Jacobi equation (3.15) with the following conventions:

  • Assume that x̊2=0\mathring{x}_{2}=0 and x22f(0,0)<0\partial_{x_{2}}^{2}f(0,0)<0.

  • Write x=x2x=x_{2} and ξ=ξ2\xi=\xi_{2}. Moreover, a prime () denotes x2\partial_{x_{2}}.

  • Introduce the shorthand γλ0(ξ)=γ(λ0,ξ)\gamma_{\lambda_{0}}(\xi)=\gamma(\lambda_{0},\xi).

With the above reductions, the Hamilton–Jacobi equation is simply given by

tΦ+p(t,x,xΦ)=0 in t×x,\partial_{t}\Phi+p(t,x,\partial_{x}\Phi)=0\qquad\hbox{ in }\mathbb{R}_{t}\times\mathbb{R}_{x}, (4.1)

where

p(t,x,ξ)=f(t,x)λ0γλ0(ξ)+Γf(t,x)λ0.p(t,x,\xi)=f^{\prime}(t,x)\lambda_{0}\gamma_{\lambda_{0}}(\xi)+\Gamma f^{\prime}(t,x)\lambda_{0}. (4.2)

The key point of our analysis is that f(t,x)f^{\prime}(t,x) is assumed to vanish linearly at some point for each tt. As a result, there exist bicharacteristics (X(t),Ξ(t))(X(t),\Xi(t)) associated with p(t,x,ξ)p(t,x,\xi) that exhibits rapid growth of |Ξ||{\Xi}|. Our aim is to construct a solution Φ\Phi to (4.1), which we refer to as a phase function, that exhibits such a behavior for a sufficiently long time.

Parameters in the construction. In this section, we are given a symbol γ\gamma and a function ff that satisfy the assumptions laid out in Section 1.5. We are also given λ0\lambda_{0}\in\mathbb{Z}, M1M\geq 1, δ0>0\delta_{0}>0 and σ00\sigma_{0}\geq 0 that satisfy (1.18)–(1.20)888To be absolutely precise, instead of (1.19) we only need τM1\tau_{M}\leq 1 in this section. The full condition is needed in Section 6 below.. When ff is time-dependent, we assume furthermore that ff is even and that we are given δ1>0\delta_{1}>0 such that (1.25) holds.

We fix a nonincreasing function ϵ(λ0)\epsilon(\lambda_{0}) for λ0Ξ0\lambda_{0}\geq\Xi_{0} such that, for λ0Λ\lambda_{0}\geq\Lambda,

max{1λ0σ0+16δ0,1γλ0(λ0)112δ0}ϵ(λ0)min{1100,1(10+22β0)supM[1,M]γλ0(Mλ0)MτM}.\max\left\{\frac{1}{\lambda_{0}^{\sigma_{0}+\frac{1}{6}\delta_{0}}},\,\frac{1}{\gamma_{\lambda_{0}}(\lambda_{0})^{1-\frac{1}{2}\delta_{0}}}\right\}\leq\epsilon(\lambda_{0})\leq\min\left\{\frac{1}{100},\frac{1}{(10+2^{2\beta_{0}})\sup_{M^{\prime}\in[1,M]}\frac{\gamma_{\lambda_{0}}(M^{\prime}\lambda_{0})}{M^{\prime}}\tau_{M^{\prime}}}\right\}. (4.3)

By the second nondissipative growth condition (1.18), we see that there always exists such a function, i.e., if ϵ(λ0)\epsilon(\lambda_{0}) satisfies

max{1λ0σ0+16δ0,1γλ0(λ0)112δ0}ϵ(λ0)12β0+1(2β0+3)min{γλ0(λ0)1δ0,λ0σ0},\max\left\{\frac{1}{\lambda_{0}^{\sigma_{0}+\frac{1}{6}\delta_{0}}},\,\frac{1}{\gamma_{\lambda_{0}}(\lambda_{0})^{1-\frac{1}{2}\delta_{0}}}\right\}\leq\epsilon(\lambda_{0})\leq\frac{1}{2^{\beta_{0}+1}(2^{\beta_{0}}+3)\min\{\gamma_{\lambda_{0}}(\lambda_{0})^{1-\delta_{0}},\lambda_{0}^{\sigma_{0}}\}},

for λ0Ξ0\lambda_{0}\geq\Xi_{0}, then it would obey (4.3).

The primary role of ϵ(λ0)\epsilon(\lambda_{0}) is as the relative physical localization scale of the wave packet: the support of a(t=0,x)a(t=0,x) will be contained in an interval of size Δx0\Delta x_{0} that is comparable to x0ϵ(λ0)x_{0}\epsilon(\lambda_{0}) up to a logarithmic power of λ0\lambda_{0}, where x0x_{0} is the initial location of the wave packet. For technical simplicity, we shall also choose (see (4.11) and (4.27) below)

x0=cx0ϵ(λ0),x_{0}=c_{x_{0}}\epsilon(\lambda_{0}),

where we shall retain the freedom to choose cx0>0c_{x_{0}}>0 throughout this section (it will only be fixed in Section 6.1).

Remark 4.1.

The restriction σ013(12δ0)\sigma_{0}\leq\frac{1}{3}(1-2\delta_{0}) in Theorems A and B arise from the uncertainty principle at t=0t=0 (which requires Δx0λ01\Delta x_{0}\lambda_{0}\gtrsim 1) and the wave packet error estimate; see Proposition 6.3 below. We note that the requirement x0ϵ(λ0)x_{0}\simeq\epsilon(\lambda_{0}) can be weakened if f(0,0)=0f^{\prime\prime\prime}(0,0)=0, in which case the restriction σ013(12δ0)\sigma_{0}\leq\frac{1}{3}(1-2\delta_{0}) can be relaxed; see Remark 4.2 below.

In addition to δ0\delta_{0} and δ1\delta_{1}, we are given δ2\delta_{2} and N01N_{0}^{-1}, which will be chosen in Sections 6 and 7 (in fact, δ2=δ010\delta_{2}=\delta_{0}^{10} and N0=104max{δ21(1+β0),1+α0,1+s,1+s}N_{0}=10^{4}\max\{\delta_{2}^{-1}(1+\beta_{0}),1+\alpha_{0},{1+s,1+s^{\prime}}\}; see Proposition 6.3 and Section 7.2 below). In the course of this section, we will choose additional small parameters in the following order (each choice may also depend on γ\gamma and ff):

δ0,δ1,δ2,N01δ3δ4δ5.\delta_{0},\delta_{1},\delta_{2},N_{0}^{-1}\to\delta_{3}\to\delta_{4}\to\delta_{5}.

In this section, we also require that

τMT,λ0Λ\tau_{M}\leq T,\qquad\lambda_{0}\geq\Lambda

where we shall retain the freedom to choose 0<T10<T\leq 1 and Λ1\Lambda\geq 1 throughout this section (TT may be fixed at the end of this section, while Λ\Lambda will be fixed in Section 6.2).

4.2 The case of a shear steady state background

Choice of Φ\Phi. When θ̊=f(x2)\mathring{\theta}=f(x_{2}) is time-independent, i.e., tf=0\partial_{t}f^{\prime}=0, we may obtain a useful solution to the Hamilton–Jacobi equation by separation of variables: Consider the ansatz

Φ(t,x)=Eλ0t+G(x).\Phi(t,x)=E\lambda_{0}t+G(x).

Then G(x)G(x) obeys the equation

p(x,xG)=Eλ0.p(x,\partial_{x}G)=-E\lambda_{0}. (4.4)

By the positivity of ξξγλ0\xi\partial_{\xi}\gamma_{\lambda_{0}}, γλ0()\gamma_{\lambda_{0}}(\cdot) is invertible on (,λ0](-\infty,-\lambda_{0}] or [λ0,)[\lambda_{0},\infty). In order to force xΦ=xG\partial_{x}\Phi=\partial_{x}G to grow (as tt increases) on characteristic curves, we choose xG\partial_{x}G to be positive (indeed, this sign relation may be read off from the bicharacteristic equations). In what follows, we denote by γλ01\gamma_{\lambda_{0}}^{-1} the inverse of the restriction γλ0:[λ0,)(0,)\gamma_{\lambda_{0}}:[\lambda_{0},\infty)\to(0,\infty). Hence, we obtain the explicit formula

xΦ=xG=γλ01(E+Γff).\partial_{x}\Phi=\partial_{x}G=\gamma_{\lambda_{0}}^{-1}\left(\frac{E+\Gamma f^{\prime}}{-f^{\prime}}\right). (4.5)

To begin with, we proceed to show that there is a choice of EE in (4.5) so that

λ0xΦ(0,x)(1+ϵ)λ0\lambda_{0}\leq\partial_{x}\Phi(0,x)\leq(1+\epsilon)\lambda_{0} (4.6)

holds for x0<x<x1x_{0}<x<x_{1} and λ0Λ\lambda_{0}\geq\Lambda, provided that x0x_{0} and x1x_{1} are sufficiently small (with respect to ff^{\prime}, λ0\lambda_{0}, and ϵ\epsilon) and Λ\Lambda is sufficiently large (depending on γ\gamma and |f(0)|1Γf|{f^{\prime\prime}(0)}|^{-1}\Gamma f^{\prime\prime}). Since γ1λ0\gamma^{-1}_{\lambda_{0}} is increasing, (4.6) is equivalent to having

γλ0(λ0)E+Γf(x)f(x)γλ0((1+ϵ(λ0))λ0),x0<x<x1.\begin{split}\gamma_{\lambda_{0}}(\lambda_{0})\leq\frac{E+\Gamma f^{\prime}(x)}{-f^{\prime}(x)}\leq\gamma_{\lambda_{0}}((1+\epsilon(\lambda_{0}))\lambda_{0}),\qquad x_{0}<x<x_{1}.\end{split} (4.7)

We first take x1>0x_{1}>0 to be sufficiently small (recall that f(0)<0f^{\prime\prime}(0)<0), so that

(1+ϵ10)f(0)<f(x)<(1ϵ10)f(0) for x(0,x1).(1+\tfrac{\epsilon}{10})f^{\prime\prime}(0)<f^{\prime\prime}(x)<(1-\tfrac{\epsilon}{10})f^{\prime\prime}(0)\qquad\hbox{ for $x\in(0,x_{1})$}. (4.8)

In particular, we have (1+ϵ10)f(0)x<f(x)<(1ϵ10)f(0)x(1+\tfrac{\epsilon}{10})f^{\prime\prime}(0)x<f^{\prime}(x)<(1-\tfrac{\epsilon}{10})f^{\prime\prime}(0)x for x(0,x1)x\in(0,x_{1}). Then, we define

E=f(x1)γλ0((1+12ϵ)λ0)Γf(x1).\begin{split}E=-f^{\prime}(x_{1})\gamma_{\lambda_{0}}((1+\tfrac{1}{2}\epsilon)\lambda_{0})-\Gamma f^{\prime}(x_{1}).\end{split}

With this choice of EE, (4.7) is trivially satisfied at x=x1x=x_{1}. We may now choose x0<x1x_{0}<x_{1} in a way that (4.7) is satisfied for all x(x0,x1)x\in(x_{0},x_{1}): we take

x1x01=110γλ0((1+34ϵ)λ0)γλ0((1+12ϵ)λ0)γλ0((1+12ϵ)λ0).\begin{split}\frac{x_{1}}{x_{0}}-1=\frac{1}{10}\frac{\gamma_{\lambda_{0}}((1+\tfrac{3}{4}\epsilon)\lambda_{0})-\gamma_{\lambda_{0}}((1+\tfrac{1}{2}\epsilon)\lambda_{0})}{\gamma_{\lambda_{0}}((1+\tfrac{1}{2}\epsilon)\lambda_{0})}.\end{split} (4.9)

With these choices, we estimate

f(x1)f(x0)=1+x0x1f(x)dx0x0f(x)dx1+10|1x1x0|=γλ0((1+34ϵ)λ0)γλ0((1+12ϵ)λ0).\begin{split}\frac{f^{\prime}(x_{1})}{f^{\prime}(x_{0})}=1+\frac{\int_{x_{0}}^{x_{1}}f^{\prime\prime}(x)\,\mathrm{d}x}{\int_{0}^{x_{0}}f^{\prime\prime}(x)\,\mathrm{d}x}\leq 1+10\left|1-\frac{x_{1}}{x_{0}}\right|=\frac{\gamma_{\lambda_{0}}((1+\tfrac{3}{4}\epsilon)\lambda_{0})}{\gamma_{\lambda_{0}}((1+\tfrac{1}{2}\epsilon)\lambda_{0})}.\end{split}

where we take λ0Λ\lambda_{0}\geq\Lambda larger if necessary to decrease ϵ\epsilon. Then,

E+Γf(x)f(x)=f(x1)γλ0((1+12ϵ)λ0)Γf(x1)+Γf(x)f(x)f(x1)f(x0)γλ0((1+12ϵ)λ0)+2f(0)ΓfL(x0,x1)|x0x1|x0γλ0((1+34ϵ)λ0)+20ΓfL(x0,x1)|f(0)|γλ0((1+34ϵ)λ0)γλ0((1+12ϵ)λ0)γλ0((1+12ϵ)λ0).\begin{split}\frac{E+\Gamma f^{\prime}(x)}{-f^{\prime}(x)}&=\frac{-f^{\prime}(x_{1})\gamma_{\lambda_{0}}((1+\tfrac{1}{2}\epsilon)\lambda_{0})-\Gamma f^{\prime}(x_{1})+\Gamma f^{\prime}(x)}{-f^{\prime}(x)}\\ &\leq\frac{f^{\prime}(x_{1})}{f^{\prime}(x_{0})}\gamma_{\lambda_{0}}((1+\tfrac{1}{2}\epsilon)\lambda_{0})+\frac{2}{f^{\prime\prime}(0)}\|{\Gamma f^{\prime\prime}}\|_{L^{\infty}(x_{0},x_{1})}\frac{|x_{0}-x_{1}|}{x_{0}}\\ &\leq\gamma_{\lambda_{0}}((1+\tfrac{3}{4}\epsilon)\lambda_{0})+\frac{20\|{\Gamma f^{\prime\prime}}\|_{L^{\infty}(x_{0},x_{1})}}{|{f^{\prime\prime}(0)}|}\frac{\gamma_{\lambda_{0}}((1+\tfrac{3}{4}\epsilon)\lambda_{0})-\gamma_{\lambda_{0}}((1+\tfrac{1}{2}\epsilon)\lambda_{0})}{\gamma_{\lambda_{0}}((1+\tfrac{1}{2}\epsilon)\lambda_{0})}.\end{split}

To estimate the last line by γλ0((1+ϵ)λ0)\gamma_{\lambda_{0}}((1+\epsilon)\lambda_{0}), it suffices to verify

20ΓfL(x0,x1)|f(0)|γλ0((1+34ϵ)λ0)γλ0((1+12ϵ)λ0)γλ0((1+12ϵ)λ0)γλ0((1+ϵ)λ0)γλ0((1+34ϵ)λ0).\begin{split}\frac{20\|{\Gamma f^{\prime\prime}}\|_{L^{\infty}(x_{0},x_{1})}}{|{f^{\prime\prime}(0)}|}\frac{\gamma_{\lambda_{0}}((1+\tfrac{3}{4}\epsilon)\lambda_{0})-\gamma_{\lambda_{0}}((1+\tfrac{1}{2}\epsilon)\lambda_{0})}{\gamma_{\lambda_{0}}((1+\tfrac{1}{2}\epsilon)\lambda_{0})}\leq\gamma_{\lambda_{0}}((1+\epsilon)\lambda_{0})-\gamma_{\lambda_{0}}((1+\tfrac{3}{4}\epsilon)\lambda_{0}).\end{split} (4.10)

From our assumptions on γ\gamma, this is guaranteed for λ0Λ0\lambda_{0}\geq\Lambda_{0} with sufficiently large Λ0\Lambda_{0}. Indeed, by the mean value theorem, there exist α1,α2[0,1]\alpha_{1},\alpha_{2}\in[0,1] and α3[2+α14,3+α24]\alpha_{3}\in[\tfrac{2+\alpha_{1}}{4},\tfrac{3+\alpha_{2}}{4}] such that

γλ0((1+34ϵ)λ0)γλ0((1+12ϵ)λ0)γλ0((1+ϵ)λ0)γλ0((1+34ϵ)λ0)=1+ξγλ0((1+2+α14ϵ)λ0)ξγλ0((1+3+α24ϵ)λ0)ξγλ0((1+3+α24ϵ)λ0)1+ϵλ0|ξ2γλ0((1+α3ϵ)λ0)|ξγλ0((1+3+α24ϵ)λ0)C,\begin{split}\frac{\gamma_{\lambda_{0}}((1+\tfrac{3}{4}\epsilon)\lambda_{0})-\gamma_{\lambda_{0}}((1+\tfrac{1}{2}\epsilon)\lambda_{0})}{\gamma_{\lambda_{0}}((1+\epsilon)\lambda_{0})-\gamma_{\lambda_{0}}((1+\tfrac{3}{4}\epsilon)\lambda_{0})}&=1+\frac{\partial_{\xi}\gamma_{\lambda_{0}}((1+\tfrac{2+\alpha_{1}}{4}\epsilon)\lambda_{0})-\partial_{\xi}\gamma_{\lambda_{0}}((1+\tfrac{3+\alpha_{2}}{4}\epsilon)\lambda_{0})}{\partial_{\xi}\gamma_{\lambda_{0}}((1+\tfrac{3+\alpha_{2}}{4}\epsilon)\lambda_{0})}\\ &\leq 1+\epsilon\lambda_{0}\frac{|{\partial_{\xi}^{2}\gamma_{\lambda_{0}}((1+\alpha_{3}\epsilon)\lambda_{0})}|}{\partial_{\xi}\gamma_{\lambda_{0}}((1+\tfrac{3+\alpha_{2}}{4}\epsilon)\lambda_{0})}\leq C,\end{split}

where we have used the ellipticity of ξξγλ0\xi\partial_{\xi}\gamma_{\lambda_{0}} for the last inequality. Since γλ0\gamma_{\lambda_{0}}\to\infty as λ0\lambda_{0}\to\infty, the desired bound (4.10) follows by taking λ0Λ\lambda_{0}\geq\Lambda large enough. We have verified the right inequality of (4.7). The proof of the left inequality is only easier since f(x)-f^{\prime}(x) is increasing in (x0,x1)(x_{0},x_{1}). We hence omit the proof.

We have shown that if x1x_{1} and x0x_{0} are chosen so that (4.8) and (4.9) and Λ\Lambda is sufficiently large depending on γ\gamma and |f(0)|1Γf|{f^{\prime\prime}(0)}|^{-1}\Gamma f^{\prime\prime}, then EE can be chosen so that (4.6) holds. In what follows, we choose x0x_{0}, x1x_{1} to be of the form

x0=cx0ϵ(λ0),x1=x0+Δx0, where ccx0(logλ0)2ϵ(λ0)|Δx0x0|cx0ϵ(λ0),\begin{split}x_{0}={c_{x_{0}}\epsilon(\lambda_{0})},\quad{x_{1}=x_{0}+\Delta x_{0}},\quad\hbox{ where }{\frac{cc_{x_{0}}}{(\log\lambda_{0})^{2}}\epsilon(\lambda_{0})}\leq\left|\frac{\Delta x_{0}}{x_{0}}\right|\leq c_{x_{0}}\epsilon(\lambda_{0}),\end{split} (4.11)

where cc is a constant depending only on γ\gamma and ff, and cx1>0c_{x_{1}}>0 obeys

f(x)L(0,cx1ϵ(λ0))cx0120|f(0)|.\|{f^{\prime\prime\prime}(x)}\|_{L^{\infty}(0,c_{x_{1}}\epsilon(\lambda_{0}))}c_{x_{0}}\leq\frac{1}{20}|{f^{\prime\prime}(0)}|. (4.12)

Clearly, such a choice of x0x_{0} and x1x_{1} ensures (4.8). In the following, there are several occasions in which we need to take x0x_{0} and x1x_{1} sufficiently small. To achieve this, we shall retain the freedom to shrink cx0>0c_{x_{0}}>0 until the end of this section; note that this action only makes the LHS of (4.12) smaller. The second inequality in (4.11) follows from (4.9), mean value theorem, and the almost comparability of ξξγλ0\xi\partial_{\xi}\gamma_{\lambda_{0}} and γλ0\gamma_{\lambda_{0}}.

Remark 4.2.

If ff^{\prime\prime\prime} vanishes at 0, then we may choose x0x_{0} be be larger. More precisely,

x0=cx0ϵ(λ0)1nf\begin{split}x_{0}={c_{x_{0}}\epsilon(\lambda_{0})^{\frac{1}{n_{f}}}}\end{split} (4.13)

would work, where nfn_{f} is the smallest natural number such that f(nf+2)(0)0f^{(n_{f}+2)}(0)\neq 0.

Bicharacteristics associated with Φ\Phi. We consider the bicharacteristic ODE

{X˙=f(X)λ0γλ0(Ξ),Ξ˙=f(X)λ0γλ0(Ξ)+(Γf)(X)λ0\left\{\begin{aligned} \dot{X}&=f^{\prime}(X)\lambda_{0}\gamma_{\lambda_{0}}^{\prime}(\Xi),\\ \dot{\Xi}&=-f^{\prime\prime}(X)\lambda_{0}\gamma_{\lambda_{0}}(\Xi)+(\Gamma f^{\prime\prime})(X)\lambda_{0}\end{aligned}\right. (4.14)

with initial data (X0,Ξ0)(X_{0},\Xi_{0}) satisfying x0<X0<x1x_{0}<X_{0}<x_{1} and λ0Ξ0=Φ(0,X0)(1+ϵ)λ0\lambda_{0}\leq\Xi_{0}=\partial\Phi(0,X_{0})\leq(1+\epsilon)\lambda_{0}.

Control of Ξ(t)\Xi(t). To obtain bounds on Ξ(t)\Xi(t), we introduce λ¯(t)\overline{\lambda}(t) and λ(t)\lambda(t), which are solutions to the ODEs

{λ¯˙(t)=(1+ϵ)f(0)λ0γλ0(λ¯),λ¯(0)=(1+ϵ)λ0,\left\{\begin{aligned} \dot{\overline{\lambda}}(t)&=-(1+\epsilon)f^{\prime\prime}(0)\lambda_{0}\gamma_{\lambda_{0}}(\overline{\lambda}),\\ \overline{\lambda}(0)&={(1+\epsilon)}\lambda_{0},\end{aligned}\right. (4.15)

and

{λ˙(t)=(1ϵ)f(0)λ0γλ0(λ),λ(0)=λ0,\left\{\begin{aligned} \dot{\lambda}(t)&=-(1-\epsilon)f^{\prime\prime}(0)\lambda_{0}\gamma_{\lambda_{0}}(\lambda),\\ \lambda(0)&={\lambda_{0}},\end{aligned}\right. (4.16)

respectively. We remind the reader that ϵ=ϵ(λ0)\epsilon=\epsilon(\lambda_{0}) obeys (4.3). We now claim that

λ(t)Ξ(t)λ¯(t).\lambda(t)\leq\Xi(t)\leq\overline{\lambda}(t). (4.17)

To begin with, observe that X˙=0\dot{X}=0 when X=0X=0, from which we see that no solution (X(t),Ξ(t))(X(t),\Xi(t)) with X(0)>0X(0)>0 can traverse to the region {X<0}\{X<0\}. Note furthermore that in the region {X>0}\{X>0\}, we have X˙<0\dot{X}<0. In conclusion, we see that 0<X(t)<X(0)<x10<X(t)<X(0)<x_{1} for any t>0t>0.

Next, note that

(1ϵ)f(0)λ0γλ0(λ)f(X)λ0γλ0(Ξ)+(Γf)(X)λ0(1+ϵ)f(0)λ0γλ0(λ¯)\begin{split}-(1-\epsilon)f^{\prime\prime}(0)\lambda_{0}\gamma_{\lambda_{0}}(\lambda)\leq-f^{\prime\prime}(X)\lambda_{0}\gamma_{\lambda_{0}}(\Xi)+(\Gamma f^{\prime\prime})(X)\lambda_{0}\leq-(1+\epsilon)f^{\prime\prime}(0)\lambda_{0}\gamma_{\lambda_{0}}(\overline{\lambda})\end{split}

for any X(0,x1)X\in(0,x_{1}) and λ0λΞλ¯\lambda_{0}\leq\lambda\leq\Xi\leq\overline{\lambda}. Indeed, by the first inequality of (4.3), we have ϵγλ0(λ)\epsilon\gamma_{\lambda_{0}}(\lambda)\to\infty as λ0\lambda_{0}\to\infty. By taking Λ0>0\Lambda_{0}>0 larger if necessary (in a way depending only on |f(0)|1ΓfL(0,x1)|{f^{\prime\prime}(0)}|^{-1}\|{\Gamma f^{\prime\prime}}\|_{L^{\infty}(0,x_{1})}, γ\gamma and ϵ\epsilon), we may guarantee that the preceding inequalities hold. Then, comparing the equations for Ξ˙\dot{\Xi}, λ˙\dot{\lambda} and λ¯˙\dot{\overline{\lambda}}, we obtain (4.17), as desired.

We now prove the following lemma.

Lemma 4.3.

The following statements hold.

  1. 1.

    λ(t)=Mλ0\lambda(t)=M\lambda_{0} exactly at tM:=11ϵ|f(0)|1τMt_{M}:=\frac{1}{1-\epsilon}|{f^{\prime\prime}(0)}|^{-1}\tau_{M}, and λ0λ(t)Mλ0\lambda_{0}\leq\lambda(t)\leq M\lambda_{0} for 0ttM0\leq t\leq t_{M}.

  2. 2.

    As long as λ(t)Mλ0\lambda(t)\leq M\lambda_{0}, we have

    λ(t)Ξ(t)2λ(t).\lambda(t)\leq\Xi(t)\leq 2\lambda(t). (4.18)
Proof.

It will be convenient to introduce M¯(t)\overline{M}^{\prime}(t) and M(t)M^{\prime}(t): M¯:=λ¯/λ0\overline{M}^{\prime}:=\overline{\lambda}/\lambda_{0} and M:=λ/λ0M^{\prime}:=\lambda/\lambda_{0}. We have M¯(t)M(t)1\overline{M}^{\prime}(t)\geq M^{\prime}(t)\geq 1. we start with the identities

λ0Mλ01γλ0(λ)dλλ0=f(0)(1ϵ)t,(1+ϵ)λ0M¯λ01γλ0(λ)dλλ0=f(0)(1+ϵ)t,\int_{\lambda_{0}}^{M^{\prime}\lambda_{0}}\frac{1}{\gamma_{\lambda_{0}}(\lambda)}\frac{\mathrm{d}\lambda}{\lambda_{0}}=-f^{\prime\prime}(0)(1-\epsilon)t,\qquad\int_{(1+\epsilon)\lambda_{0}}^{\overline{M}^{\prime}\lambda_{0}}\frac{1}{\gamma_{\lambda_{0}}(\lambda)}\frac{\mathrm{d}\lambda}{\lambda_{0}}=-f^{\prime\prime}(0)(1+\epsilon)t, (4.19)

which follow from (4.16) and (4.15), respectively. Note that the LHS of the first identity is exactly τM\tau_{M^{\prime}}; from this observation, the first statement easily follows. To prove (4.18), noting that MM^{\prime} is strictly increasing in time, we shall split the proof into two time intervals, depending on whether M<2M^{\prime}<2 or M2M^{\prime}\geq 2.

Case 1. Assume that M(t)<2M^{\prime}(t)<2. We are going to prove that M¯(t)<2M(t)\overline{M}^{\prime}(t)<2M^{\prime}(t) holds in this time interval by a bootstrap argument. Towards a contradiction, we may assume that there exists some 0<T<tM0<T<t_{M}, such that M(t)<2,M¯(t)<2M(t)M^{\prime}(t)<2,\overline{M}^{\prime}(t)<2M^{\prime}(t) on [0,T)[0,T) and M¯(T)=2M(T)\overline{M}^{\prime}(T)=2M^{\prime}(T).

We now restrict tt to [0,T][0,T]. Since λ\lambda and λ¯\overline{\lambda} are strictly increasing functions of time, we may consider λ¯=λ¯(λ)\overline{\lambda}=\overline{\lambda}(\lambda) (with some abuse of notation) and obtain

dλ¯dλ=1+ϵ1ϵγ(λ(λ¯/λ))γ(λ)(1+3ϵ)(λ¯λ)C\begin{split}\frac{\mathrm{d}\overline{\lambda}}{\mathrm{d}\lambda}=\frac{1+\epsilon}{1-\epsilon}\frac{\gamma(\lambda\cdot(\overline{\lambda}/\lambda))}{\gamma(\lambda)}\leq(1+3\epsilon)\left(\frac{\overline{\lambda}}{\lambda}\right)^{C}\end{split}

where we may assume that C>1C>1 since 2λλ¯λ2\lambda\geq\overline{\lambda}\geq\lambda. Integrating in λ\lambda and recalling that λ¯(λ0)=(1+ϵ)λ0\overline{\lambda}(\lambda_{0})=(1+\epsilon)\lambda_{0}, we obtain

1λ¯C11((1+ϵ)λ0)C1(1+3ϵ)(1λC11λ0C1).\begin{split}\frac{1}{\overline{\lambda}^{C-1}}-\frac{1}{((1+\epsilon)\lambda_{0})^{C-1}}\geq(1+3\epsilon)\left(\frac{1}{\lambda^{C-1}}-\frac{1}{\lambda_{0}^{C-1}}\right).\end{split}

Applying this inequality at t=Tt=T, we have that λ¯=2λ\overline{\lambda}=2\lambda and obtain

((1+3ϵ)12C1)1λC1((1+3ϵ)1(1+ϵ)C1)1λ0C1=OC(ϵ)1λ0C1.\begin{split}\left((1+3\epsilon)-\frac{1}{2^{C-1}}\right)\frac{1}{\lambda^{C-1}}\leq\left((1+3\epsilon)-\frac{1}{(1+\epsilon)^{C-1}}\right)\frac{1}{\lambda_{0}^{C-1}}=O_{C}(\epsilon)\frac{1}{\lambda_{0}^{C-1}}.\end{split}

and we can take Λ0\Lambda_{0} larger so that (1+3ϵ)>12C1(1+3\epsilon)>\frac{1}{2^{C-1}}. This gives (λ/λ0)C11(\lambda/\lambda_{0})^{C-1}\gg 1, which is a contradiction to M(T)<2M^{\prime}(T)<2. This guarantees (4.18) in this case.

Case 2. When M(t)2M^{\prime}(t)\geq 2, we now combine both identities in (4.19) to obtain

λ0(1+ϵ)λ01γλ0(λ)dλλ0+2ϵ1+ϵ(1+ϵ)λ0Mλ01γλ0(λ)dλλ0=1ϵ1+ϵMλ0M¯λ01γλ0(λ)dλλ0.\begin{split}\int_{\lambda_{0}}^{(1+\epsilon)\lambda_{0}}\frac{1}{\gamma_{\lambda_{0}}(\lambda)}\frac{\mathrm{d}\lambda}{\lambda_{0}}+\frac{2\epsilon}{1+\epsilon}\int_{(1+\epsilon)\lambda_{0}}^{M^{\prime}\lambda_{0}}\frac{1}{\gamma_{\lambda_{0}}(\lambda)}\frac{\mathrm{d}\lambda}{\lambda_{0}}=\frac{1-\epsilon}{1+\epsilon}\int_{M^{\prime}\lambda_{0}}^{\overline{M}^{\prime}\lambda_{0}}\frac{1}{\gamma_{\lambda_{0}}(\lambda)}\frac{\mathrm{d}\lambda}{\lambda_{0}}.\end{split}

Note that the LHS is bounded by

λ0(1+ϵ)λ01γλ0(λ)dλλ0+2ϵ1+ϵ(1+ϵ)λ0Mλ01γλ0(λ)dλλ0ϵγλ0(λ0)+2ϵλ0Mλ01γλ0(λ)dλλ0(2β0+1)ϵλ02λ01γλ0(λ)dλλ0+2ϵλ0Mλ01γλ0(λ)dλλ0(2β0+3)ϵτM,\begin{split}&\int_{\lambda_{0}}^{(1+\epsilon)\lambda_{0}}\frac{1}{\gamma_{\lambda_{0}}(\lambda)}\frac{\mathrm{d}\lambda}{\lambda_{0}}+\frac{2\epsilon}{1+\epsilon}\int_{(1+\epsilon)\lambda_{0}}^{M^{\prime}\lambda_{0}}\frac{1}{\gamma_{\lambda_{0}}(\lambda)}\frac{\mathrm{d}\lambda}{\lambda_{0}}\\ &\leq\frac{\epsilon}{\gamma_{\lambda_{0}}(\lambda_{0})}+2\epsilon\int_{\lambda_{0}}^{M^{\prime}\lambda_{0}}\frac{1}{\gamma_{\lambda_{0}}(\lambda)}\frac{\mathrm{d}\lambda}{\lambda_{0}}\\ &\leq(2^{\beta_{0}}+1)\epsilon\int_{\lambda_{0}}^{2\lambda_{0}}\frac{1}{\gamma_{\lambda_{0}}(\lambda)}\frac{\mathrm{d}\lambda}{\lambda_{0}}+2\epsilon\int_{\lambda_{0}}^{M^{\prime}\lambda_{0}}\frac{1}{\gamma_{\lambda_{0}}(\lambda)}\frac{\mathrm{d}\lambda}{\lambda_{0}}\\ &\leq(2^{\beta_{0}}+3)\epsilon\tau_{M^{\prime}},\end{split}

where we used the quantitative slow-variance condition (1.11) for the second inequality. By the second inequality in (4.3) and the assumption that MMM^{\prime}\leq M, we have

2β0+1(2β0+3)ϵγλ0(Mλ0)MτM1.\begin{split}2^{\beta_{0}+1}(2^{\beta_{0}}+3)\epsilon\frac{\gamma_{\lambda_{0}}(M^{\prime}\lambda_{0})}{M^{\prime}}\tau_{M^{\prime}}\leq 1.\end{split}

As a result,

Mλ0M¯λ01γλ0(λ)dλλ02(2β0+3)ϵτM2β0Mγλ0(Mλ0)Mλ02Mλ01γλ0(λ)dλλ0,\begin{split}\int_{M^{\prime}\lambda_{0}}^{\overline{M}^{\prime}\lambda_{0}}\frac{1}{\gamma_{\lambda_{0}}(\lambda)}\frac{\mathrm{d}\lambda}{\lambda_{0}}\leq 2(2^{\beta_{0}}+3)\epsilon\tau_{M^{\prime}}\leq 2^{-\beta_{0}}\frac{M^{\prime}}{\gamma_{\lambda_{0}}(M^{\prime}\lambda_{0})}\leq\int_{M^{\prime}\lambda_{0}}^{2M^{\prime}\lambda_{0}}\frac{1}{\gamma_{\lambda_{0}}(\lambda)}\frac{\mathrm{d}\lambda}{\lambda_{0}},\end{split}

where we used (1.11) for the last inequality. This implies M¯2M\overline{M}^{\prime}\leq 2M^{\prime}. Then, by (4.17), (4.18) follows. ∎

Control of X(t)X(t). With Lemma 4.3 for Ξ(t)\Xi(t), we may obtain bounds on X(t)X(t) using the conservation of the Hamiltonian:

p(X(t),xΦ(X(t)))=p(X(0),xΦ(X(0))).p(X(t),\partial_{x}\Phi(X(t)))=p(X(0),\partial_{x}\Phi(X(0))).

More specifically, we have

f(X(t))=f(X(0))γλ0(Ξ(0))+Γf(X(0))Γf(X(t))γλ0(Ξ(t)).f^{\prime}(X(t))=\frac{f^{\prime}(X(0))\gamma_{\lambda_{0}}(\Xi(0))+\Gamma f^{\prime}(X(0))-\Gamma f^{\prime}(X(t))}{\gamma_{\lambda_{0}}(\Xi(t))}.

Recall that 0<X(t)<X(0)0<X(t)<X(0). Therefore, we have

|Γf(X(0))Γf(X(t))|ΓfL(0,x1)|X(0)X(t)|ΓfL(0,x1)X(0).|{\Gamma f^{\prime}(X(0))-\Gamma f^{\prime}(X(t))}|\leq\|{\Gamma f^{\prime\prime}}\|_{L^{\infty}(0,x_{1})}|{X(0)-X(t)}|\leq\|{\Gamma f^{\prime\prime}}\|_{L^{\infty}(0,x_{1})}X(0).

On the other hand, |f(X(0))|(1ϵ10)|f(0)|X(0)|{f^{\prime}(X(0))}|\geq(1-\frac{\epsilon}{10})|{f^{\prime\prime}(0)}|X(0). Therefore, choosing Λ\Lambda sufficiently large depending on γ\gamma and |f(0)|1Γf|{f^{\prime\prime}(0)}|^{-1}\Gamma f^{\prime\prime}, we obtain

12β0+1X(0)γλ0(λ0)γλ0(λ(t))X(t)X(0)γλ0(λ0)γλ0(λ(t)) for λ0Λ.\begin{split}\frac{1}{2^{\beta_{0}+1}}\frac{X(0)\gamma_{\lambda_{0}}(\lambda_{0})}{\gamma_{\lambda_{0}}(\lambda(t))}\leq X(t)\leq\frac{X(0)\gamma_{\lambda_{0}}(\lambda_{0})}{\gamma_{\lambda_{0}}(\lambda(t))}\quad\hbox{ for }\lambda_{0}\geq\Lambda.\end{split} (4.20)

Control of x2Φ(X(t))\partial_{x}^{2}\Phi(X(t)). As we will see, a fundamental quantity for controlling the geometry of nearby characteristic curves is x2Φ\partial_{x}^{2}\Phi. While this quantity may also be computed explicitly by differentiating (4.5), it is more convenient to obtain an implicit formula by differentiating (4.1). Indeed, from (4.1) we see that

xp(x,xΦ)+ξp(x,xΦ)x2Φ=0,\partial_{x}p(x,\partial_{x}\Phi)+\partial_{\xi}p(x,\partial_{x}\Phi)\partial_{x}^{2}\Phi=0,

so that

x2Φ=xp(x,xΦ)ξp(x,xΦ)=f(x)γλ0(xΦ)+Γf(x)f(x)γλ0(xΦ).\partial_{x}^{2}\Phi=-\frac{\partial_{x}p(x,\partial_{x}\Phi)}{\partial_{\xi}p(x,\partial_{x}\Phi)}=-\frac{f^{\prime\prime}(x)\gamma_{\lambda_{0}}(\partial_{x}\Phi)+\Gamma f^{\prime\prime}(x)}{f^{\prime}(x)\gamma_{\lambda_{0}}^{\prime}(\partial_{x}\Phi)}. (4.21)

By taking Λ0\Lambda_{0} larger if necessary (in a way depending only on γ\gamma and |f(0)|1Γf|{f^{\prime\prime}(0)}|^{-1}\Gamma f^{\prime\prime}), we obtain the bound

0x2Φ(X(t))Cxγλ0(xΦ(X(t)))ξγλ0(xΦ(X(t)))Cx0(γλ0(Ξ(t)))2γλ0(λ0)γλ0(Ξ(t)).0\leq-\partial_{x}^{2}\Phi(X(t))\leq\frac{C}{x}\frac{\gamma_{\lambda_{0}}(\partial_{x}\Phi(X(t)))}{\partial_{\xi}\gamma_{\lambda_{0}}(\partial_{x}\Phi(X(t)))}\leq\frac{C}{x_{0}}\frac{(\gamma_{\lambda_{0}}(\Xi(t)))^{2}}{\gamma_{\lambda_{0}}(\lambda_{0})\gamma_{\lambda_{0}}^{\prime}(\Xi(t))}.
Remark 4.4.

It is worth noting that x2Φ(X(t))-\partial_{x}^{2}\Phi(X(t)) remains finite as long as (X(t),Ξ(t))(X(t),\Xi(t)) exists. Geometrically, this reflects the fact that there are no focal points along each characteristic associated to our Φ\Phi constructed via separation of variables. This fact is not a-priori clear at the level of the ODE for x2Φ(X(t))\partial_{x}^{2}\Phi(X(t)), which is of Ricatti-type (see (4.32) below); hence, it should be seen as a benefit of our separation of variables approach. The above sharp bound for x2Φ(X(t))-\partial_{x}^{2}\Phi(X(t)) will serve as the basis for the sharp estimates for higher derivatives of Φ\Phi and the amplitude function aa (see (3.16)) along characteristics in Sections 4.44.5.

For the ensuing argument, we record separately the bound at the initial time t=0t=0:

0x2Φ(x)Cxγλ0(xΦ(x))ξγλ0(xΦ(x))Cx0λ0(logλ0)2.0\leq-\partial_{x}^{2}\Phi(x)\leq\frac{C}{x}\frac{\gamma_{\lambda_{0}}(\partial_{x}\Phi(x))}{\partial_{\xi}\gamma_{\lambda_{0}}(\partial_{x}\Phi(x))}\leq\frac{C}{x_{0}}\lambda_{0}(\log\lambda_{0})^{2}.

Higher derivatives of Φ\Phi in (x0,x1)(x_{0},x_{1}). We now consider higher derivatives of Φ\Phi in (x0,x1)(x_{0},x_{1}). For any k2k\geq 2, we claim that

|xkΦ(x)|Ckx0k+1λ0(logλ0)2(k1), for x0<x<x1.\begin{split}|\partial_{x}^{k}\Phi(x)|\leq C_{k}x_{0}^{-k+1}\lambda_{0}(\log\lambda_{0})^{2(k-1)},\quad\hbox{ for }x_{0}<x<x_{1}.\end{split} (4.22)

Note that this corresponds to the bound for xkΦ(X(0))\partial_{x}^{k}\Phi(X(0)) at the initial time. We will obtain bounds for xkΦ(X(t))\partial_{x}^{k}\Phi(X(t)) with t>0t>0 later in Section 4.4, based on the bounds for x2Φ(X(t))\partial_{x}^{2}\Phi(X(t)) and (4.22).

We shall prove (4.22) with an induction in kk; assuming it holds for k=2,,k0+1k=2,\cdots,k_{0}+1 for some k0k_{0}, we compute

xk0+2Φ=xk0(f(x)γλ0(xΦ)+Γf(x)f(x)ξγλ0(xΦ))==0k0C,k0x1ξγλ0(xΦ)(xk0(Γff)+j=0k0Cj,k0xk0j(ff)xjγλ0(xΦ)),\begin{split}\partial_{x}^{k_{0}+2}\Phi&=\partial_{x}^{k_{0}}\left(-\frac{f^{\prime\prime}(x)\gamma_{\lambda_{0}}(\partial_{x}\Phi)+\Gamma f^{\prime\prime}(x)}{f^{\prime}(x)\partial_{\xi}\gamma_{\lambda_{0}}(\partial_{x}\Phi)}\right)\\ &=-\sum_{\ell=0}^{k_{0}}C_{\ell,k_{0}}\partial_{x}^{\ell}\frac{1}{\partial_{\xi}\gamma_{\lambda_{0}}(\partial_{x}\Phi)}\left(\partial_{x}^{k_{0}-\ell}\left(\frac{\Gamma f^{\prime\prime}}{f^{\prime}}\right)+\sum_{j=0}^{k_{0}-\ell}C_{j,k_{0}-\ell}\partial_{x}^{k_{0}-\ell-j}\left(\frac{f^{\prime\prime}}{f^{\prime}}\right)\partial_{x}^{j}\gamma_{\lambda_{0}}(\partial_{x}\Phi)\right),\end{split}

where C,k0C_{\ell,k_{0}} and Cj,k0C_{j,k_{0}-\ell} are combinatorial coefficients. To begin with, it is not difficult to see that

xk0(Γff)Ck0x0k0+1,xk0j(ff)Ck0x0k0++j1.\begin{split}\partial_{x}^{k_{0}-\ell}\left(\frac{\Gamma f^{\prime\prime}}{f^{\prime}}\right)\leq C_{k_{0}}x_{0}^{-k_{0}+\ell-1},\qquad\partial_{x}^{k_{0}-\ell-j}\left(\frac{f^{\prime\prime}}{f^{\prime}}\right)\leq C_{k_{0}}x_{0}^{-k_{0}+\ell+j-1}.\end{split}

Next, we may expand xjγλ0(xΦ)\partial_{x}^{j}\gamma_{\lambda_{0}}(\partial_{x}\Phi) using Faà di Bruno’s formula:

xjγλ0(xΦ)=𝔞:a1+2a2++jaj=jj!a1!1!a1aj!j!aj(a1+ajξγλ0)(xΦ)b=1j(x1+bΦ)ab.\begin{split}\partial_{x}^{j}\gamma_{\lambda_{0}}(\partial_{x}\Phi)=\sum_{\mathfrak{a}:a_{1}+2a_{2}+\cdots+ja_{j}=j}\frac{j!}{a_{1}!1!^{a_{1}}\cdots a_{j}!j!^{a_{j}}}(\partial^{a_{1}\cdots+a_{j}}_{\xi}\gamma_{\lambda_{0}})(\partial_{x}\Phi)\prod_{b=1}^{j}(\partial_{x}^{1+b}\Phi)^{a_{b}}.\end{split}

In the above, the summation is over jj-tuples 𝔞=(a1,,aj)\mathfrak{a}=(a_{1},\cdots,a_{j}) with non-negative integer entries satisfying a1+2a2++jaj=ja_{1}+2a_{2}+\cdots+ja_{j}=j. Using the induction assumption (4.22) and the ellipticity assumption for ξγλ0\partial_{\xi}\gamma_{\lambda_{0}}, we obtain that

|xjγλ0(xΦ)|x0j(logλ0)2j𝔞C𝔞λ0a1+aj(a1+ajξγλ0)(xΦ)Cjx0j(logλ0)2jλ0ξγλ0(xΦ).\begin{split}|\partial_{x}^{j}\gamma_{\lambda_{0}}(\partial_{x}\Phi)|&\leq x_{0}^{-j}(\log\lambda_{0})^{2j}\sum_{\mathfrak{a}}C_{\mathfrak{a}}\lambda_{0}^{a_{1}\cdots+a_{j}}(\partial^{a_{1}\cdots+a_{j}}_{\xi}\gamma_{\lambda_{0}})(\partial_{x}\Phi)\\ &\leq C_{j}x_{0}^{-j}(\log\lambda_{0})^{2j}\lambda_{0}\partial_{\xi}\gamma_{\lambda_{0}}(\partial_{x}\Phi).\end{split}

Similarly, we expand

x1ξγλ0(xΦ)=𝔞:a1+2a2++a=!a1!1!a1a!!a(a1+aξ1ξγλ0)(xΦ)b=1(x1+bΦ)ab.\begin{split}\partial_{x}^{\ell}\frac{1}{\partial_{\xi}\gamma_{\lambda_{0}}(\partial_{x}\Phi)}=\sum_{\mathfrak{a}:a_{1}+2a_{2}+\cdots+\ell a_{\ell}=\ell}\frac{\ell!}{a_{1}!1!^{a_{1}}\cdots a_{\ell}!\ell!^{a_{\ell}}}(\partial^{a_{1}\cdots+a_{\ell}}_{\xi}\frac{1}{\partial_{\xi}\gamma_{\lambda_{0}}})(\partial_{x}\Phi)\prod_{b=1}^{\ell}(\partial_{x}^{1+b}\Phi)^{a_{b}}.\end{split}

Again, using the ellipticity assumption on ξξγλ0\xi\partial_{\xi}\gamma_{\lambda_{0}}, we see that

|(nξ1ξγλ0)(xΦ)|Cnλ0n1ξγλ0(xΦ).\begin{split}\left|(\partial^{n}_{\xi}\frac{1}{\partial_{\xi}\gamma_{\lambda_{0}}})(\partial_{x}\Phi)\right|\leq C_{n}\lambda_{0}^{-n}\frac{1}{\partial_{\xi}\gamma_{\lambda_{0}}}(\partial_{x}\Phi).\end{split}

This gives

|x1ξγλ0(xΦ)|Cx0(logλ0)21ξγλ0.\begin{split}\left|\partial_{x}^{\ell}\frac{1}{\partial_{\xi}\gamma_{\lambda_{0}}(\partial_{x}\Phi)}\right|\leq C_{\ell}x_{0}^{-\ell}(\log\lambda_{0})^{2\ell}\frac{1}{\partial_{\xi}\gamma_{\lambda_{0}}}.\end{split}

Collecting the bounds, we conclude that

|xk0+2Φ|Ck0x0k01λ0(logλ0)2(k0+1)\begin{split}\left|\partial_{x}^{k_{0}+2}\Phi\right|\leq C_{k_{0}}x_{0}^{-k_{0}-1}\lambda_{0}(\log\lambda_{0})^{2(k_{0}+1)}\end{split}

holds, which is just (4.22) with k=k0+2k=k_{0}+2. Therefore, we have arrived at the following

Proposition 4.5.

Let xΦ\partial_{x}\Phi be defined as in (4.5). Then for any k1k\geq 1, xkxΦ\partial_{x}^{k}\partial_{x}\Phi satisfy

|xkxΦ(x)|Ckx0k(logλ0)2kλ0 for x0<x<x1.\begin{split}{|{\partial_{x}^{k}\partial_{x}\Phi(x)}|\leq C_{k}x_{0}^{-k}(\log\lambda_{0})^{2k}\lambda_{0}\qquad\hbox{ for }x_{0}<x<x_{1}.}\end{split} (4.23)

4.3 The case of a time-dependent background

We now generalize our construction of the phase function to the case when ff^{\prime} has time dependence. This time, we employ the method of characteristics to analyze (4.1). The explicit computation we performed in the time-independent case in Section 4.2 will serve as a very useful guide.

Remark 4.6.

A small modification of our scheme can handle ff^{\prime} with a moving zero, by working with the new variable x~=xx̊(t)\tilde{x}=x-\mathring{x}(t) to fix the position of the zero. Then we need to add a term of the form x̊(t)xΦ\mathring{x}^{\prime}(t)\partial_{x}\Phi in the Hamiltonian, which may be handled perturbatively (dominated by the first term in pp) in the regime Mγ(λ0)M\ll\gamma(\lambda_{0}).

In this subsection, we assume that (1.18)–(1.20), as well as (1.24)–(1.25) (i.e., the hypotheses for Theorem B) hold.

Choice of Φ(0,x)\Phi(0,x). Let Λ\Lambda be a large positive parameter to be fixed below. Introduce TT so that

0<T1,tf(t,0)Lt(0,10099tf(T))tf(T)110|f(0,0)|,0<T\leq 1,\qquad\|{\partial_{t}f^{\prime\prime}(t^{\prime},0)}\|_{L^{\infty}_{t^{\prime}}(0,\frac{100}{99}t_{f}(T))}t_{f}(T)\leq\frac{1}{10}|{f^{\prime\prime}(0,0)}|, (4.24)

where we remind the reader that tf(τ)t_{f}(\tau) is defined by τ=0tf(τ)f(t,0)dt\tau=\int_{0}^{t_{f}(\tau)}-f^{\prime\prime}(t^{\prime},0)\,\mathrm{d}t^{\prime}. We then introduce cx0>0c_{x_{0}}>0 such that

supt[0,10099tf(T)]f(t,x)L(0,2cx0ϵ(λ0))cx1110|f(t,0)|,\sup_{t\in[0,\frac{100}{99}t_{f}(T)]}\|{f^{\prime\prime\prime}(t,x)}\|_{L^{\infty}(0,2c_{x_{0}}\epsilon(\lambda_{0}))}c_{x_{1}}\leq\frac{1}{10}|{f^{\prime\prime}(t,0)}|, (4.25)

where ϵ(λ0)\epsilon(\lambda_{0}) is an nonincreasing function of λ0[Ξ0,)\lambda_{0}\in[\Xi_{0},\infty) obeying (4.3) for λ0Λ\lambda_{0}\geq\Lambda. We emphasize that we retain the freedom to shrink T,cx1>0T,c_{x_{1}}>0 and enlarge Λ\Lambda until the end of this section.

We look for the solution Φ\Phi to (4.1) with the initial data

xΦ(0,x)=γλ01(f(0,x1)γλ0((1+12ϵ)λ0)+Γf(0,x)Γf(0,x1)f(0,x))\partial_{x}\Phi(0,x)=\gamma_{\lambda_{0}}^{-1}\left(\frac{f^{\prime}(0,x_{1})\gamma_{\lambda_{0}}((1+\tfrac{1}{2}\epsilon)\lambda_{0})+\Gamma f^{\prime}(0,x)-\Gamma f^{\prime}(0,x_{1})}{-f^{\prime}(0,x)}\right) (4.26)

on {t=0}×(x0,x1)\{t=0\}\times(x_{0},x_{1}), where

x0=cx0ϵ(λ0),x1=x0+Δx0,Δx0x0=110(γλ0((1+34ϵ(λ0))λ0)γλ0((1+12ϵ(λ0))λ0)1),{x_{0}=c_{x_{0}}\epsilon(\lambda_{0}),\quad x_{1}=x_{0}+\Delta x_{0},\quad\frac{\Delta x_{0}}{x_{0}}=\frac{1}{10}\left(\frac{\gamma_{\lambda_{0}}((1+\tfrac{3}{4}\epsilon(\lambda_{0}))\lambda_{0})}{\gamma_{\lambda_{0}}((1+\tfrac{1}{2}\epsilon(\lambda_{0}))\lambda_{0})}-1\right),} (4.27)

and cx0c_{x_{0}} obeys (4.25). Note that (4.26)–(4.27) are precisely (4.5), (4.11) with ff^{\prime} frozen at t=0t=0. The relevance of this choice will be evident in the estimate for x2Φ\partial_{x}^{2}\Phi below. As before, observe that for Λ\Lambda sufficiently large,

λ0xΦ(0,x)(1+ϵ)λ0 on {t=0}×(x0,x1) for λ0Λ.\lambda_{0}\leq\partial_{x}\Phi(0,x)\leq(1+\epsilon)\lambda_{0}\quad\hbox{ on }\{t=0\}\times(x_{0},x_{1})\hbox{ for }\lambda_{0}\geq\Lambda.

Setup for the method of characteristics. We now set up the method of characteristics for (4.1). Let X(t)X(t) be a characteristic curve parametrized by tt, and introduce Ξ(t)=xΦ(t,X(t))\Xi(t)=\partial_{x}\Phi(t,X(t)). The bicharacteristic ODEs for (X,Ξ)(X,\Xi) read as follows:

X˙\displaystyle\dot{X} =ξp(t,X,Ξ)=f(t,X)λ0γλ0(Ξ),\displaystyle=\partial_{\xi}p(t,X,\Xi)=f^{\prime}(t,X)\lambda_{0}\gamma_{\lambda_{0}}^{\prime}(\Xi), (4.28)
Ξ˙\displaystyle\dot{\Xi} =xp(t,X,Ξ)=f(t,X)λ0γλ0(Ξ)+(Γf)(t,X)λ0.\displaystyle=-\partial_{x}p(t,X,\Xi)=-f^{\prime\prime}(t,X)\lambda_{0}\gamma_{\lambda_{0}}(\Xi)+(\Gamma f^{\prime\prime})(t,X)\lambda_{0}. (4.29)

Observe that these equations constitute the Hamiltonian ODEs corresponding to the time-dependent Hamiltonian p(t,x,ξ)p(t,x,\xi). In the following, it will be always assumed that the initial data satisfy x0<X(0)<x1x_{0}<X(0)<x_{1} and Λ0λ0Ξ(0)(1+ϵ)λ0\Lambda_{0}\leq\lambda_{0}\leq\Xi(0)\leq(1+\epsilon)\lambda_{0}.

Continuation criterion and set up for the bootstrap argument. Before we continue, we briefly discuss how our bootstrap argument for constructing and estimating Φ\Phi is set up. Given Φ\Phi, denote by X(t;x¯)X(t;\underline{x}) the solution to X˙=ξp(t,X,xΦ(t,X))\dot{X}=\partial_{\xi}p(t,X,\partial_{x}\Phi(t,X)). Observe that the method of characteristics guarantees the existence of Φ(t,x)\Phi(t,x) for X(t;x0)<x<X(t;x1)X(t;x_{0})<x<X(t;x_{1}), initially for some short time interval [0,t)[0,t^{\ast}). In what follows, we will prove x2Φ(t,x)L(X(t;(x0,x1)))\|{\partial_{x}^{2}\Phi(t,x)}\|_{L^{\infty}(X(t;(x_{0},x_{1})))} is uniformly bounded for t<tt<t^{\ast}, provided that t<10099tf(τM)t^{\ast}<\frac{100}{99}t_{f}(\tau_{M}) (the time scale in Theorem B). Then x¯X(t;x¯)\underline{x}\mapsto X(t^{\ast};\underline{x}) is a bi-Lipschitz isomorphism, and X(t),Ξ(t)X(t),\Xi(t) and Φ(t,X(t))\Phi(t,X(t)) exist on a longer time interval; this allows us to set up a continuous induction (bootstrap) argument for constructing and estimating Φ\Phi.

With such details in mind, in what follows, for the simplicity of exposition, we will simply assume the existence of X(t),Ξ(t)X(t),\Xi(t) and Φ(t,X(t))\Phi(t,X(t)) for 0t10099tf(τM)0\leq t\leq\frac{100}{99}t_{f}(\tau_{M}) and demonstrate how to derive bounds for X(t),Ξ(t)X(t),\Xi(t) and x2Φ(t,X(t))\partial_{x}^{2}\Phi(t,X(t)).

Control of Ξ(t)\Xi(t). To control Ξ(t)\Xi(t), we employ a comparison argument that is similar to the steady case. We introduce λ(t)\lambda(t) and λ¯(t)\overline{\lambda}(t), which are solutions to the ODEs

{λ˙(t)=(1ϵ)f(t,0)λ0γλ0(λ),λ(0)=λ0,\left\{\begin{aligned} \dot{\lambda}(t)&=-(1-\epsilon)f^{\prime\prime}(t,0)\lambda_{0}\gamma_{\lambda_{0}}(\lambda),\\ \lambda(0)&=\lambda_{0},\end{aligned}\right.

and

{λ¯˙(t)=(1+ϵ)f(t,0)λ0γλ0(λ¯),λ¯(0)=(1+ϵ)λ0.\left\{\begin{aligned} \dot{\overline{\lambda}}(t)&=-(1+\epsilon)f^{\prime\prime}(t,0)\lambda_{0}\gamma_{\lambda_{0}}(\overline{\lambda}),\\ \overline{\lambda}(0)&=(1+\epsilon)\lambda_{0}.\end{aligned}\right.

Recall also that tf(τ)t_{f}(\tau) is defined by the relation 0tf(τ)(f)(t,0)dt=τ\int_{0}^{t_{f}(\tau)}(-f^{\prime\prime})(t,0)\,\mathrm{d}t=\tau. For 0t10099tf(T)0\leq t\leq\frac{100}{99}t_{f}(T), we claim that, for ΛλΞ(t)λ¯\Lambda\leq\lambda\leq\Xi(t)\leq\overline{\lambda} and 0<X(t)<x10<X(t)<x_{1},

(1ϵ)f(t,0)λ0γλ0(λ)\displaystyle-(1-\epsilon)f^{\prime\prime}(t,0)\lambda_{0}\gamma_{\lambda_{0}}(\lambda) f(t,X(t))λ0γλ0(Ξ(t))+(Γf)(t,X(t))λ0\displaystyle\leq-f^{\prime\prime}(t,X(t))\lambda_{0}\gamma_{\lambda_{0}}(\Xi(t))+(\Gamma f^{\prime\prime})(t,X(t))\lambda_{0}
(1+ϵ)f(t,0)λ0γλ0(λ¯),\displaystyle\leq-(1+\epsilon)f^{\prime\prime}(t,0)\lambda_{0}\gamma_{\lambda_{0}}(\overline{\lambda}),

provided that Λ\Lambda is sufficiently large depending on γ\gamma and |f(0,0)|Γf|{f^{\prime\prime}(0,0)}|\Gamma f^{\prime\prime}. The proof is similar to the steady case using (4.24), (4.25) and (4.27). The first inequality implies, in particular, that Ξ˙>0\dot{\Xi}>0 for λ0Λ\lambda_{0}\leq\Lambda. Observing furthermore that X˙<0\dot{X}<0 but X(t)>0X(t)>0, and comparing the ODEs for λ(t)\lambda(t), Ξ(t)\Xi(t) and λ¯(t)\overline{\lambda}(t), we arrive at

λ(t)Ξ(t)λ¯(t) for 0t10099tf(T).\lambda(t)\leq\Xi(t)\leq\overline{\lambda}(t)\qquad\hbox{ for }0\leq t\leq\frac{100}{99}t_{f}(T).

The following analogue of Lemma 4.3 holds:

Lemma 4.7.

Let λ0\lambda_{0}, MM and τM\tau_{M} obey (1.18)–(1.20) as well as τMT\tau_{M}\leq T. Then the following statements hold.

  1. 1.

    λ(t)=Mλ0\lambda(t)=M\lambda_{0} exactly at tM:=11ϵtf(τM)t_{M}:=\frac{1}{1-\epsilon}t_{f}(\tau_{M}), and λ0λ(t)Mλ0\lambda_{0}\leq\lambda(t)\leq M\lambda_{0} for 0ttM0\leq t\leq t_{M}.

  2. 2.

    As long as λ(t)Mλ0\lambda(t)\leq M\lambda_{0}, we have

    λ(t)Ξ(t)=xΦ(t,X(t))2λ(t).\lambda(t)\leq\Xi(t)=\partial_{x}\Phi(t,X(t))\leq 2\lambda(t). (4.30)
Proof.

Using the identities

λ0Mλ01γλ0(λ)dλλ0=(1ϵ)tf(t),λ0M¯λ01γλ0(λ)dλλ0=(1+ϵ)tf(t)\int_{\lambda_{0}}^{M^{\prime}\lambda_{0}}\frac{1}{\gamma_{\lambda_{0}}(\lambda)}\frac{\mathrm{d}\lambda}{\lambda_{0}}=(1-\epsilon)t_{f}(t),\qquad\int_{\lambda_{0}}^{\overline{M}^{\prime}\lambda_{0}}\frac{1}{\gamma_{\lambda_{0}}(\lambda)}\frac{\mathrm{d}\lambda}{\lambda_{0}}=(1+\epsilon)t_{f}(t)

in lieu of (4.19) (where Mλ0=λ(t)M^{\prime}\lambda_{0}=\lambda(t) and M¯λ0=λ¯(t)\overline{M}^{\prime}\lambda_{0}=\overline{\lambda}(t) as before), the proof of this lemma proceeds exactly as in that of Lemma 4.3. We omit the details. ∎

Control of X(t)X(t). Next, we aim to obtain a bound for X(t)X(t) that is comparable to (4.20) in the steady case. Instead of a simple argument involving the Hamiltonian (which is exactly preserved in the steady case), here we need an analysis of the bicharacteristic ODEs.

Let 0t11ϵtf(τM)0\leq t\leq\frac{1}{1-\epsilon}t_{f}(\tau_{M}) with τMT\tau_{M}\leq T. Recall the soft fact that 0<X(t)X(0)<x10<X(t)\leq X(0)<x_{1}. Using the bicharacteristic ODEs, we may write

X˙(t)X(t)\displaystyle\frac{\dot{X}(t)}{X(t)} =f(t,X(t))X(t)λ0γλ0(Ξ(t))\displaystyle=\frac{f^{\prime}(t,X(t))}{X(t)}\lambda_{0}\gamma_{\lambda_{0}}^{\prime}(\Xi(t))
=|f(t,X(t))||f(t,X(t))|X(t)γλ0(Ξ(t))γλ0(Ξ(t))(|f(t,X(t))|λ0γλ0(Ξ(t))|f(t,X(t))|λ0γλ0(Ξ(t))+Γf(t,X(t))λ0)Ξ˙(t)\displaystyle=-\frac{|{f^{\prime}(t,X(t))}|}{|{f^{\prime\prime}(t,X(t))}|X(t)}\frac{\gamma_{\lambda_{0}}^{\prime}(\Xi(t))}{\gamma_{\lambda_{0}}(\Xi(t))}\left(\frac{|{f^{\prime\prime}(t,X(t))}|\lambda_{0}\gamma_{\lambda_{0}}(\Xi(t))}{|{f^{\prime\prime}(t,X(t))}|\lambda_{0}\gamma_{\lambda_{0}}(\Xi(t))+\Gamma f^{\prime\prime}(t,X(t))\lambda_{0}}\right)\dot{\Xi}(t)
=|f(t,X(t))||f(t,X(t))|X(t)γλ0(Ξ(t))γλ0(Ξ(t))(1Γf(t,X(t))|f(t,X(t))|γλ0(Ξ(t))+Γf(t,X(t)))Ξ˙(t).\displaystyle=-\frac{|{f^{\prime}(t,X(t))}|}{|{f^{\prime\prime}(t,X(t))}|X(t)}\frac{\gamma_{\lambda_{0}}^{\prime}(\Xi(t))}{\gamma_{\lambda_{0}}(\Xi(t))}\left(1-\frac{\Gamma f^{\prime\prime}(t,X(t))}{|{f^{\prime\prime}(t,X(t))}|\gamma_{\lambda_{0}}(\Xi(t))+\Gamma f^{\prime\prime}(t,X(t))}\right)\dot{\Xi}(t).

By (4.24) and (4.25), for λ0Λ\lambda_{0}\geq\Lambda with Λ\Lambda sufficiently large depending on γ\gamma and |f(0,0)|Γf|{f^{\prime\prime}(0,0)}|\Gamma f^{\prime\prime} on [0,10099tf(T)]×(0,x1)[0,\frac{100}{99}t_{f}(T)]\times(0,x_{1}), we have

||f(t,X(t))||f(t,0)|X(t)1|C0X(t),|Γf(t,X(t))|f(t,X(t))|γλ0(Ξ(t))+Γf(t,X(t))|C11γλ0(Ξ(t)),\left|{\frac{|{f^{\prime}(t,X(t))}|}{|{f^{\prime\prime}(t,0)}|X(t)}-1}\right|\lesssim C_{0}X(t),\qquad\left|{\frac{\Gamma f^{\prime\prime}(t,X(t))}{|{f^{\prime\prime}(t,X(t))}|\gamma_{\lambda_{0}}(\Xi(t))+\Gamma f^{\prime\prime}(t,X(t))}}\right|\lesssim C_{1}\frac{1}{\gamma_{\lambda_{0}}(\Xi(t))},

where the implicit constants are absolute, C0=|f(0,0)|1fL([0,10099tf(T)]×(0,x1))C_{0}=|{f^{\prime\prime}(0,0)}|^{-1}\|{f^{\prime\prime\prime}}\|_{L^{\infty}([0,\frac{100}{99}t_{f}(T)]\times(0,x_{1}))} and C1=|f(0,0)|1ΓfL([0,10099tf(T)]×(0,x1))C_{1}=|{f^{\prime\prime}(0,0)}|^{-1}\|{\Gamma f^{\prime\prime}}\|_{L^{\infty}([0,\frac{100}{99}t_{f}(T)]\times(0,x_{1}))}. By integration in time, it follows that

logX(t)X(0)CC0(X(0)X(t))\displaystyle\log\frac{X(t)}{X(0)}-CC_{0}(X(0)-X(t)) logγλ0(Ξ(0))γλ0(Ξ(t))+CC1(1γλ0(Ξ(0))1γλ0(Ξ(t))),\displaystyle\leq\log\frac{\gamma_{\lambda_{0}}(\Xi(0))}{\gamma_{\lambda_{0}}(\Xi(t))}+CC_{1}\left(\frac{1}{\gamma_{\lambda_{0}}(\Xi(0))}-\frac{1}{\gamma_{\lambda_{0}}(\Xi(t))}\right),
logγλ0(Ξ(0))γλ0(Ξ(t))CC1(1γλ0(Ξ(0))1γλ0(Ξ(t)))\displaystyle\log\frac{\gamma_{\lambda_{0}}(\Xi(0))}{\gamma_{\lambda_{0}}(\Xi(t))}-CC_{1}\left(\frac{1}{\gamma_{\lambda_{0}}(\Xi(0))}-\frac{1}{\gamma_{\lambda_{0}}(\Xi(t))}\right) logX(t)X(0)+CC0(X(0)X(t)),\displaystyle\leq\log\frac{X(t)}{X(0)}+CC_{0}(X(0)-X(t)),

for some absolute constant C>0C>0. By the monotonicity properties of γλ0\gamma_{\lambda_{0}}, XX and Ξ\Xi, observe that all non-logarithmic terms may be made arbitrarily small by taking cx0>0c_{x_{0}}>0 small and Λ\Lambda large depending on γ\gamma, C0C_{0} and C1C_{1}. Finally, by the slow-variance of γλ0\gamma_{\lambda_{0}} and Lemma 4.7, we may replace γλ0(Ξ(0))\gamma_{\lambda_{0}}(\Xi(0)) and γλ0(Ξ(t))\gamma_{\lambda_{0}}(\Xi(t)) by γλ0(λ0)\gamma_{\lambda_{0}}(\lambda_{0}) and γλ0(λ(t))\gamma_{\lambda_{0}}(\lambda(t)), respectively.

In conclusion, for 0t11ϵtf(τM)0\leq t\leq\frac{1}{1-\epsilon}t_{f}(\tau_{M}), τMT\tau_{M}\leq T and λ0Λ\lambda_{0}\geq\Lambda, we obtain

12β0+1X(0)γλ0(λ0)γλ0(λ(t))X(t)2β0+1X(0)γλ0(λ0)γλ0(λ(t)).\begin{split}\frac{1}{{2^{\beta_{0}+1}}}\frac{X(0)\gamma_{\lambda_{0}}(\lambda_{0})}{\gamma_{\lambda_{0}}(\lambda(t))}\leq X(t)\leq{2^{\beta_{0}+1}}\frac{X(0)\gamma_{\lambda_{0}}(\lambda_{0})}{\gamma_{\lambda_{0}}(\lambda(t))}.\end{split} (4.31)

as long as cx0>0c_{x_{0}}>0 is sufficiently small and Λ\Lambda is sufficiently large depending on γ\gamma, |f(0,0)|1f|{f^{\prime\prime}(0,0)}|^{-1}f^{\prime\prime\prime} and |f(0,0)|1Γf|{f^{\prime\prime}(0,0)}|^{-1}\Gamma f^{\prime\prime}.

Control of x2Φ(t,X(t))\partial_{x}^{2}\Phi(t,X(t)). Next, we turn to x2Φ(t,X(t))\partial_{x}^{2}\Phi(t,X(t)). Note that

ddtx2Φ(t,X(t))=tx2Φ(t,X(t))+X˙(t)x3Φ(t,X(t)).\frac{\mathrm{d}}{\mathrm{d}t}\partial_{x}^{2}\Phi(t,X(t))=\partial_{t}\partial_{x}^{2}\Phi(t,X(t))+\dot{X}(t)\partial_{x}^{3}\Phi(t,X(t)).

By the equation

tx2Φ(t,x)+ξp(t,x,xΦ)x3Φ(t,x)+x2p(t,x,xΦ)+2xξp(x,xΦ)x2Φ(t,x)+ξ2p(t,x,xΦ)(x2Φ)2=0,\partial_{t}\partial_{x}^{2}\Phi(t,x)+\partial_{\xi}p(t,x,\partial_{x}\Phi)\partial_{x}^{3}\Phi(t,x)+\partial_{x}^{2}p(t,x,\partial_{x}\Phi)+2\partial_{x}\partial_{\xi}p(x,\partial_{x}\Phi)\partial_{x}^{2}\Phi(t,x)+\partial_{\xi}^{2}p(t,x,\partial_{x}\Phi)(\partial_{x}^{2}\Phi)^{2}=0,

it follows that

ddtx2Φ(t,X(t))\displaystyle\frac{\mathrm{d}}{\mathrm{d}t}\partial_{x}^{2}\Phi(t,X(t)) =x2p(t,X(t),Ξ(t))2xξp(t,X(t),Ξ(t))x2Φ(t,X(t))\displaystyle=-\partial_{x}^{2}p(t,X(t),\Xi(t))-2\partial_{x}\partial_{\xi}p(t,X(t),\Xi(t))\partial_{x}^{2}\Phi(t,X(t)) (4.32)
ξ2p(t,X(t),Ξ(t))(x2Φ(t,X(t)))2.\displaystyle\mathrel{\phantom{=}}-\partial_{\xi}^{2}p(t,X(t),\Xi(t))(\partial_{x}^{2}\Phi(t,X(t)))^{2}.

Note that (4.32) is a Ricatti-type ODE. As is well-known, this ODE is prone to blowing up in finite time, and indeed this is the reflection of the possibility of formation of focal or conjugate points along bicharacteristics in the Hamilton–Jacobi formulation. While it is possible to play with the initial data for Φ\Phi and analyze (4.32) directly to obtain control on x2Φ(t,X(t))\partial_{x}^{2}\Phi(t,X(t)), it is not clear how to obtain control on a sufficiently long time interval needed for our purposes.

Instead, motivated by (4.21), we perform the following inspired change of variables:

h(t):=ξp(t,X(t),Ξ(t))xp(t,X(t),Ξ(t))x2Φ(t,X(t))+1.h(t):=\frac{\partial_{\xi}p(t,X(t),\Xi(t))}{\partial_{x}p(t,X(t),\Xi(t))}\partial_{x}^{2}\Phi(t,X(t))+1. (4.33)

Our choice of the initial data (4.26) is such that hh is initially zero. More precisely, by (4.26), we have

xf(0,0)x1λ0γλ0(λ0)+p(0,x,xΦ(0,x))=0,\partial_{x}f^{\prime}(0,0)x_{1}\lambda_{0}\gamma_{\lambda_{0}}(\lambda_{0})+p(0,x,\partial_{x}\Phi(0,x))=0,

hence by differentiating in xx, it follows that

h(0)=0 for X(0)=x(x0,x1) and Ξ(0)=xΦ(0,x).h(0)=0\qquad\hbox{ for }X(0)=x\in(x_{0},x_{1})\,\mbox{ and }\,\Xi(0)=\partial_{x}\Phi(0,x). (4.34)

We shall show that the variable hh obeys the following remarkably nice evolution equation:

Lemma 4.8.

We have

h˙=s(q+r+s)h+qh2,\dot{h}=s-(q+r+s)h+qh^{2}, (4.35)

where

s:=tξpξp+txpxp,r:=x2pξpxp,q:=ξ2pxpξp,s:=-\frac{\partial_{t}\partial_{\xi}p}{\partial_{\xi}p}+\frac{\partial_{t}\partial_{x}p}{\partial_{x}p},\qquad r:=\partial_{x}^{2}p\frac{\partial_{\xi}p}{\partial_{x}p},\qquad q:=-\partial_{\xi}^{2}p\frac{\partial_{x}p}{\partial_{\xi}p}, (4.36)

which are all evaluated at (t,x,ξ)=(t,X(t),Ξ(t))(t,x,\xi)=(t,X(t),\Xi(t)).

Note that s=0s=0 when pp is time-independent. By (4.34), it is then clear that when pp is time independent, hh remains zero; this is precisely the computation (4.21) in the steady case! The advantage of (4.35) over (4.32) is that we may now incorporate the time dependence of pp in a perturbative manner in the estimate for x2Φ(t,X(t))\partial_{x}^{2}\Phi(t,X(t)).

Proof.

To simplify the notation, we omit the dependence of various derivatives of pp on (t,X(t),Ξ(t))(t,X(t),\Xi(t)), as well as the dependence of x2Φ\partial_{x}^{2}\Phi on (t,X(t))(t,X(t)). We begin by computing

ddtξpxp\displaystyle\frac{\mathrm{d}}{\mathrm{d}t}\frac{\partial_{\xi}p}{\partial_{x}p} =tξp+xξpX˙+ξ2pΞ˙xp(txp+x2pX˙+xξpΞ˙)ξp(xp)2\displaystyle=\frac{\partial_{t}\partial_{\xi}p+\partial_{x}\partial_{\xi}p\dot{X}+\partial_{\xi}^{2}p\dot{\Xi}}{\partial_{x}p}-\frac{(\partial_{t}\partial_{x}p+\partial_{x}^{2}p\dot{X}+\partial_{x}\partial_{\xi}p\dot{\Xi})\partial_{\xi}p}{(\partial_{x}p)^{2}}
=(tξpξp+xξpξ2pxpξptxpxpx2pξpxp+xξp)ξpxp\displaystyle=\left(\frac{\partial_{t}\partial_{\xi}p}{\partial_{\xi}p}+\partial_{x}\partial_{\xi}p-\partial_{\xi}^{2}p\frac{\partial_{x}p}{\partial_{\xi}p}-\frac{\partial_{t}\partial_{x}p}{\partial_{x}p}-\partial_{x}^{2}p\frac{\partial_{\xi}p}{\partial_{x}p}+\partial_{x}\partial_{\xi}p\right)\frac{\partial_{\xi}p}{\partial_{x}p}
=(2xξpξ2pxpξpx2pξpxp+tξpξptxpxp)ξpxp.\displaystyle=\left(2\partial_{x}\partial_{\xi}p-\partial_{\xi}^{2}p\frac{\partial_{x}p}{\partial_{\xi}p}-\partial_{x}^{2}p\frac{\partial_{\xi}p}{\partial_{x}p}+\frac{\partial_{t}\partial_{\xi}p}{\partial_{\xi}p}-\frac{\partial_{t}\partial_{x}p}{\partial_{x}p}\right)\frac{\partial_{\xi}p}{\partial_{x}p}.

Therefore,

ddth\displaystyle\frac{\mathrm{d}}{\mathrm{d}t}h =(ddtξpxp)x2Φ(t,X(t))+ξpxpddtx2Φ(t,X(t))\displaystyle=\left(\frac{\mathrm{d}}{\mathrm{d}t}\frac{\partial_{\xi}p}{\partial_{x}p}\right)\partial_{x}^{2}\Phi(t,X(t))+\frac{\partial_{\xi}p}{\partial_{x}p}\frac{\mathrm{d}}{\mathrm{d}t}\partial_{x}^{2}\Phi(t,X(t))
=(2xξpξ2pxpξpx2pξpxp+tξpξptxpxp)ξpxpx2Φ(t,X(t))\displaystyle=\left(2\partial_{x}\partial_{\xi}p-\partial_{\xi}^{2}p\frac{\partial_{x}p}{\partial_{\xi}p}-\partial_{x}^{2}p\frac{\partial_{\xi}p}{\partial_{x}p}+\frac{\partial_{t}\partial_{\xi}p}{\partial_{\xi}p}-\frac{\partial_{t}\partial_{x}p}{\partial_{x}p}\right)\frac{\partial_{\xi}p}{\partial_{x}p}\partial_{x}^{2}\Phi(t,X(t))
x2pξpxp2xξpξpxpx2Φ(t,X(t))ξ2pξpxp(x2Φ(t,X(t)))2.\displaystyle\mathrel{\phantom{=}}-\partial_{x}^{2}p\frac{\partial_{\xi}p}{\partial_{x}p}-2\partial_{x}\partial_{\xi}p\frac{\partial_{\xi}p}{\partial_{x}p}\partial_{x}^{2}\Phi(t,X(t))-\partial_{\xi}^{2}p\frac{\partial_{\xi}p}{\partial_{x}p}(\partial_{x}^{2}\Phi(t,X(t)))^{2}.

Observe the cancellation of the term 2xξpξpxpx2Φ(t,X(t))2\partial_{x}\partial_{\xi}p\frac{\partial_{\xi}p}{\partial_{x}p}\partial_{x}^{2}\Phi(t,X(t)). Writing ξpxpx2Φ(t,X(t))=h1\frac{\partial_{\xi}p}{\partial_{x}p}\partial_{x}^{2}\Phi(t,X(t))=h-1, we moreover observe that

ddth\displaystyle\frac{\mathrm{d}}{\mathrm{d}t}h =(ξ2pxpξpx2pξpxp+tξpξptxpxp)(h1)\displaystyle=\left(-\partial_{\xi}^{2}p\frac{\partial_{x}p}{\partial_{\xi}p}-\partial_{x}^{2}p\frac{\partial_{\xi}p}{\partial_{x}p}+\frac{\partial_{t}\partial_{\xi}p}{\partial_{\xi}p}-\frac{\partial_{t}\partial_{x}p}{\partial_{x}p}\right)(h-1)
x2pξpxpξ2pxpξp(h1)2\displaystyle\mathrel{\phantom{=}}-\partial_{x}^{2}p\frac{\partial_{\xi}p}{\partial_{x}p}-\partial_{\xi}^{2}p\frac{\partial_{x}p}{\partial_{\xi}p}(h-1)^{2}
=(ξ2pxpξp+x2pξpxptξpξp+txpxp)x2pξpxpξ2pxpξp\displaystyle=\left(\partial_{\xi}^{2}p\frac{\partial_{x}p}{\partial_{\xi}p}+\partial_{x}^{2}p\frac{\partial_{\xi}p}{\partial_{x}p}-\frac{\partial_{t}\partial_{\xi}p}{\partial_{\xi}p}+\frac{\partial_{t}\partial_{x}p}{\partial_{x}p}\right)-\partial_{x}^{2}p\frac{\partial_{\xi}p}{\partial_{x}p}-\partial_{\xi}^{2}p\frac{\partial_{x}p}{\partial_{\xi}p}
(ξ2pxpξpx2pξpxp+tξpξptxpxp+2ξ2pxpξp)hξ2pxpξph2,\displaystyle\mathrel{\phantom{=}}\left(-\partial_{\xi}^{2}p\frac{\partial_{x}p}{\partial_{\xi}p}-\partial_{x}^{2}p\frac{\partial_{\xi}p}{\partial_{x}p}+\frac{\partial_{t}\partial_{\xi}p}{\partial_{\xi}p}-\frac{\partial_{t}\partial_{x}p}{\partial_{x}p}+2\partial_{\xi}^{2}p\frac{\partial_{x}p}{\partial_{\xi}p}\right)h-\partial_{\xi}^{2}p\frac{\partial_{x}p}{\partial_{\xi}p}h^{2},

from which (4.35) follows.∎

We now analyze (4.35) to obtain a uniform control on hh under the additional assumption (1.25) compared to the steady case. Then, by (4.33), x2Φ(t,X(t))\partial_{x}^{2}\Phi(t,X(t)) would enjoy similar estimates as in the steady case.

Proposition 4.9.

Let λ0\lambda_{0}, MM and τM\tau_{M} satisfy (1.18)–(1.20), τMT1\tau_{M}\leq T\leq 1 as well as (1.25). Then for any δ5>0\delta_{5}>0, by taking cx0c_{x_{0}} smaller and Λ\Lambda larger depending on δ5\delta_{5}, γ\gamma and ff (for the precise dependence, see C0C_{0}, C0C_{0}^{\prime} and C1C_{1} in the proof), we have

|h(t)|δ5 for 0t11ϵtf(τM) and λ0Λ.|{h(t)}|\leq\delta_{5}\qquad\hbox{ for }0\leq t\leq\frac{1}{1-\epsilon}t_{f}(\tau_{M})\hbox{ and }\lambda_{0}\geq\Lambda. (4.37)
Proof.

In this proof, all implicit constants are absolute unless otherwise stated. Consider the ODE (4.35). Note that qq can be alternatively written as

q=ξ2γλ0(Ξ(t))ξγλ0(Ξ(t))(xp)(t,X(t),Ξ(t))=(ξlogγλ0)(Ξ(t))Ξ˙(t).q=\frac{\partial_{\xi}^{2}\gamma_{\lambda_{0}}(\Xi(t))}{\partial_{\xi}\gamma_{\lambda_{0}}(\Xi(t))}(-\partial_{x}p)(t,X(t),\Xi(t))=\left(\partial_{\xi}\log\gamma_{\lambda_{0}}^{\prime}\right)(\Xi(t))\dot{\Xi}(t). (4.38)

We begin by expanding the terms in (4.35) that involve t\partial_{t}:

s=tξpξptxpxp\displaystyle-s=\frac{\partial_{t}\partial_{\xi}p}{\partial_{\xi}p}-\frac{\partial_{t}\partial_{x}p}{\partial_{x}p} =tfftxfγλ0(ξ)+tΓffγλ0(ξ)+Γf\displaystyle=\frac{\partial_{t}f^{\prime}}{f^{\prime}}-\frac{\partial_{t}\partial_{x}f^{\prime}\gamma_{\lambda_{0}}(\xi)+\partial_{t}\Gamma f^{\prime\prime}}{f^{\prime\prime}\gamma_{\lambda_{0}}(\xi)+\Gamma f^{\prime\prime}}
=tfftff+tffΓff(γλ0(ξ)+Γf)+tΓff(γλ0(ξ)+Γf).\displaystyle=\frac{\partial_{t}f^{\prime}}{f^{\prime}}-\frac{\partial_{t}f^{\prime\prime}}{f^{\prime\prime}}+\frac{\partial_{t}f^{\prime\prime}}{f^{\prime\prime}}\frac{\Gamma f^{\prime\prime}}{f^{\prime\prime}(\gamma_{\lambda_{0}}(\xi)+\Gamma f^{\prime\prime})}+\frac{\partial_{t}\Gamma f^{\prime\prime}}{f^{\prime\prime}(\gamma_{\lambda_{0}}(\xi)+\Gamma f^{\prime\prime})}.

Observe that, in the first two terms, the terms of order O(1)O(1) cancel and we are left with O(x)O(x). On the other hand, the remaining two terms are bounded by O(1γλ0(ξ))O(\frac{1}{\gamma_{\lambda_{0}}(\xi)}). Combined with Lemma 4.7 and (4.31), we obtain (for λ0Λ\lambda_{0}\geq\Lambda sufficiently large)

|s|\displaystyle|{s}| |f(0,0)|(C1X+C0γλ0(Ξ(t)))|f(0,0)|γλ0(λ0)γλ0(λ(t))(C1cx0ϵ(λ0)+C0γλ0(λ0))\displaystyle\lesssim|{f^{\prime\prime}(0,0)}|\left(C_{1}X+\frac{C_{0}}{\gamma_{\lambda_{0}}(\Xi(t))}\right)\lesssim|{f^{\prime\prime}(0,0)}|\frac{\gamma_{\lambda_{0}}(\lambda_{0})}{\gamma_{\lambda_{0}}(\lambda(t))}\left(C_{1}c_{x_{0}}\epsilon(\lambda_{0})+\frac{C_{0}}{\gamma_{\lambda_{0}}(\lambda_{0})}\right)
γλ0(λ0)γλ0(λ(t))2λ˙(t)λ0(C1cx0ϵ(λ0)+C0γλ0(λ0)),\displaystyle\lesssim\frac{\gamma_{\lambda_{0}}(\lambda_{0})}{\gamma_{\lambda_{0}}(\lambda(t))^{2}}\frac{\dot{\lambda}(t)}{\lambda_{0}}\left(C_{1}c_{x_{0}}\epsilon(\lambda_{0})+\frac{C_{0}}{\gamma_{\lambda_{0}}(\lambda_{0})}\right), (4.39)

where

C0\displaystyle C_{0} =|f(0,0)|2tΓfL([0,11ϵtf(τM)]×(0,2cx0ϵ))+|f(0,0)|1ΓfL([0,11ϵtf(τM)]×(0,2cx0ϵ)),\displaystyle=|{f^{\prime\prime}(0,0)}|^{-2}\|{\partial_{t}\Gamma f^{\prime\prime}}\|_{L^{\infty}([0,\tfrac{1}{1-\epsilon}t_{f}(\tau_{M})]\times(0,2c_{x_{0}}\epsilon))}+|{f^{\prime\prime}(0,0)}|^{-1}\|{\Gamma f^{\prime\prime}}\|_{L^{\infty}([0,\tfrac{1}{1-\epsilon}t_{f}(\tau_{M})]\times(0,2c_{x_{0}}\epsilon))},
C1\displaystyle C_{1} =|f(0,0)|2tfL([0,11ϵtf(τM)]×(0,2cx0ϵ))+|f(0,0)|1fL([0,11ϵtf(τM)]×(0,2cx0ϵ)).\displaystyle=|{f^{\prime\prime}(0,0)}|^{-2}\|{\partial_{t}f^{\prime\prime\prime}}\|_{L^{\infty}([0,\tfrac{1}{1-\epsilon}t_{f}(\tau_{M})]\times(0,2c_{x_{0}}\epsilon))}+|{f^{\prime\prime}(0,0)}|^{-1}\|{f^{\prime\prime\prime}}\|_{L^{\infty}([0,\tfrac{1}{1-\epsilon}t_{f}(\tau_{M})]\times(0,2c_{x_{0}}\epsilon))}.

Next, for the term rr, we again use Lemma 4.7 and (4.31) to estimate

|r|C1X(t)γλ0(λ(t))γλ0(λ(t))λ(t)˙C1cx0ϵ(λ0)γλ0(λ(t))γλ0(λ0)γλ0(λ(t))2λ˙(t),|{r}|\lesssim C_{1}^{\prime}X(t)\frac{\gamma_{\lambda_{0}}^{\prime}(\lambda(t))}{\gamma_{\lambda_{0}}(\lambda(t))}\dot{\lambda(t)}\lesssim C_{1}^{\prime}c_{x_{0}}\epsilon(\lambda_{0})\frac{\gamma_{\lambda_{0}}^{\prime}(\lambda(t))\gamma_{\lambda_{0}}(\lambda_{0})}{\gamma_{\lambda_{0}}(\lambda(t))^{2}}\dot{\lambda}(t), (4.40)

where

C1=C1+γλ0(λ0)1|f(0,0)|1ΓfL([0,11ϵtf(τM)]×(0,2cx0ϵ)).C_{1}^{\prime}=C_{1}+\gamma_{\lambda_{0}}(\lambda_{0})^{-1}|{f^{\prime\prime}(0,0)}|^{-1}\|{\Gamma f^{\prime\prime\prime}}\|_{L^{\infty}([0,\tfrac{1}{1-\epsilon}t_{f}(\tau_{M})]\times(0,2c_{x_{0}}\epsilon))}.

We are now ready to set up a bootstrap argument (continuous induction in time) to prove (4.37). Assume, without any loss of generality, that δ5<δ1\delta_{5}<\delta_{1}. Initially, recall that h(0)=0h(0)=0. As a bootstrap assumption, assume that |h(t)|<δ1|{h(t)}|<\delta_{1} on some time interval [0,t][0,t^{\ast}]. By the method of integrating factor, (4.38) and the bootstrap assumption, we estimate, for any t[0,t]t\in[0,t^{\ast}],

|h(t)|\displaystyle|{h(t)}| |0texp(tt(q+r+sqh)(t)dt)s(t)dt|\displaystyle\leq\left|{\int_{0}^{t}\exp\left(\int_{t^{\prime}}^{t}(q+r+s-qh)(t^{\prime\prime})\,\mathrm{d}t^{\prime\prime}\right)s(t^{\prime})\mathrm{d}t^{\prime}}\right|
0texp(tt(q+|r|+|s|+|q|δ1)(t)dt)|s(t)|dt\displaystyle\leq\int_{0}^{t}\exp\left(\int_{t^{\prime}}^{t}(q+|{r}|+|{s}|+|{q}|\delta_{1})(t^{\prime\prime})\,\mathrm{d}t^{\prime\prime}\right)|{s(t^{\prime})}|\mathrm{d}t^{\prime}
0tmax{γλ0(λ(t))γλ0(M(t)λ0),γλ0(M(t)λ0)γλ0(λ(t))}δ1γλ0(λ(t))γλ0(M(t)λ0)|s(t)|dt\displaystyle\leq\int_{0}^{t}\max\left\{\frac{\gamma_{\lambda_{0}}^{\prime}(\lambda(t^{\prime}))}{\gamma_{\lambda_{0}}^{\prime}(M^{\prime}(t)\lambda_{0})},\frac{\gamma_{\lambda_{0}}^{\prime}(M^{\prime}(t)\lambda_{0})}{\gamma_{\lambda_{0}}^{\prime}(\lambda(t^{\prime}))}\right\}^{\delta_{1}}\frac{\gamma_{\lambda_{0}}^{\prime}(\lambda(t^{\prime}))}{\gamma_{\lambda_{0}}^{\prime}(M^{\prime}(t)\lambda_{0})}|{s(t^{\prime})}|\,\mathrm{d}t^{\prime}
×exp(0t(|s|+|r|)(t)dt).\displaystyle\mathrel{\phantom{\leq}}\times\exp\left(\int_{0}^{t}(|{s}|+|{r}|)(t^{\prime\prime})\,\mathrm{d}t^{\prime\prime}\right).

We first claim that the last term is uniformly bounded in tt. Indeed, by (4.39), monotonicity of γλ0\gamma_{\lambda_{0}}, Lemma 4.7 (by which M(t)MM^{\prime}(t)\leq M) and τMT1\tau_{M}\leq T\leq 1,

0t|s|(t)dt\displaystyle\int_{0}^{t}|{s}|(t^{\prime})\,\mathrm{d}t^{\prime} λ0M(t)λ0γλ0(λ0)γλ0(λ(t))2dλλ0(C1cx0ϵ(λ0)+C0γλ0(λ0))\displaystyle\lesssim\int_{\lambda_{0}}^{M^{\prime}(t)\lambda_{0}}\frac{\gamma_{\lambda_{0}}(\lambda_{0})}{\gamma_{\lambda_{0}}(\lambda(t))^{2}}\frac{\mathrm{d}\lambda}{\lambda_{0}}\left(C_{1}c_{x_{0}}\epsilon(\lambda_{0})+\frac{C_{0}}{\gamma_{\lambda_{0}}(\lambda_{0})}\right)
λ0M(t)λ01γλ0(λ(t))dλλ0(C1cx0ϵ(λ0)+C0γλ0(λ0))\displaystyle\leq\int_{\lambda_{0}}^{M^{\prime}(t)\lambda_{0}}\frac{1}{\gamma_{\lambda_{0}}(\lambda(t))}\frac{\mathrm{d}\lambda}{\lambda_{0}}\left(C_{1}c_{x_{0}}\epsilon(\lambda_{0})+\frac{C_{0}}{\gamma_{\lambda_{0}}(\lambda_{0})}\right)
(C1cx0ϵ(λ0)+C0γλ0(λ0)).\displaystyle\lesssim\left(C_{1}c_{x_{0}}\epsilon(\lambda_{0})+\frac{C_{0}}{\gamma_{\lambda_{0}}(\lambda_{0})}\right). (4.41)

Next, by (4.40) and M(t)MM^{\prime}(t)\leq M

0t|r|(t)dtC1cx0ϵ(λ0)λ0M(t)λ0γλ0(λ0)γλ0(λ)γλ0(λ)2dλC1cx0ϵ(λ0).\displaystyle\int_{0}^{t}|{r}|(t^{\prime})\,\mathrm{d}t^{\prime}\lesssim C_{1}^{\prime}c_{x_{0}}\epsilon(\lambda_{0})\int_{\lambda_{0}}^{M^{\prime}(t)\lambda_{0}}\frac{\gamma_{\lambda_{0}}(\lambda_{0})\gamma_{\lambda_{0}}^{\prime}(\lambda)}{\gamma_{\lambda_{0}}(\lambda)^{2}}\,\mathrm{d}\lambda\leq C_{1}^{\prime}c_{x_{0}}\epsilon(\lambda_{0}). (4.42)

In particular, by taking cx0>0c_{x_{0}}>0 smaller and Λ\Lambda larger depending on γ\gamma, C0C_{0}, C1C_{1} and C1C_{1}^{\prime}, we may ensure that the RHSs of (4.41) and (4.42) are smaller than, say, 11. Thus,

|h(t)|\displaystyle|{h(t)}| λ0M(t)λ0max{γλ0(λ)γλ0(M(t)λ0),γλ0(M(t)λ0)γλ0(λ)}δ1γλ0(λ)γλ0(M(t)λ0)γλ0(λ0)γλ0(λ)2dλλ0\displaystyle\lesssim\int_{\lambda_{0}}^{M^{\prime}(t)\lambda_{0}}\max\left\{\frac{\gamma_{\lambda_{0}}^{\prime}(\lambda)}{\gamma_{\lambda_{0}}^{\prime}(M^{\prime}(t)\lambda_{0})},\frac{\gamma_{\lambda_{0}}^{\prime}(M^{\prime}(t)\lambda_{0})}{\gamma_{\lambda_{0}}^{\prime}(\lambda)}\right\}^{\delta_{1}}\frac{\gamma_{\lambda_{0}}^{\prime}(\lambda)}{\gamma_{\lambda_{0}}^{\prime}(M^{\prime}(t)\lambda_{0})}\frac{\gamma_{\lambda_{0}}(\lambda_{0})}{\gamma_{\lambda_{0}}(\lambda)^{2}}\frac{\mathrm{d}\lambda}{\lambda_{0}}
×(C1cx0ϵ(λ0)+C0γλ0(λ0)).\displaystyle\mathrel{\phantom{\leq}}\times\left(C_{1}c_{x_{0}}\epsilon(\lambda_{0})+\frac{C_{0}}{\gamma_{\lambda_{0}}(\lambda_{0})}\right).

We split the λ\lambda-integral as follows. Define E:={λ(0,M(t)λ0):γλ0(M(t)λ0)1γλ0(λ)1}E:=\{\lambda\in(0,M^{\prime}(t)\lambda_{0}):\gamma_{\lambda_{0}}^{\prime}(M^{\prime}(t)\lambda_{0})^{-1}\gamma_{\lambda_{0}}^{\prime}(\lambda)\leq 1\}. As in the proof of (4.41), we have

(0,M(t)λ0)Emax{γλ0(λ)γλ0(M(t)λ0),γλ0(M(t)λ0)γλ0(λ)}δ1γλ0(λ)γλ0(M(t)λ0)γλ0(λ0)γλ0(λ)2dλλ0\displaystyle\int_{(0,M^{\prime}(t)\lambda_{0})\cap E}\max\left\{\frac{\gamma_{\lambda_{0}}^{\prime}(\lambda)}{\gamma_{\lambda_{0}}^{\prime}(M^{\prime}(t)\lambda_{0})},\frac{\gamma_{\lambda_{0}}^{\prime}(M^{\prime}(t)\lambda_{0})}{\gamma_{\lambda_{0}}^{\prime}(\lambda)}\right\}^{\delta_{1}}\frac{\gamma_{\lambda_{0}}^{\prime}(\lambda)}{\gamma_{\lambda_{0}}^{\prime}(M^{\prime}(t)\lambda_{0})}\frac{\gamma_{\lambda_{0}}(\lambda_{0})}{\gamma_{\lambda_{0}}(\lambda)^{2}}\frac{\mathrm{d}\lambda}{\lambda_{0}}
0M(t)λγλ0(λ0)γλ0(λ)2dλλ0τMT1.\displaystyle\leq\int_{0}^{M^{\prime}(t)\lambda}\frac{\gamma_{\lambda_{0}}(\lambda_{0})}{\gamma_{\lambda_{0}}(\lambda)^{2}}\frac{\mathrm{d}\lambda}{\lambda_{0}}\leq\tau_{M}\leq T\leq 1.

On the other hand, the λ\lambda-integral on (0,M(t)λ0)E(0,M^{\prime}(t)\lambda_{0})\setminus E is bounded from above by the LHS of (1.25). In conclusion,

|h(t)|\displaystyle|{h(t)}| (C1cx0ϵ(λ0)+C0γλ0(λ0)).\displaystyle\lesssim\left(C_{1}c_{x_{0}}\epsilon(\lambda_{0})+\frac{C_{0}}{\gamma_{\lambda_{0}}(\lambda_{0})}\right).

Taking cx0>0c_{x_{0}}>0 smaller and Λ\Lambda larger depending on δ5\delta_{5}, γ\gamma, C0C_{0} and C1C_{1}, we obtain |h(t)|δ5|{h(t)}|\leq\delta_{5}, which improves the bootstrap assumption. ∎

4.4 Estimates for the phase along characteristic curves

We pick up from the end of either Section 4.2 and 4.3; that is, we assume that the hypotheses, and therefore the conclusions, of one of these sections hold. Our goal in this subsection is to obtain sharp bounds for the derivatives of Φ\Phi, which are essential for sharp estimates for the amplitude function in Section 4.5 below. Our key idea is again to consider a renormalization of the form (4.33).

To this end, we consider the transport operator

t:=t+ξp(t,x,xΦ)x.\nabla_{t}:=\partial_{t}+\partial_{\xi}p(t,x,\partial_{x}\Phi)\partial_{x}.

Observe that the characteristics for this operator are precisely X(t)X(t) in Sections 4.2 and 4.3. Moreover, the solutions to tϕ=0\nabla_{t}\phi=0 obey a-priori LL^{\infty}-bounds. We introduce the commutator notation

[t,x]=Ax, or equivalently, A=x(ξp(t,x,xΦ(t,x))).[\nabla_{t},\partial_{x}]=A\partial_{x},\quad\hbox{ or equivalently, }A=-\partial_{x}\left(\partial_{\xi}p(t,x,\partial_{x}\Phi(t,x))\right). (4.43)

We generalize the characteristic-wise definition (4.33) of hh by

h(t,x)=ξp(t,x,xΦ(t,x))xp(t,x,xΦ(t,x))x2Φ(t,x)+1.h(t,x)=\frac{\partial_{\xi}p(t,x,\partial_{x}\Phi(t,x))}{\partial_{x}p(t,x,\partial_{x}\Phi(t,x))}\partial_{x}^{2}\Phi(t,x)+1.

The transport equations for hh and its the derivatives then follow from Lemma 4.8:

txkh=kAxkh(q+r+s)xkh+2qhxkh+xks+=1k1Ck,1xAxkh+=1kCk,2x(q+r+s)xkh+=1km=0kC3k,,mxqxmhxkmh.\begin{split}\nabla_{t}\partial_{x}^{k}h&=kA\partial_{x}^{k}h-(q+r+s)\partial_{x}^{k}h+2qh\partial_{x}^{k}h+\partial_{x}^{k}s\\ &\mathrel{\phantom{=}}+\sum_{\ell=1}^{k-1}C_{k,\ell}^{1}\partial_{x}^{\ell}A\partial_{x}^{k-\ell}h+\sum_{\ell=1}^{k}C_{k,\ell}^{2}\partial_{x}^{\ell}\left(q+r+s\right)\partial_{x}^{k-\ell}h\\ &\mathrel{\phantom{=}}+\sum_{\ell=1}^{k}\sum_{m=0}^{k-\ell}C^{3}_{k,\ell,m}\partial_{x}^{\ell}q\partial_{x}^{m}h\partial_{x}^{k-\ell-m}h.\end{split} (4.44)

In the above, qq, rr and ss are defined as in (4.36) but evaluated at (t,x,ξ)=(t,x,xΦ(t,x))(t,x,\xi)=(t,x,\partial_{x}\Phi(t,x)), and Ck,1C_{k,\ell}^{1}, Ck,2C_{k,\ell}^{2} and Ck,,m3C_{k,\ell,m}^{3} are some combinatorial coefficients.

Before we continue, we note the following key commutator computation:

[t,x]=Ax\displaystyle{}[\nabla_{t},\partial_{x}]=A\partial_{x} =(xξp(x,xΦ)ξ2p(x,xΦ)xp(x,xΦ)ξp(x,xΦ)(1h))x\displaystyle=-\left(\partial_{x}\partial_{\xi}p(x,\partial_{x}\Phi)-\partial_{\xi}^{2}p(x,\partial_{x}\Phi)\frac{\partial_{x}p(x,\partial_{x}\Phi)}{\partial_{\xi}p(x,\partial_{x}\Phi)}(1-h)\right)\partial_{x} (4.45)
=(ξxp(x,xΦ)xp(x,xΦ)ξ2p(x,xΦ)ξp(x,xΦ)(1h))xp(x,xΦ)x\displaystyle=-\left(\frac{\partial_{\xi}\partial_{x}p(x,\partial_{x}\Phi)}{\partial_{x}p(x,\partial_{x}\Phi)}-\frac{\partial_{\xi}^{2}p(x,\partial_{x}\Phi)}{\partial_{\xi}p(x,\partial_{x}\Phi)}(1-h)\right)\partial_{x}p(x,\partial_{x}\Phi)\partial_{x}
=(γλ0(Ξ)γλ0(Ξ)γλ0(Ξ)γλ0(Ξ)+O(δ5+γλ0(λ0)1)Ξ)Ξ˙x,\displaystyle=\left(\frac{\gamma^{\prime}_{\lambda_{0}}(\Xi)}{\gamma_{\lambda_{0}}(\Xi)}-\frac{\gamma^{\prime\prime}_{\lambda_{0}}(\Xi)}{\gamma^{\prime}_{\lambda_{0}}(\Xi)}+\frac{O(\delta_{5}+\gamma_{\lambda_{0}}(\lambda_{0})^{-1})}{\Xi}\right)\dot{\Xi}\partial_{x},

where O()O(\cdot) refers to a term whose absolute value is bounded by C()C(\cdot) with CC depending on γ\gamma and |f(0,0)|1Γf|{f^{\prime\prime}(0,0)}|^{-1}\Gamma f^{\prime\prime}. In what follows, we shall take Λ0\Lambda_{0} larger so that γλ0(λ0)1<δ5\gamma_{\lambda_{0}}(\lambda_{0})^{-1}<\delta_{5}.

Proposition 4.10.

Let N0N_{0} be a positive integer. There exist δ3δ2N0\delta_{3}\ll\frac{\delta_{2}}{N_{0}} and δ5δ0δ3\delta_{5}\ll\delta_{0}\delta_{3} such that, the following holds. If |h(t,X(t))|δ5|{h(t,X(t))}|\leq\delta_{5} for 0t11ϵtf(τM)0\leq t\leq\frac{1}{1-\epsilon}t_{f}(\tau_{M}), where τM1\tau_{M}\leq 1, then for k=1,,N0k=1,\ldots,N_{0},

|xkxΦ(t,X(t))|\displaystyle|{\partial_{x}^{k}\partial_{x}\Phi(t,X(t))}| μ(t)k1(x2Φ(t,X(t))),\displaystyle\leq\mu(t)^{k-1}(-\partial_{x}^{2}\Phi(t,X(t))), (4.46)
|xkxΦ(t,X(t))|\displaystyle|{\partial_{x}^{k}\partial_{x}\Phi(t,X(t))}| μ(t)kλ(t),\displaystyle\leq\mu(t)^{k}\lambda(t), (4.47)

where

μ(t)=λ02δ3N0x0ϵ(λ0)γλ0(λ(t))γλ0(λ0)γλ0(λ0)γλ0(λ(t)).\mu(t)=\frac{\lambda_{0}^{2\delta_{3}N_{0}}}{x_{0}\epsilon(\lambda_{0})}\frac{\gamma_{\lambda_{0}}(\lambda(t))}{\gamma_{\lambda_{0}}(\lambda_{0})}\frac{\gamma_{\lambda_{0}}^{\prime}(\lambda_{0})}{\gamma_{\lambda_{0}}^{\prime}(\lambda(t))}. (4.48)
Remark 4.11.

We remark that the requirement |h(t,X(t))|δ5|{h(t,X(t))}|\leq\delta_{5} is vacuous in the steady case (Section 4.2), whereas in the time-dependent case it may be ensured by taking Λ\Lambda larger and cx0c_{x_{0}} smaller depending on δ5\delta_{5} (Proposition 4.9 in Section 4.3).

Proof.

In the proof, we fix a bicharacteristic curve (X(t),Ξ(t))(X(t),\Xi(t)) with x0<X(0)<x1x_{0}<X(0)<x_{1} and Ξ(0)=xΦ(0,X(0))\Xi(0)=\partial_{x}\Phi(0,X(0)).

An integration factor. For simplicity, we set

I(t):=0t(ξxpxpξ2pξp(1h))(τ,X(τ),Ξ(τ))(xp)(τ,X(τ),Ξ(τ))dτ.\begin{split}I(t):=\int_{0}^{t}\left(\frac{\partial_{\xi}\partial_{x}p}{\partial_{x}p}-\frac{\partial_{\xi}^{2}p}{\partial_{\xi}p}(1-h)\right)(\tau,X(\tau),\Xi(\tau))(-\partial_{x}p)(\tau,X(\tau),\Xi(\tau))\,\mathrm{d}\tau.\end{split} (4.49)

Recalling the computations following (4.45), we have the bounds

γλ0(Ξ)γλ0(λ0)γλ0(λ0)γλ0(Ξ)(Ξλ0)δ4exp(I(t))γλ0(Ξ)γλ0(λ0)γλ0(λ0)γλ0(Ξ)(Ξλ0)δ4\begin{split}\frac{\gamma_{\lambda_{0}}(\Xi)}{\gamma_{\lambda_{0}}(\lambda_{0})}\frac{\gamma_{\lambda_{0}}^{\prime}(\lambda_{0})}{\gamma_{\lambda_{0}}^{\prime}(\Xi)}\left(\frac{\Xi}{\lambda_{0}}\right)^{-\delta_{4}}\leq\exp(I(t))\leq\frac{\gamma_{\lambda_{0}}(\Xi)}{\gamma_{\lambda_{0}}(\lambda_{0})}\frac{\gamma_{\lambda_{0}}^{\prime}(\lambda_{0})}{\gamma_{\lambda_{0}}^{\prime}(\Xi)}\left(\frac{\Xi}{\lambda_{0}}\right)^{\delta_{4}}\end{split} (4.50)

where δ4=Cδ5\delta_{4}=C\delta_{5} with some constant C>0C>0 depending on γ\gamma and |f(0,0)|1Γf|{f^{\prime\prime}(0,0)}|^{-1}\Gamma f^{\prime\prime}. We shall establish the claimed bounds by propagating a sharp estimate based on the quantity I(t)I(t).

Induction base case. The following bound for hh, which could be weaker than (4.37), is more convenient as an induction hypothesis:

|h(t,X(t))|λ0δ341γλ0(Ξ(t))1λ0.\begin{split}|{h(t,X(t))}|\leq\lambda_{0}^{\frac{\delta_{3}}{4}}\frac{1}{\gamma_{\lambda_{0}}^{\prime}(\Xi(t))}\frac{1}{\lambda_{0}}.\end{split} (4.51)

To prove this, we introduce h(0)(t)=λ0γλ0(Ξ(t))h(t,X(t))h^{(0)}(t)=\lambda_{0}\gamma_{\lambda_{0}}^{\prime}(\Xi(t))h(t,X(t)), which incorporates an integrating factor for the term qhqh. Then

ddth(0)+(r+s2qh)h(0)=λ0γλ0(Ξ(t))s.\frac{\mathrm{d}}{\mathrm{d}t}h^{(0)}+(r+s-2qh)h^{(0)}=\lambda_{0}\gamma_{\lambda_{0}}^{\prime}(\Xi(t))s.

By the ellipticity assumption for ξξγλ0\xi\partial_{\xi}\gamma_{\lambda_{0}} and Ξ˙=xp\dot{\Xi}=-\partial_{x}p, we have

2|qh|Cδ5Ξ˙Ξ,2|{qh}|\leq C\delta_{5}\frac{\dot{\Xi}}{\Xi}, (4.52)

where CC depends on γ\gamma. Using (4.39), (4.41) and (4.42), we obtain

|h(0)(t)|\displaystyle|{h^{(0)}(t)}| 0texp(tt|r|+|s|+2|qh|dt)λ0γλ0(Ξ)|s|dt\displaystyle\leq\int_{0}^{t}\exp\left(\int_{t^{\prime}}^{t}|{r}|+|{s}|+2|{qh}|\,\mathrm{d}t^{\prime\prime}\right)\lambda_{0}\gamma_{\lambda_{0}}^{\prime}(\Xi)|{s}|\,\mathrm{d}t^{\prime}
(Ξ(t)λ0)Cδ50tγλ0(Ξ)γλ0(λ0)γλ0(Ξ)2Ξ˙dt.\displaystyle\lesssim\left(\frac{\Xi(t)}{\lambda_{0}}\right)^{C\delta_{5}}\int_{0}^{t}\frac{\gamma_{\lambda_{0}}^{\prime}(\Xi)\gamma_{\lambda_{0}}(\lambda_{0})}{\gamma_{\lambda_{0}}(\Xi)^{2}}\dot{\Xi}\,\mathrm{d}t^{\prime}.

The first factor is bounded by λ0Cδ5δ0\lambda_{0}^{\frac{C\delta_{5}}{\delta_{0}}}, where as the tt^{\prime}-integral is evaluated and bounded as

0tγλ0(Ξ)γλ0(λ0)γλ0(Ξ)2Ξ˙dt=λ0Ξ(t)γλ0(Ξ)γλ0(λ0)γλ0(Ξ)2dΞ=1γλ0(λ0)γλ0(Ξ(t))1.\int_{0}^{t}\frac{\gamma_{\lambda_{0}}^{\prime}(\Xi)\gamma_{\lambda_{0}}(\lambda_{0})}{\gamma_{\lambda_{0}}(\Xi)^{2}}\dot{\Xi}\,\mathrm{d}t^{\prime}=\int_{\lambda_{0}}^{\Xi(t)}\frac{\gamma_{\lambda_{0}}^{\prime}(\Xi)\gamma_{\lambda_{0}}(\lambda_{0})}{\gamma_{\lambda_{0}}(\Xi)^{2}}\,\mathrm{d}\Xi=1-\frac{\gamma_{\lambda_{0}}(\lambda_{0})}{\gamma_{\lambda_{0}}(\Xi(t))}\leq 1. (4.53)

Hence, by taking δ5δ0δ3\delta_{5}\ll\delta_{0}\delta_{3} and returning to hh, (4.51) follows.

Induction hypothesis. We now turn to the case k01k_{0}\geq 1. When k0>1k_{0}>1, we assume the following in addition to |h|δ5|{h}|\leq\delta_{5}: For k=1,,k01k=1,\ldots,k_{0}-1,

|kxh(t,X(t))|\displaystyle|\partial^{k}_{x}h(t,X(t))| λ0δ3k2x0kϵ(λ0)kexp(kI(t))1γλ0(Ξ(t))1λ0,\displaystyle\leq\lambda_{0}^{\delta_{3}k^{2}}x_{0}^{-k}{\epsilon(\lambda_{0})^{-k}}\exp(kI(t))\frac{1}{\gamma_{\lambda_{0}}^{\prime}(\Xi(t))}\frac{1}{\lambda_{0}}, (4.54)
|kxx2Φ(t,X(t))|\displaystyle|\partial^{k}_{x}\partial_{x}^{2}\Phi(t,X(t))| λ0δ3((k+1)21)x0kϵ(λ0)kexp(kI(t))(x2Φ(t,X(t))).\displaystyle\leq\lambda_{0}^{\delta_{3}((k+1)^{2}-1)}x_{0}^{-k}{\epsilon(\lambda_{0})^{-k}}\exp(kI(t))(-\partial_{x}^{2}\Phi(t,X(t))). (4.55)

To handle the quadratic term in hh, it is easier to work with the following simplification of (4.54). Note that

1λ0γλ0(Ξ)Ξλ0γλ0(Ξ)(logΞ)2max{1γλ0(λ0),τM}δ02(logλ0)2δ02(logλ0)2.\frac{1}{\lambda_{0}\gamma_{\lambda_{0}}^{\prime}(\Xi)}\lesssim\frac{\Xi}{\lambda_{0}\gamma_{\lambda_{0}}(\Xi)}(\log\Xi)^{2}\lesssim\max\left\{\frac{1}{\gamma_{\lambda_{0}}(\lambda_{0})},\tau_{M}\right\}\delta_{0}^{-2}(\log\lambda_{0})^{2}\lesssim\delta_{0}^{-2}(\log\lambda_{0})^{2}. (4.56)

In the second inequality, we used the fact that Ξλ0γλ0(Ξ)\frac{\Xi}{\lambda_{0}\gamma_{\lambda_{0}}(\Xi)} is clearly bounded by 1γλ0(λ0)\frac{1}{\gamma_{\lambda_{0}}(\lambda_{0})} for Ξ4\Xi\leq 4, and by τM\tau_{M} for Ξ4\Xi\geq 4 thanks to λΞ2λ\lambda\leq\Xi\leq 2\lambda and λMλ0\lambda\leq M\lambda_{0}. As a consequence of |h|δ5|h|\leq\delta_{5} for k=0k=0 and (4.54) and (4.56) for k1k\geq 1, we obtain, for k=0,,k01k=0,\ldots,k_{0}-1,

|x0kϵ(λ0)kexp(kI(t))kxh(t,X(t))|δ0λ0δ3k2(logλ0)2.|{x_{0}^{k}{\epsilon(\lambda_{0})^{k}}\exp(-kI(t))\partial^{k}_{x}h(t,X(t))}|\lesssim_{\delta_{0}}\lambda_{0}^{\delta_{3}k^{2}}(\log\lambda_{0})^{2}. (4.57)

Induction argument. Our goal now is to prove (4.54) and (4.55) for k=k0k=k_{0}.

To use (4.44) to establish (4.54) for k=k0k=k_{0}, we work with new variables that incorporate integrating factors for cancelling the large coefficient kAqkA-q. For k1k\geq 1, define

h(k)(t)\displaystyle h^{(k)}(t) =x0kϵ(λ0)kexp(kI(t))λ0γλ0(Ξ(t))xkh(t,X(t)),\displaystyle=x_{0}^{k}{\epsilon(\lambda_{0})^{k}}\exp(-kI(t))\lambda_{0}\gamma_{\lambda_{0}}^{\prime}(\Xi(t))\partial_{x}^{k}h(t,X(t)), A(k)(t)\displaystyle A^{(k)}(t) =x0kϵ(λ0)kexp(kI(t))xkA(t,X(t)),\displaystyle=x_{0}^{k}{\epsilon(\lambda_{0})^{k}}\exp(-kI(t))\partial_{x}^{k}A(t,X(t)),
s(k)(t)\displaystyle s^{(k)}(t) =x0kϵ(λ0)kexp(kI(t))xks(t,X(t)),\displaystyle=x_{0}^{k}{\epsilon(\lambda_{0})^{k}}\exp(-kI(t))\partial_{x}^{k}s(t,X(t)), r(k)(t)\displaystyle r^{(k)}(t) =x0kϵ(λ0)kexp(kI(t))xkr(t,X(t)),\displaystyle=x_{0}^{k}{\epsilon(\lambda_{0})^{k}}\exp(-kI(t))\partial_{x}^{k}r(t,X(t)),
q(k)(t)\displaystyle q^{(k)}(t) =x0kϵ(λ0)kexp(kI(t))xkq(t,X(t)).\displaystyle=x_{0}^{k}{\epsilon(\lambda_{0})^{k}}\exp(-kI(t))\partial_{x}^{k}q(t,X(t)).

Observe that (4.51) and (4.54) are equivalent to

|h(0)|λ0δ34,|h(k)|λ0δ3k2 for 1k<k0.|{h^{(0)}}|\leq\lambda_{0}^{\frac{\delta_{3}}{4}},\quad|{h^{(k)}}|\leq\lambda_{0}^{\delta_{3}k^{2}}\quad\hbox{ for }1\leq k<k_{0}. (4.58)

Moreover, evaluating along a characteristic curve, we note that (4.44) can be written as

ddth(k)+(r+s2qh)h(k)\displaystyle\frac{\mathrm{d}}{\mathrm{d}t}h^{(k)}+(r+s-2qh)h^{(k)} =λ0γλ0(Ξ(t))s(k)+=1k1C1k,A()h(k)\displaystyle=\lambda_{0}\gamma_{\lambda_{0}}^{\prime}(\Xi(t))s^{(k)}+\sum_{\ell=1}^{k-1}C^{1}_{k,\ell}A^{(\ell)}h^{(k-\ell)} (4.59)
+=1kC2k,(q()+r()+s())h(k)\displaystyle\mathrel{\phantom{=}}+\sum_{\ell=1}^{k}C^{2}_{k,\ell}\left(q^{(\ell)}+r^{(\ell)}+s^{(\ell)}\right)h^{(k-\ell)}
+=1km=0kC3k,,mq()λ0γλ0(Ξ)h(m)h(km)\displaystyle\mathrel{\phantom{=}}+\sum_{\ell=1}^{k}\sum_{m=0}^{k-\ell}C^{3}_{k,\ell,m}\frac{q^{(\ell)}}{\lambda_{0}\gamma_{\lambda_{0}}^{\prime}(\Xi)}h^{(m)}h^{(k-\ell-m)}
+m=1k1C3k,0,mqλ0γλ0(Ξ)h(m)h(km).\displaystyle\mathrel{\phantom{=}}+\sum_{m=1}^{k-1}C^{3}_{k,0,m}\frac{q}{\lambda_{0}\gamma_{\lambda_{0}}^{\prime}(\Xi)}h^{(m)}h^{(k-m)}.

We shall prove the following estimates for the coefficients of (4.59): for 1kk01\leq k\leq k_{0},

|s(k)(t)|\displaystyle|{s^{(k)}(t)}| λ0δ3(k212)γλ0(λ0)γλ0(Ξ(t))2λ0Ξ˙(t)(Ck,1x0+Ck,0γλ0(λ0)),\displaystyle\leq\lambda_{0}^{\delta_{3}(k^{2}-\frac{1}{2})}\frac{\gamma_{\lambda_{0}}(\lambda_{0})}{\gamma_{\lambda_{0}}(\Xi(t))^{2}\lambda_{0}}\dot{\Xi}(t)\left(C_{k,1}x_{0}+\frac{C_{k,0}}{\gamma_{\lambda_{0}}(\lambda_{0})}\right), (4.60)
|r(k)(t)|\displaystyle|{r^{(k)}(t)}| Ck,0Ck,1λ0δ3(k212)γλ0(Ξ(t))γλ0(λ0)γλ0(Ξ(t))2Ξ˙(t)x0,\displaystyle\leq C_{k,0}C_{k,1}\lambda_{0}^{\delta_{3}(k^{2}-\frac{1}{2})}\frac{\gamma_{\lambda_{0}}^{\prime}(\Xi(t))\gamma_{\lambda_{0}}(\lambda_{0})}{\gamma_{\lambda_{0}}(\Xi(t))^{2}}\dot{\Xi}(t)x_{0}, (4.61)
|q(k)(t)|\displaystyle|{q^{(k)}(t)}| Ck,0λ0δ3(k212)Ξ˙(t)Ξ(t),\displaystyle\leq C_{k,0}\lambda_{0}^{\delta_{3}(k^{2}-\frac{1}{2})}\frac{\dot{\Xi}(t)}{\Xi(t)}, (4.62)

and for k>1k>1,

|A(k1)(t)|\displaystyle|{A^{(k-1)}(t)}| Ck1,0λ0δ3((k1)212)Ξ˙(t)Ξ(t).\displaystyle\leq C_{k-1,0}\lambda_{0}^{\delta_{3}((k-1)^{2}-\frac{1}{2})}\frac{\dot{\Xi}(t)}{\Xi(t)}. (4.63)

Here, Ck,1C_{k,1} depends on |f(0,0)|1(xx)kf|{f^{\prime\prime}(0,0)}|^{-1}(x\partial_{x})^{k^{\prime}}f^{\prime\prime\prime}, |f(0,0)|1(xx)kΓf|{f^{\prime\prime}(0,0)}|^{-1}(x\partial_{x})^{k^{\prime}}\Gamma f^{\prime\prime\prime} and |f(0,0)|2(xx)ktf|{f^{\prime\prime}(0,0)}|^{-2}(x\partial_{x})^{k^{\prime}}\partial_{t}f^{\prime\prime\prime} for 0kk0\leq k^{\prime}\leq k; and Ck,0C_{k,0} depends on |f(0,0)|1(xx)kf|{f^{\prime\prime}(0,0)}|^{-1}(x\partial_{x})^{k^{\prime}}f^{\prime\prime}, |f(0,0)|2(xx)ktf|{f^{\prime\prime}(0,0)}|^{-2}(x\partial_{x})^{k^{\prime}}\partial_{t}f^{\prime\prime}, |f(0,0)|1(xx)kΓf|{f^{\prime\prime}(0,0)}|^{-1}(x\partial_{x})^{k^{\prime}}\Gamma f^{\prime\prime} and |f(0,0)|2(xx)ktΓf|{f^{\prime\prime}(0,0)}|^{-2}(x\partial_{x})^{k^{\prime}}\partial_{t}\Gamma f^{\prime\prime} for 0kk0\leq k^{\prime}\leq k. Without loss of generality, we may assume Ck,jCk,jC_{k^{\prime},j}\leq C_{k,j} for kkk^{\prime}\leq k and j=0,1j=0,1.

Assuming (4.60)–(4.63), we can improve the induction hypothesis (4.54) for hh as follows. Using (4.41), (4.42), (4.52) and (4.59), we have

|h(k0)|\displaystyle|{h^{(k_{0})}}| 0texp(tt(|r|+|s|+|q|δ5dt)|(RHS of (4.59))|dt(Ξ(t)λ0)Cδ50t|(RHS of (4.59))|dt.\displaystyle\leq\int_{0}^{t}\exp\left(\int_{t^{\prime}}^{t}(|{r}|+|{s}|+|{q}|\delta_{5}\,\mathrm{d}t^{\prime\prime}\right)|{\hbox{(RHS of \eqref{eq:HJE-h-der2})}}|\,\mathrm{d}t^{\prime}\lesssim\left(\frac{\Xi(t)}{\lambda_{0}}\right)^{C\delta_{5}}\int_{0}^{t}|{\hbox{(RHS of \eqref{eq:HJE-h-der2})}}|\,\mathrm{d}t^{\prime}.

By (4.58) for k<k0k<k_{0}, (4.60)–(4.63), as well as (4.56), we obatin

|(RHS of (4.59))|\displaystyle|{\hbox{(RHS of \eqref{eq:HJE-h-der2})}}| Ck0,0,Ck0,1,δ01λ0δ3(k0212)(γλ0(λ0)γλ0(Ξ)γλ0(Ξ)2+λ014(logλ0)21Ξ)Ξ˙\displaystyle\lesssim_{C_{k_{0},0},C_{k_{0},1},\delta_{0}^{-1}}\lambda_{0}^{\delta_{3}(k_{0}^{2}-\frac{1}{2})}\left(\frac{\gamma_{\lambda_{0}}(\lambda_{0})\gamma_{\lambda_{0}}^{\prime}(\Xi)}{\gamma_{\lambda_{0}}(\Xi)^{2}}+\lambda_{0}^{\frac{1}{4}}(\log\lambda_{0})^{2}\frac{1}{\Xi}\right)\dot{\Xi}
+m=1k01λ0δ3(m2+(k0m)2)(logλ0)2Ξ˙Ξ,\displaystyle\phantom{\lesssim_{C_{k_{0},0},C_{k_{0},1},\delta_{0}^{-1}}}+\sum_{m=1}^{k_{0}-1}\lambda_{0}^{\delta_{3}(m^{2}+(k_{0}-m)^{2})}(\log\lambda_{0})^{2}\frac{\dot{\Xi}}{\Xi},

where the last term arises from the contribution of the last sum in (4.59) (this sum is vacuous when k0=1k_{0}=1). Using m2+(k0m)2k022m^{2}+(k_{0}-m)^{2}\leq k_{0}^{2}-2 for 1mk011\leq m\leq k_{0}-1, we obtain

|h(k0)|\displaystyle|{h^{(k_{0})}}| Ck0,0,Ck0,1,δ01λ0δ3(k0214)(logλ0)2(Ξ(t)λ0)Cδ5(1γλ0(λ0)γλ0(Ξ(t))+logΞ(t)λ0).\displaystyle\lesssim_{C_{k_{0},0},C_{k_{0},1},\delta_{0}^{-1}}\lambda_{0}^{\delta_{3}(k_{0}^{2}-\frac{1}{4})}(\log\lambda_{0})^{2}\left(\frac{\Xi(t)}{\lambda_{0}}\right)^{C\delta_{5}}\left(1-\frac{\gamma_{\lambda_{0}}(\lambda_{0})}{\gamma_{\lambda_{0}}(\Xi(t))}+\log\frac{\Xi(t)}{\lambda_{0}}\right).

Taking δ5δ0δ3\delta_{5}\ll\delta_{0}\delta_{3} and requiring λ0Λ\lambda_{0}\geq\Lambda, where Λ\Lambda is sufficiently large, we obtain |h(k0)|λ0δ3k02|{h^{(k_{0})}}|\leq\lambda_{0}^{\delta_{3}k_{0}^{2}}, which is equivalent to (4.54) for k=k0k=k_{0}, as desired.

Next, we improve (4.55). We begin with the formula

xk0x2Φ(t,x)=(xpξp)xk0h+xk0(xpξp)(h1)+=1k01x(xpξp)xkh,\partial_{x}^{k_{0}}\partial_{x}^{2}\Phi(t,x)=\left(\frac{\partial_{x}p}{\partial_{\xi}p}\right)\partial_{x}^{k_{0}}h+\partial_{x}^{k_{0}}\left(\frac{\partial_{x}p}{\partial_{\xi}p}\right)(h-1)+\sum_{\ell=1}^{k_{0}-1}\partial_{x}^{\ell}\left(\frac{\partial_{x}p}{\partial_{\xi}p}\right)\partial_{x}^{k-\ell}h, (4.64)

where xp\partial_{x}p and ξp\partial_{\xi}p are evaluated at (t,x,ξ)=(t,x,xΦ(t,x))(t,x,\xi)=(t,x,\partial_{x}\Phi(t,x)). We will prove that, for 0kk00\leq k\leq k_{0},

|x0kϵ(λ0)kexp(kI(t))xk(xpξp)(t,X(t),Ξ(t))|Ck,0λ0δ3k2(x2Φ),\left|{x_{0}^{k}{\epsilon(\lambda_{0})^{k}}\exp(-kI(t))\partial_{x}^{k}\left(\frac{\partial_{x}p}{\partial_{\xi}p}\right)(t,X(t),\Xi(t))}\right|\leq C_{k,0}\lambda_{0}^{\delta_{3}k^{2}}(-\partial_{x}^{2}\Phi), (4.65)

where Ck,0C_{k,0} depends on |f(0,0)|1(xx)kΓf|{f^{\prime\prime}(0,0)}|^{-1}(x\partial_{x})^{k^{\prime}}\Gamma f^{\prime\prime} and |f(0,0)|2(xx)ktΓf|{f^{\prime\prime}(0,0)}|^{-2}(x\partial_{x})^{k^{\prime}}\partial_{t}\Gamma f^{\prime\prime} for 0kk0\leq k^{\prime}\leq k. Without loss of generality, we may assume that these constants are the same as those in (4.60)–(4.63). In the proof of (4.65), it is important that xk(xpξp)\partial_{x}^{k}(\frac{\partial_{x}p}{\partial_{\xi}p}) involves only xkΦ\partial_{x}^{k^{\prime}}\Phi with kk+1<k0+2k^{\prime}\leq k+1<k_{0}+2, so that we may apply the induction hypothesis (4.55). We postpone the details until later.

Assuming (4.65), and also using |h|δ5|{h}|\leq\delta_{5}, (4.57) for kk0k\leq k_{0} (the case k=k0k=k_{0} has just been established) and (4.68) for kk01k\leq k_{0}-1, we may estimate the RHS of (4.64) by

|xk0x2Φ(t,X(t)|δ0,Ck0,0λ0δ3k02(logλ0)2x0k0ϵ(λ0)k0exp(k0I)(x2Φ),|{\partial_{x}^{k_{0}}\partial_{x}^{2}\Phi(t,X(t)}|\lesssim_{\delta_{0},C_{k_{0},0}}\lambda_{0}^{\delta_{3}k_{0}^{2}}(\log\lambda_{0})^{2}x_{0}^{-k_{0}}{\epsilon(\lambda_{0})^{-k_{0}}}\exp(k_{0}I)(-\partial_{x}^{2}\Phi),

where used the simple inequality (k0)2+2k02(k_{0}-\ell)^{2}+\ell^{2}\leq k_{0}^{2}. Since the exponent δ3((k0+1)21)\delta_{3}((k_{0}+1)^{2}-1) on λ0\lambda_{0} in (4.55) is strictly greater than δ3k02\delta_{3}k_{0}^{2}, taking λ0\lambda_{0} large enough depending on δ0\delta_{0}, δ3\delta_{3} and Ck0,0C_{k_{0},0}, we obtain (4.55) for k=k0k=k_{0}.

Proof of (4.60)–(4.63), (4.65). In the proof, we fix some k1k\geq 1. We may assume that the bounds (4.54) and (4.55) are available for any \ell satisfying <k\ell<k.

Some preliminary computations. We claim that

|x2Φ(t,X(t))|\displaystyle|{\partial_{x}^{2}\Phi(t,X(t))}| Cλ0δ32x0ϵ0(λ0)exp(I(t))Ξ(t),\displaystyle\leq\frac{C\lambda_{0}^{\frac{\delta_{3}}{2}}}{x_{0}{\epsilon_{0}(\lambda_{0})}}\exp(I(t))\Xi(t), (4.66)
1X(t)\displaystyle\frac{1}{X(t)} Cλ0δ32x0ϵ(λ0)exp(I(t)),\displaystyle\leq\frac{C\lambda_{0}^{\frac{\delta_{3}}{2}}}{x_{0}{\epsilon(\lambda_{0})}}\exp(I(t)), (4.67)

where CC depends on γ\gamma. Note that (4.66) combined with (4.55) would lead to, for any k1\ell\leq k-1,

|+1xxΦ(t,X(t))|λ0δ3((+1)212)x0(+1)ϵ(λ0)(+1)exp((+1)I(t))Ξ(t).|{\partial^{\ell+1}_{x}\partial_{x}\Phi(t,X(t))}|\leq\lambda_{0}^{\delta_{3}((\ell+1)^{2}-\frac{1}{2})}x_{0}^{-(\ell+1)}{\epsilon(\lambda_{0})^{-(\ell+1)}}\exp((\ell+1)I(t))\Xi(t). (4.68)

Ignoring the powers of λ0δ3\lambda_{0}^{\delta_{3}}, (4.68) tells us that every derivative of xΦ\partial_{x}\Phi loses x01ϵ(λ0)1exp(I)x_{0}^{-1}\epsilon(\lambda_{0})^{-1}\exp(I), which is essentially μ\mu in (4.48); (4.67) says that the loss x1x^{-1} is more favorable than x01ϵ(λ0)1exp(I)x_{0}^{-1}\epsilon(\lambda_{0})^{-1}\exp(I) along characteristics.

Essential to the proofs of both (4.66) and (4.67) is the following consequence of (4.3), which is connected to the assumption (1.18): If 2λ0λ(t)Mλ02\lambda_{0}\leq\lambda(t)\leq M\lambda_{0}, we have

ϵ0(λ0)λ0γλ0(λ0)C(logλ0)2ϵ0(λ0)γ(λ0)Cϵ0(λ0)(logλ0)2τMC(logλ0)2Mγλ0(Mλ0),\frac{\epsilon_{0}(\lambda_{0})}{\lambda_{0}\gamma_{\lambda_{0}}^{\prime}(\lambda_{0})}\leq C\frac{(\log\lambda_{0})^{2}\epsilon_{0}(\lambda_{0})}{\gamma(\lambda_{0})}\leq C\epsilon_{0}(\lambda_{0})(\log\lambda_{0})^{2}\tau_{M^{\prime}}\leq C(\log\lambda_{0})^{2}\frac{M^{\prime}}{\gamma_{\lambda_{0}}(M^{\prime}\lambda_{0})}, (4.69)

where M=λ01λ(t)M^{\prime}=\lambda_{0}^{-1}\lambda(t).

To prove (4.66), we first use (4.33), |h|δ5|{h}|\leq\delta_{5}, (4.20) or (4.31), and (4.50) to estimate, for λ0Λ\lambda_{0}\geq\Lambda sufficiently large depending on |f(0,0)|1Γf|{f^{\prime\prime}(0,0)}|^{-1}\Gamma f^{\prime\prime},

x0ϵ0(λ0)|x2Φ(t,X(t))|Ξ(t)exp(I(t))\displaystyle\frac{x_{0}{\epsilon_{0}(\lambda_{0})}|{\partial_{x}^{2}\Phi(t,X(t))}|}{\Xi(t)\exp(I(t))} Cϵ0(λ0)Ξ(t)exp(I(t))xp(t,X(t),Ξ(t))ξp(t,X(t),Ξ(t))\displaystyle\leq\frac{C{\epsilon_{0}(\lambda_{0})}}{\Xi(t)\exp(I(t))}\frac{\partial_{x}p(t,X(t),\Xi(t))}{\partial_{\xi}p(t,X(t),\Xi(t))}
Cϵ0(λ0)(λ0Ξ(t))1δ4γλ0(Ξ(t))λ0γλ0(λ0).\displaystyle\leq C{\epsilon_{0}(\lambda_{0})}\left(\frac{\lambda_{0}}{\Xi(t)}\right)^{1-\delta_{4}}\frac{\gamma_{\lambda_{0}}(\Xi(t))}{\lambda_{0}\gamma_{\lambda_{0}}^{\prime}(\lambda_{0})}.

For Ξ(t)4λ0\Xi(t)\leq 4\lambda_{0}, (4.66) is obvious, and for Ξ(t)4λ0\Xi(t)\geq 4\lambda_{0}, we apply (4.69) and λ(t)Ξ(t)2λ(t)\lambda(t)\leq\Xi(t)\leq 2\lambda(t) (from Lemma 4.3 or 4.7).

Next, to prove (4.67), we use (4.20) or (4.31) and (4.50) to estimate

x0ϵ(λ0)X(t)exp(I(t))\displaystyle\frac{x_{0}{\epsilon(\lambda_{0})}}{X(t)\exp(I(t))} ϵ0(λ0)(Ξ(t)λ0)δ4γλ0(Ξ(t))γλ0(λ0).\displaystyle\lesssim{\epsilon_{0}(\lambda_{0})}\left(\frac{\Xi(t)}{\lambda_{0}}\right)^{\delta_{4}}\frac{\gamma_{\lambda_{0}}^{\prime}(\Xi(t))}{\gamma_{\lambda_{0}}^{\prime}(\lambda_{0})}.

Again, for Ξ(t)4λ0\Xi(t)\leq 4\lambda_{0}, (4.67) is obvious. For Ξ(t)4λ0\Xi(t)\geq 4\lambda_{0}, we apply (4.69) and λ(t)Ξ(t)2λ(t)\lambda(t)\leq\Xi(t)\leq 2\lambda(t) (from Lemma 4.3 or 4.7) and estimate

ϵ0(λ0)(Ξ(t)λ0)δ4γλ0(Ξ(t))γλ0(λ0)C(Ξ(t)λ0)δ4(logλ0)2Ξ(t)γλ0(Ξ(t))γλ0(Ξ(t))Cλ0δ4δ0(logλ0)2,\displaystyle{\epsilon_{0}(\lambda_{0})}\left(\frac{\Xi(t)}{\lambda_{0}}\right)^{\delta_{4}}\frac{\gamma_{\lambda_{0}}^{\prime}(\Xi(t))}{\gamma_{\lambda_{0}}^{\prime}(\lambda_{0})}\leq C\left(\frac{\Xi(t)}{\lambda_{0}}\right)^{\delta_{4}}(\log\lambda_{0})^{2}\frac{\Xi(t)\gamma_{\lambda_{0}}^{\prime}(\Xi(t))}{\gamma_{\lambda_{0}}(\Xi(t))}\leq C\lambda_{0}^{\frac{\delta_{4}}{\delta_{0}}}(\log\lambda_{0})^{2},

which is acceptable.

Derivatives of qq, rr, ss and AA. We start with some preparations. By Faà di Bruno’s formula,

xγλ0(xΦ)=𝔞:a1+2a2++a=!a1!1!a1a!!a(ξa1++aγλ0)(xΦ)b=1(x1+bΦ)ab.\partial_{x}^{\ell}\gamma_{\lambda_{0}}(\partial_{x}\Phi)=\sum_{\mathfrak{a}:a_{1}+2a_{2}+\cdots+\ell a_{\ell}=\ell}\frac{\ell!}{a_{1}!1!^{a_{1}}\cdots a_{\ell}!\ell!^{a_{\ell}}}(\partial_{\xi}^{a_{1}+\cdots+a_{\ell}}\gamma_{\lambda_{0}})(\partial_{x}\Phi)\prod_{b=1}^{\ell}(\partial_{x}^{1+b}\Phi)^{a_{b}}.

Observe that we see at most \ell derivatives falling on xΦ\partial_{x}\Phi, to which (4.68) applies as long as k\ell\leq k. Using also the ellipticity assumption for γλ0\gamma_{\lambda_{0}} and b=1(bab)22\sum_{b=1}^{\ell}(ba_{b})^{2}\leq\ell^{2}, we obtain, for any 0k0\leq\ell\leq k,

|x(γλ0(xΦ))|(t,x)=(t,X(t))|γ,λ0δ3(212)x0ϵ(λ0)exp(I)γλ0(Ξ).\left|{\left.\partial_{x}^{\ell}(\gamma_{\lambda_{0}}(\partial_{x}\Phi))\right|_{(t,x)=(t,X(t))}}\right|\lesssim_{\gamma,\ell}\lambda_{0}^{\delta_{3}(\ell^{2}-\frac{1}{2})}x_{0}^{-\ell}{\epsilon(\lambda_{0})^{-\ell}}\exp(\ell I)\gamma_{\lambda_{0}}(\Xi). (4.70)

Similarly, but using instead the ellipticity assumption for ξξγλ0\xi\partial_{\xi}\gamma_{\lambda_{0}}, we also obtain, for any 0k0\leq\ell\leq k,

|x(γλ0(xΦ))|(t,x)=(t,X(t))|\displaystyle\left|{\left.\partial_{x}^{\ell}(\gamma_{\lambda_{0}}^{\prime}(\partial_{x}\Phi))\right|_{(t,x)=(t,X(t))}}\right| γ,λ0δ3(212)x0ϵ(λ0)exp(I)γλ0(Ξ).\displaystyle\lesssim_{\gamma,\ell}\lambda_{0}^{\delta_{3}(\ell^{2}-\frac{1}{2})}x_{0}^{-\ell}{\epsilon(\lambda_{0})^{-\ell}}\exp(\ell I)\gamma_{\lambda_{0}}^{\prime}(\Xi). (4.71)

Next, we introduce the notation f~=(f(0,0))1f\tilde{f}=(-f^{\prime\prime}(0,0))^{-1}f. In what follows, we write C,1C_{\ell,1} (resp. C,0C_{\ell,0}) for a constant, that may vary from line to line, that depends on (xx)kf~(x\partial_{x})^{k^{\prime}}\tilde{f}^{\prime\prime\prime}, (xx)kΓf~(x\partial_{x})^{k^{\prime}}\Gamma\tilde{f}^{\prime\prime\prime} and (f(0,0))1(xx)ktf~(-f^{\prime\prime}(0,0))^{-1}(x\partial_{x})^{k^{\prime}}\partial_{t}\tilde{f}^{\prime\prime\prime} (resp. (xx)kf~(x\partial_{x})^{k^{\prime}}\tilde{f}^{\prime\prime}, (f(0,0))1(xx)ktf~(-f^{\prime\prime}(0,0))^{-1}(x\partial_{x})^{k^{\prime}}\partial_{t}\tilde{f}^{\prime\prime}, (xx)kΓf~(x\partial_{x})^{k^{\prime}}\Gamma\tilde{f}^{\prime\prime} and
(f(0,0))1(xx)ktΓf~(-f^{\prime\prime}(0,0))^{-1}(x\partial_{x})^{k^{\prime}}\partial_{t}\Gamma\tilde{f}^{\prime\prime}) for 0k0\leq k^{\prime}\leq\ell. Then thanks to (4.67), we easily have, for any 0\ell\geq 0,

|f~(+2)(t,X(t))|+|tf~(+2)(t,X(t))f(0,0)|+|Γf~(+2)(t,X(t))|+|tΓf~(+2)(t,X(t))f(0,0)|\displaystyle\left|{\tilde{f}^{(\ell+2)}(t,X(t))}\right|+\left|{\frac{\partial_{t}\tilde{f}^{(\ell+2)}(t,X(t))}{-f^{\prime\prime}(0,0)}}\right|+\left|{\Gamma\tilde{f}^{(\ell+2)}(t,X(t))}\right|+\left|{\frac{\partial_{t}\Gamma\tilde{f}^{(\ell+2)}(t,X(t))}{-f^{\prime\prime}(0,0)}}\right| (4.72)
C,0λ0δ32x0ϵ(λ0)exp(I),\displaystyle\leq C_{\ell,0}\lambda_{0}^{\delta_{3}\frac{\ell}{2}}x_{0}^{-\ell}{\epsilon(\lambda_{0})^{-\ell}}\exp(\ell I),
|f~(+3)(t,X(t))|+|Γf~(+3)(t,X(t))|+|tf~(+3)(t,X(t))f(0,0)|C,1λ0δ32x0ϵ(λ0)exp(I).\left|{\tilde{f}^{(\ell+3)}(t,X(t))}\right|+\left|{\Gamma\tilde{f}^{(\ell+3)}(t,X(t))}\right|+\left|{\frac{\partial_{t}\tilde{f}^{(\ell+3)}(t,X(t))}{-f^{\prime\prime}(0,0)}}\right|\leq C_{\ell,1}\lambda_{0}^{\delta_{3}\frac{\ell}{2}}x_{0}^{-\ell}{\epsilon(\lambda_{0})^{-\ell}}\exp(\ell I). (4.73)

We are now ready to prove (4.60)–(4.63). We start with (4.62) for qq, which may be written as

q(t,x)=γλ0(xΦ)γλ0(xΦ)(f)(0,0)λ0γλ0(xΦ)(f~(t,x)+γλ0(xΦ)1Γf~(t,x)).q(t,x)=\frac{\gamma_{\lambda_{0}}^{\prime\prime}(\partial_{x}\Phi)}{\gamma_{\lambda_{0}}^{\prime}(\partial_{x}\Phi)}(-f^{\prime\prime})(0,0)\lambda_{0}\gamma_{\lambda_{0}}(\partial_{x}\Phi)\left(\tilde{f}^{\prime\prime}(t,x)+\gamma_{\lambda_{0}}(\partial_{x}\Phi)^{-1}\Gamma\tilde{f}^{\prime\prime}(t,x)\right).

By (4.71) and the ellipticity property of ξξγλ0\xi\partial_{\xi}\gamma_{\lambda_{0}}, we have, for any 0k0\leq\ell\leq k,

|x0ϵ(λ0)exp(I)xγλ0(xΦ)γλ0(xΦ)|(t,x)=(t,X(t))|γ,λ0δ3(212)1Ξ.\left|{x_{0}^{\ell}{\epsilon(\lambda_{0})^{\ell}}\exp(-\ell I)\left.\partial_{x}^{\ell}\frac{\gamma_{\lambda_{0}}^{\prime\prime}(\partial_{x}\Phi)}{\gamma_{\lambda_{0}}^{\prime}(\partial_{x}\Phi)}\right|_{(t,x)=(t,X(t))}}\right|\lesssim_{\gamma,\ell}\lambda_{0}^{\delta_{3}(\ell^{2}-\frac{1}{2})}\frac{1}{\Xi}.

By (4.70) and Ξ˙=xp\dot{\Xi}=-\partial_{x}p, we also see that

|x0ϵ(λ0)exp(I)x(f(0,0))λ0γλ0(xΦ)|(t,x)=(t,X(t))|γ,λ0δ3(212)Ξ˙.\left|{x_{0}^{\ell}{\epsilon(\lambda_{0})^{\ell}}\exp(-\ell I)\left.\partial_{x}^{\ell}(-f^{\prime\prime}(0,0))\lambda_{0}\gamma_{\lambda_{0}}(\partial_{x}\Phi)\right|_{(t,x)=(t,X(t))}}\right|\lesssim_{\gamma,\ell}\lambda_{0}^{\delta_{3}(\ell^{2}-\frac{1}{2})}\dot{\Xi}.

Derivatives of the last factor are easily bounded using (4.70) and (4.72). Putting together these bounds, (4.62) follows.

Next, we prove (4.63) for AA, which we write as

A(t,x)\displaystyle A(t,x) =ξxp(t,x,xΦ)q(t,x)(1h(t,x))\displaystyle=-\partial_{\xi}\partial_{x}p(t,x,\partial_{x}\Phi)-q(t,x)(1-h(t,x))
=f(0,0)λ0γλ0(xΦ)f~(t,x)q(t,x)(1h(t,x)).\displaystyle=-f^{\prime\prime}(0,0)\lambda_{0}\gamma_{\lambda_{0}}^{\prime}(\partial_{x}\Phi)\tilde{f}^{\prime\prime}(t,x)-q(t,x)(1-h(t,x)).

We now apply xk1\partial_{x}^{k-1} to the above expression. The contribution of q(1h)q(1-h) is already acceptable thanks to the previous bound for derivatives of qq, |h|δ5|{h}|\leq\delta_{5} and (4.57) for k1k-1 (indeed, note that we need to control only up to k1k-1 derivatives of hh). By (4.71), ellipticity of γλ0\gamma_{\lambda_{0}} and Ξ˙=xp\dot{\Xi}=-\partial_{x}p, we may also bound the contribution of the first term by Ck1,0λ0δ3((k1)212)Ξ1Ξ˙C_{k-1,0}\lambda_{0}^{\delta_{3}((k-1)^{2}-\frac{1}{2})}\Xi^{-1}\dot{\Xi}, which is acceptable.

To prove (4.61) for rr, we write

r(t,x)\displaystyle r(t,x) =x2pxpξp(t,x,xΦ)=(f(0,0))xγλ0(xΦ)f~(t,x)+γλ0(xΦ)1Γf~(t,x)f~(t,x)+γλ0(xΦ)1Γf~(t,x)f~(t,x)x,\displaystyle=\frac{\partial_{x}^{2}p}{\partial_{x}p}\partial_{\xi}p(t,x,\partial_{x}\Phi)=(-f^{\prime\prime}(0,0))x\gamma_{\lambda_{0}}^{\prime}(\partial_{x}\Phi)\frac{\tilde{f}^{\prime\prime\prime}(t,x)+\gamma_{\lambda_{0}}(\partial_{x}\Phi)^{-1}\Gamma\tilde{f}^{\prime\prime\prime}(t,x)}{\tilde{f}^{\prime\prime}(t,x)+\gamma_{\lambda_{0}}(\partial_{x}\Phi)^{-1}\Gamma\tilde{f}^{\prime\prime}(t,x)}\frac{\tilde{f}^{\prime}(t,x)}{x},

and apply xk\partial_{x}^{k} to the above expression. By Taylor expansion and (4.72), we have, for any 0\ell\geq 0,

|x0ϵ(λ0)exp(I)x(x1f~(t,x))|(t,x)=(t,X(t)|C,0λ0δ32.\left|{x_{0}^{\ell}{\epsilon(\lambda_{0})^{\ell}}\exp(-\ell I)\left.\partial_{x}^{\ell}\left(x^{-1}\tilde{f}^{\prime}(t,x)\right)\right|_{(t,x)=(t,X(t)}}\right|\lesssim_{\ell}C_{\ell,0}\lambda_{0}^{\delta_{3}\frac{\ell}{2}}.

On the other hand, using (4.20) or (4.31), (4.67), (4.70), (4.71), (4.72) and (4.73), it is straightforward to bound each factor appropriately and establish (4.61).

Finally, (4.60) for ss is proved by starting from the explicit form of ss given in the proof of Proposition 4.9, then applying Taylor expansion for (f)1tf(f)1tf(f^{\prime})^{-1}\partial_{t}f^{\prime}-(f^{\prime\prime})^{-1}\partial_{t}f^{\prime\prime}, (4.70), (4.72) and (4.73), as well as (4.20) or (4.31) and Ξ˙=xp\dot{\Xi}=-\partial_{x}p.

Proof of (4.65). We begin by writing

xpξp=f~(t,x)+γλ0(xΦ)1Γf~(t,x)x1f~(t,x)γλ0(xΦ)xγλ0(xΦ),\displaystyle\frac{\partial_{x}p}{\partial_{\xi}p}=\frac{\tilde{f}^{\prime\prime}(t,x)+\gamma_{\lambda_{0}}(\partial_{x}\Phi)^{-1}\Gamma\tilde{f}^{\prime\prime}(t,x)}{x^{-1}\tilde{f}^{\prime}(t,x)}\frac{\gamma_{\lambda_{0}}(\partial_{x}\Phi)}{x\gamma_{\lambda_{0}}^{\prime}(\partial_{x}\Phi)},

and applying xk\partial_{x}^{k} to the above expression. By (4.67), (4.70), (4.71) and (4.72) (as well as a Taylor-expansion argument for x1fx^{-1}f^{\prime} as before), we have

|x0kϵ(λ0)kexp(kI)xk(xpξp)(t,X(t),Ξ(t))|Ck,0λ0δ3(k212)γλ0(xΦ)xγλ0(xΦ).\displaystyle\left|{x_{0}^{k}{\epsilon(\lambda_{0})^{k}}\exp(-kI)\partial_{x}^{k}\left(\frac{\partial_{x}p}{\partial_{\xi}p}\right)(t,X(t),\Xi(t))}\right|\leq C_{k,0}\lambda_{0}^{\delta_{3}(k^{2}-\frac{1}{2})}\frac{\gamma_{\lambda_{0}}(\partial_{x}\Phi)}{x\gamma_{\lambda_{0}}^{\prime}(\partial_{x}\Phi)}.

Then by the relation x2Φ=xpξp(1h)\partial_{x}^{2}\Phi=-\frac{\partial_{x}p}{\partial_{\xi}p}(1-h) and |h|δ5|{h}|\leq\delta_{5} and the above expression for xpξp\frac{\partial_{x}p}{\partial_{\xi}p}, (4.65) follows. ∎

4.5 Estimates for the amplitude

In this section, we obtain L2L^{2}-estimates for the amplitude function aa, working under the same hypotheses as in Section 4.4. First, we rewrite the amplitude equation (3.16), incorporating the time dependence of ff and using the notations and conventions in this section:

ta+ξp(t,x,xΦ)xa+(12ξ2p(t,x,xΦ)2xΦ+s(t,x,xΦ))a=0.\begin{split}\partial_{t}a+\partial_{\xi}p(t,x,\partial_{x}\Phi)\partial_{x}a+\left(\frac{1}{2}\partial_{\xi}^{2}p(t,x,\partial_{x}\Phi)\partial^{2}_{x}\Phi+s(t,x,\partial_{x}\Phi)\right)a=0.\end{split} (4.74)

Moreover, we recall that

s(t,x,ξ)=12f(t,x)ξγλ0(ξ)λ012Γf(t,x)ξγλ0(ξ)γλ0(ξ)λ0=12xξp(t,x,ξ)12Γf(t,x)ξγλ0(ξ)γλ0(ξ)λ0.\begin{split}s(t,x,\xi)&=\frac{1}{2}f^{\prime\prime}(t,x)\partial_{\xi}\gamma_{\lambda_{0}}(\xi)\lambda_{0}-\frac{1}{2}\Gamma f^{\prime\prime}(t,x)\frac{\partial_{\xi}\gamma_{\lambda_{0}}(\xi)}{\gamma_{\lambda_{0}}(\xi)}\lambda_{0}\\ &=\frac{1}{2}\partial_{x}\partial_{\xi}p(t,x,\xi)-\frac{1}{2}\Gamma f^{\prime\prime}(t,x)\frac{\partial_{\xi}\gamma_{\lambda_{0}}(\xi)}{\gamma_{\lambda_{0}}(\xi)}\lambda_{0}.\end{split} (4.75)

It will be convenient to define

~t=t+ξp(x,xΦ)x+12(xξp(t,x,xΦ)+ξ2p(t,x,xΦ)x2Φ)\widetilde{\nabla}_{t}=\partial_{t}+\partial_{\xi}p(x,\partial_{x}\Phi)\partial_{x}+\frac{1}{2}\left(\partial_{x}\partial_{\xi}p(t,x,\partial_{x}\Phi)+\partial_{\xi}^{2}p(t,x,\partial_{x}\Phi)\partial_{x}^{2}\Phi\right)

so that solutions to ~ta=0\widetilde{\nabla}_{t}a=0 obey a-priori L2L^{2}-bounds. Then, (4.74) is simply given by

~ta=12Γfλ0ξγλ0(xΦ)γλ0(xΦ)a.\begin{split}\widetilde{\nabla}_{t}a=\frac{1}{2}\Gamma f^{\prime\prime}\frac{\lambda_{0}\partial_{\xi}\gamma_{\lambda_{0}}(\partial_{x}\Phi)}{\gamma_{\lambda_{0}}(\partial_{x}\Phi)}a.\end{split}

Commutation with x\partial_{x} again gives rise to a similar factor as before: recalling (4.43), we compute that

[x,~t]=[x,t]+12[x,xξp(x,xΦ)+ξ2p(x,xΦ)x2Φ]=Ax12xA.\begin{split}[\partial_{x},\widetilde{\nabla}_{t}]&=[\partial_{x},\nabla_{t}]+\frac{1}{2}[\partial_{x},\partial_{x}\partial_{\xi}p(x,\partial_{x}\Phi)+\partial_{\xi}^{2}p(x,\partial_{x}\Phi)\partial_{x}^{2}\Phi]{=-A\partial_{x}-\frac{1}{2}\partial_{x}A}.\end{split}

For simplicity, we also set

H(t,x):=12Γf(t,x)γλ0(xΦ(t,x))γλ0(xΦ(t,x))λ0\begin{split}H(t,x):=-\frac{1}{2}\Gamma f^{\prime\prime}(t,x)\frac{\gamma_{\lambda_{0}}^{\prime}(\partial_{x}\Phi(t,x))}{\gamma_{\lambda_{0}}(\partial_{x}\Phi(t,x))}\lambda_{0}\end{split}

so that the equations for aa, xa\partial_{x}a, and x2a\partial_{x}^{2}a read

~ta=Ha,\begin{split}\widetilde{\nabla}_{t}a=Ha,\end{split}
~txa=[t,x]a+12xAa+x(Ha),\begin{split}\widetilde{\nabla}_{t}\partial_{x}a=[\nabla_{t},\partial_{x}]a+{\frac{1}{2}\partial_{x}A}a+\partial_{x}(Ha),\end{split}

and

~tx2a=2[t,x]xa+32xAxa+12x((xA)a)+x2(Ha).\begin{split}\widetilde{\nabla}_{t}\partial_{x}^{2}a=2[\nabla_{t},\partial_{x}]\partial_{x}a+{\frac{3}{2}\partial_{x}A}\partial_{x}a+{\frac{1}{2}\partial_{x}((\partial_{x}A)a)}+\partial_{x}^{2}(Ha).\end{split}
Proposition 4.12.

Assume that a0(x)a_{0}(x) is sufficiently smooth and supported in (x0,x1)(x_{0},x_{1}). For the solution of (4.74) with initial data a(t=0)=a0a(t=0)=a_{0}, we have, for 0kN00\leq k\leq N_{0},

μkxkaL2=0k(μ01x)a0L2 for tτ\|{\mu^{-k}\partial_{x}^{k}a}\|_{L^{2}}\lesssim\sum_{\ell=0}^{k}\|{({\mu_{0}^{-1}}\partial_{x})^{\ell}a_{0}}\|_{L^{2}}\qquad\hbox{ for }t\leq\tau

where

μ0=(logλ0)2x0ϵ(λ0),\mu_{0}=\frac{(\log\lambda_{0})^{2}}{x_{0}{\epsilon(\lambda_{0})}}, (4.76)

and the implicit constant depends on kk and γ\gamma, as well as |f(0,0)|1(xx)kf|{f^{\prime\prime}(0,0)}|^{-1}(x\partial_{x})^{k^{\prime}}f^{\prime\prime}, |f(0,0)|1(xx)kΓf|{f^{\prime\prime}(0,0)}|^{-1}(x\partial_{x})^{k^{\prime}}\Gamma f^{\prime\prime} for 0kk0\leq k^{\prime}\leq k.

Proof.

In what follows, the symbol k\lesssim_{k} signifies an implicit constant that may depend on kk and γ\gamma, as well as |f(0,0)|1(xx)kf|{f^{\prime\prime}(0,0)}|^{-1}(x\partial_{x})^{k^{\prime}}f^{\prime\prime}, |f(0,0)|1(xx)kΓf|{f^{\prime\prime}(0,0)}|^{-1}(x\partial_{x})^{k^{\prime}}\Gamma f^{\prime\prime} for 0kk0\leq k^{\prime}\leq k.

To begin with, we observe the following form of the equation for xka\partial_{x}^{k}a:

~txka=k[t,x]xk1a+R~k=kAxka+R~k,\begin{split}\widetilde{\nabla}_{t}\partial_{x}^{k}a=k[\nabla_{t},\partial_{x}]\partial_{x}^{k-1}a+\widetilde{R}_{k}=kA\partial_{x}^{k}a+\widetilde{R}_{k},\end{split}

where R~k\widetilde{R}_{k} is a linear combination of the terms

xAxka,xHxka\begin{split}{\partial_{x}^{\ell}A\partial_{x}^{k-\ell}a,\quad\partial_{x}^{\ell}H\partial_{x}^{k-\ell}a}\end{split}

for 0k0\leq\ell\leq k, with the exception that the term with =0\ell=0 does not appear in the case of AA. It will be convenient to first solve the following transport equation:

tI=A,\begin{split}\nabla_{t}I=A,\end{split}

with the initial data I(t=0)=0I(t=0)=0. The solution, which is simply the integral in time of AA along each characteristic curve, is nothing but II defined in (4.49). Since ~tt\widetilde{\nabla}_{t}-\nabla_{t} is simply a multiplication operator, we have that

~t(exp(kI(t,x))xka(t,x))=exp(kI(t,x))R~k.\begin{split}\widetilde{\nabla}_{t}\left(\exp(-kI(t,x))\partial_{x}^{k}a(t,x)\right)=\exp(-kI(t,x))\widetilde{R}_{k}.\end{split} (4.77)

We now set up the induction hypothesis: For 0kk00\leq k\leq k_{0}, we shall require

sup0tτexp(kI(t,x))xka(t,x)L2kx0kϵ(λ0)kλ0δ3k20k(x0ϵ(λ0)x)ka0L2.\begin{split}\sup_{0\leq t\leq\tau}\|{\exp(-kI(t,x))\partial_{x}^{k}a(t,x)}\|_{L^{2}}\lesssim_{k}x_{0}^{-k}{\epsilon(\lambda_{0})^{-k}}\lambda_{0}^{\delta_{3}k^{2}}\sum_{0\leq\ell\leq k}\|{(x_{0}{\epsilon(\lambda_{0})}\partial_{x})^{k}a_{0}}\|_{L^{2}}.\end{split} (4.78)

We first need to check the above for k=0k=0. Recall the definition of HH. By Lemma 4.3 or 4.7 and the definition of λ˙\dot{\lambda}, we have the obvious bound

|H|γλ0(λ)γλ0(λ)2λ˙,|{H}|\lesssim\frac{\gamma_{\lambda_{0}}^{\prime}(\lambda)}{\gamma_{\lambda_{0}}(\lambda)^{2}}\dot{\lambda},

Then from the equation for aa, we have, by Grönwall’s inequality,

sup0tτa(t)L2a0L2.\begin{split}\sup_{0\leq t\leq\tau}\|{a(t)}\|_{L^{2}}\lesssim\|{a_{0}}\|_{L^{2}}.\end{split}

Let us now estimate exp(kI(t,x))xka(t,x)\exp(-kI(t,x))\partial_{x}^{k}a(t,x) in L2L^{2} for some k>0k>0, using (4.78) for all <k\ell<k. Recalling the bound (4.63) for xkA\partial_{x}^{k}A along characteristics, we obtain the uniform bound (k>0)(k>0)

|xkA|kx0kϵ(λ0)kΞ˙Ξλ0δ3(k212)exp(kI)kx0kϵ(λ0)kλ0δ3(k212)exp(kI)λ˙λ\begin{split}|\partial_{x}^{k}A|\lesssim_{k}x_{0}^{-k}{\epsilon(\lambda_{0})^{-k}}\frac{\dot{\Xi}}{\Xi}\lambda_{0}^{\delta_{3}(k^{2}-\frac{1}{2})}\exp(kI)\lesssim_{k}x_{0}^{-k}{\epsilon(\lambda_{0})^{-k}}\lambda_{0}^{\delta_{3}(k^{2}-\frac{1}{2})}\exp(kI)\frac{\dot{\lambda}}{\lambda}\end{split}

using simply that Ξ˙ΞCλ˙λ\frac{\dot{\Xi}}{\Xi}\leq C\frac{\dot{\lambda}}{\lambda}. This allows us to bound

exp(kI)xAxkaL2kx0ϵ(λ0)λ0δ3(214)λ˙λexp((k)I)xkaL2.\begin{split}\|{\exp(-kI)\partial_{x}^{\ell}A\partial_{x}^{k-\ell}a}\|_{L^{2}}\lesssim_{k}x_{0}^{-\ell}{\epsilon(\lambda_{0})^{-\ell}}\lambda_{0}^{\delta_{3}(\ell^{2}-\frac{1}{4})}\frac{\dot{\lambda}}{\lambda}\|{\exp(-(k-\ell)I)\partial_{x}^{k-\ell}a}\|_{L^{2}}.\end{split}

Finally, we need to estimate derivatives of HH. We claim that for k>0k>0,

|xkH|kx0kϵ(λ0)kexp(kI)λ0δ3(k214)γλ0(λ)γλ0(λ)2λ˙.\begin{split}|\partial_{x}^{k}H|\lesssim_{k}x_{0}^{-k}{\epsilon(\lambda_{0})^{-k}}\exp(kI)\lambda_{0}^{\delta_{3}(k^{2}-\frac{1}{4})}{\frac{\gamma_{\lambda_{0}}^{\prime}(\lambda)}{\gamma_{\lambda_{0}}(\lambda)^{2}}\dot{\lambda}}.\end{split}

We omit the proof, which can be done along the lines of the proof of (4.60)–(4.63). Applying these bounds to (4.77) and taking the L2L^{2} inner product with exp(kI(t,x))xka(t,x)\exp(-kI(t,x))\partial_{x}^{k}a(t,x), we obtain that

ddtexp(kI(t,x))xka(t,x)L2k=1kx0ϵ(λ0)λ0δ3(214)(λ˙λ+γλ0(λ)λ˙γλ0(λ)2)exp((k)I)xkaL2+γλ0(λ)λ˙γλ0(λ)2exp(kI(t,x))xka(t,x)L2kx0kϵ(λ0)kλ0δ3(k214)(λ˙λ+γλ0(λ)λ˙γλ0(λ)2)=0k1(x0ϵ(λ0)x)a0L2+γλ0(λ)λ˙γλ0(λ)2exp(kI(t,x))xka(t,x)L2.\begin{split}\frac{\mathrm{d}}{\mathrm{d}t}\|{\exp(-kI(t,x))\partial_{x}^{k}a(t,x)}\|_{L^{2}}&\lesssim_{k}\sum_{\ell={1}}^{k}x_{0}^{-\ell}{\epsilon(\lambda_{0})^{-\ell}}\lambda_{0}^{\delta_{3}(\ell^{2}-\frac{1}{4})}{\left(\frac{\dot{\lambda}}{\lambda}+\frac{\gamma_{\lambda_{0}}^{\prime}(\lambda)\dot{\lambda}}{\gamma_{\lambda_{0}}(\lambda)^{2}}\right)}\|{\exp(-(k-\ell)I)\partial_{x}^{k-\ell}a}\|_{L^{2}}\\ &\mathrel{\phantom{\lesssim}}+{\frac{\gamma_{\lambda_{0}}^{\prime}(\lambda)\dot{\lambda}}{\gamma_{\lambda_{0}}(\lambda)^{2}}}\|{\exp(-kI(t,x))\partial_{x}^{k}a(t,x)}\|_{L^{2}}\\ &\lesssim_{k}x_{0}^{-k}{\epsilon(\lambda_{0})^{-k}}\lambda_{0}^{\delta_{3}(k^{2}-\frac{1}{4})}{\left(\frac{\dot{\lambda}}{\lambda}+\frac{\gamma_{\lambda_{0}}^{\prime}(\lambda)\dot{\lambda}}{\gamma_{\lambda_{0}}(\lambda)^{2}}\right)}\sum_{\ell=0}^{{k-1}}\|{(x_{0}{\epsilon(\lambda_{0})}\partial_{x})^{\ell}a_{0}}\|_{L^{2}}\\ &\mathrel{\phantom{\lesssim}}+{\frac{\gamma_{\lambda_{0}}^{\prime}(\lambda)\dot{\lambda}}{\gamma_{\lambda_{0}}(\lambda)^{2}}}\|{\exp(-kI(t,x))\partial_{x}^{k}a(t,x)}\|_{L^{2}}.\end{split}

Applying Grönwall’s inequality finishes the proof of (4.78). Recalling the definitions of μ\mu and μ0\mu_{0} from (4.48) and (4.76), respectively, this completes the proof of the proposition (by taking Λ\Lambda larger if necessary). ∎

4.6 Cutoff and extension of the phase function

In this section, we shall extend Φ\Phi constructed in the previous subsections globally in space. For this purpose, we take points x0=2x0+x13x_{0}^{\prime}=\frac{2x_{0}+x_{1}}{3} and x1=x0+2x13x_{1}^{\prime}=\frac{x_{0}+2x_{1}}{3} so that x0<x0<x1<x1x_{0}<x_{0}^{\prime}<x_{1}^{\prime}<x_{1} and let χ(x0,x1)\chi_{(x_{0},x_{1})} be a smooth cutoff supported in (x0,x1)(x_{0},x_{1}) that equals 11 on (x0,x1)(x_{0}^{\prime},x_{1}^{\prime}). Then the support of 1χ(x0,x1)1-\chi_{(x_{0},x_{1})} has two components; we denote by χ(,x0)\chi_{(-\infty,x_{0})} and χ(x1,)\chi_{(x_{1},\infty)} the smooth cutoffs supported in the left and the right components, respectively, such that 1χ(x0,x1)=χ(,x0)+χ(x1,)1-\chi_{(x_{0},x_{1})}=\chi_{(-\infty,x_{0})}+\chi_{(x_{1},\infty)}. For x¯(x0,x1)\underline{x}\in(x_{0},x_{1}), we define

xΦglobal(t,X(t;x¯))\displaystyle\partial_{x}\Phi^{global}(t,X(t;\underline{x})) =χ(x0,x1)(x¯)xΦ(t,X(t;x¯))\displaystyle=\chi_{(x_{0},x_{1})}(\underline{x})\partial_{x}\Phi(t,X(t;\underline{x})) (4.79)
+χ(,x0)(x¯)xΦ(t,X(t;x0))+χ(x1,)(x¯)xΦ(t,X(t;x1)),\displaystyle\mathrel{\phantom{=}}+\chi_{(-\infty,x_{0})}(\underline{x})\partial_{x}\Phi(t,X(t;x_{0}))+\chi_{(x_{1},\infty)}(\underline{x})\partial_{x}\Phi(t,X(t;x_{1})),

where X(t;x¯)X(t;\underline{x}) is the characteristic curve solving X˙=ξp(t,X,xΦ(t,X))\dot{X}=\partial_{\xi}p(t,X,\partial_{x}\Phi(t,X)) with X(0;x¯)=x¯X(0;\underline{x})=\underline{x}. We furthermore normalize

Φglobal(t,X(t;x0+x12))=Φ(t,X(t;x0+x12)).\Phi^{global}(t,X(t;\tfrac{x^{\prime}_{0}+x^{\prime}_{1}}{2}))=\Phi(t,X(t;\tfrac{x^{\prime}_{0}+x^{\prime}_{1}}{2})).

Finally, for xx outside of the image of X(t;(x0,x1))X(t;(x_{0},x_{1})), we extend xΦglobal(t,x)\partial_{x}\Phi^{global}(t,x) (smoothly) by constants.

By definition,

Φglobal=Φ(t,x) for X(t,x0)xX(t,x1).\Phi^{global}=\Phi(t,x)\qquad\hbox{ for }X(t,x_{0}^{\prime})\leq x\leq X(t,x_{1}^{\prime}).

To continue, for each tt, denote by x¯(t,x)\underline{x}(t,x) the inverse of the map x¯X(t;x¯)\underline{x}\mapsto X(t;\underline{x}). Since

tx¯(t,x)=0,x¯(0,x)=x,\nabla_{t}\underline{x}(t,x)=0,\qquad\underline{x}(0,x)=x,

by an argument similar to Section 4.4 using (4.63), we have

|xkx¯(t,X(t;x¯))|kμk.|{\partial_{x}^{k}\underline{x}(t,X(t;\underline{x}))}|\lesssim_{k}\mu^{k}.

It follows that

λxΦglobal2λ,|xkxΦglobal|kμkλ for 1kN.\lambda\leq\partial_{x}\Phi^{global}\leq 2\lambda,\qquad|{\partial_{x}^{k}\partial_{x}\Phi^{global}}|\lesssim_{k}\mu^{k}\lambda\qquad\hbox{ for }1\leq k\leq N.

In Sections 6 and 7, we shall write Φ\Phi for Φglobal\Phi^{global}. This abuse of notation is minor, since our wave packet would be of the form Re(aeiΦ)\operatorname{Re}(ae^{i\Phi}), where a(t,)a(t,\cdot) is supported in (X(t,x0),X(t,x1))(X(t,x_{0}^{\prime}),X(t,x_{1}^{\prime})), on which Φglobal=Φ\Phi^{global}=\Phi.

5 Oscillatory integrals

The purpose of this section is to formulate and prove a result concerning the L2L^{2}-bound for the operators arising in the remainder in Proposition 3.3. Logically, this section is self-contained so the symbols in this section are detached from their meanings in the previous sections. On the other hand, the results of this section will be applied in Section 6 to objects that are suggested by the notation here.

To begin with, observe that each term in the remainder symbols (Φ)rp,1(x,ξ){}^{(\Phi)}r_{p,-1}(x,\xi) and (Φ)rp,2(x,ξ){}^{(\Phi)}r_{p,-2}(x,\xi) in (3.7)–(3.8) is of the form 01(Φ)q(x,ξ;σ)dσ\int_{0}^{1}{}^{(\Phi)}q(x,\xi;\sigma)\,\mathrm{d}\sigma, where

(Φ)q(x,ξ;σ):=q(x,ση+(1σ)yΦ(y))ei(Φ(y)Φ(x)+(ηξ)(xy))dydη(2π)2{}^{(\Phi)}q(x,\xi;\sigma):=\iint q(x,\sigma\eta+(1-\sigma)\partial_{y}\Phi(y))e^{i(\Phi(y)-\Phi(x)+(\eta-\xi)\cdot(x-y))}\,\mathrm{d}y\frac{\mathrm{d}\eta}{(2\pi)^{2}} (5.1)

for some symbol qq. Hence, it is expedient to formulate an L2L^{2}-bound result for an operator of the form (Φ)q(x,D;σ){}^{(\Phi)}q(x,D;\sigma) under suitable assumptions on qq.

Our main result is as follows:

Proposition 5.1.

Let λ0\lambda_{0}\in\mathbb{N}, λλ0\lambda\geq\lambda_{0}, μ>0\mu>0, λq>0\lambda_{q}>0 and N0N_{0}\in\mathbb{N}. Assume that the phase function Φ:x2\Phi:\mathbb{R}_{x_{2}}\to\mathbb{R} and the symbol q(x2,ξ2):x2×ξ2q(x_{2},\xi_{2}):\mathbb{R}_{x_{2}}\times\mathbb{R}_{\xi_{2}}\to\mathbb{C} satisfy, for any integer 0kN00\leq k\leq N_{0},

|x2kx2Φ(x)|\displaystyle|{\partial_{x_{2}}^{k}\partial_{x_{2}}\Phi(x)}| AΦ,kμkλ,\displaystyle\leq A_{\Phi,k}\mu^{k}\lambda, (5.2)
|ξ2kq(x2,ξ2)|\displaystyle|{\partial_{\xi_{2}}^{k}q(x_{2},\xi_{2})}| Aq,kλqkg(x2)M(ξ2),\displaystyle\leq A_{q,k}\lambda_{q}^{-k}g(x_{2})M(\xi_{2}), (5.3)

respectively, where AΦ,j,Aq,j>0A_{\Phi,j},A_{q,j}>0 are increasing. Furthermore, assume that λq,λ,μ\lambda_{q},\lambda,\mu satisfy

μλq1,λμλq21.\frac{\mu}{\lambda_{q}}\leq 1,\quad\frac{\lambda\mu}{\lambda_{q}^{2}}\leq 1. (5.4)

Then for any smooth a=a(x2)a=a(x_{2}), 0σ10\leq\sigma\leq 1, X0X_{0}\in\mathbb{R}, ss\in\mathbb{R} and max{104,100|s|}NN0\max\{10^{4},100|{s}|\}\leq N\leq N_{0}, we have the bound

((Φ)q(x2,D2;σ)a)μ(x2X0)sL2\displaystyle\left\|{\left({}^{(\Phi)}q(x_{2},D_{2};\sigma)a\right){\langle{\mu(x_{2}-X_{0})}\rangle^{s}}}\right\|_{L^{2}} (5.5)
s,N,AΦ,N,Aq,N=0N(g¯<μ1(x2)¯<μ1(x2;σ)(μ1x2)a(x2))μ(x2X0)sL2,\displaystyle\lesssim_{s,N,A_{\Phi,N},A_{q,N}}\sum_{\ell=0}^{N}\left\|{\left(\overline{g}_{<\mu^{-1}}(x_{2})\overline{\mathcal{M}}_{<\mu^{-1}}(x_{2};\sigma)(\mu^{-1}\partial_{x_{2}})^{\ell}a(x_{2})\right){\langle{\mu(x_{2}-X_{0})}\rangle^{s}}}\right\|_{L^{2}},

where

g¯<μ1(x2)\displaystyle\overline{g}_{<\mu^{-1}}(x_{2}) =supy2g(y2)μ(x2y2)N100,\displaystyle=\sup_{y_{2}}\frac{g(y_{2})}{\langle{\mu(x_{2}-y_{2})}\rangle^{\frac{N}{100}}},
¯<μ1(x2;σ)\displaystyle\overline{\mathcal{M}}_{<\mu^{-1}}(x_{2};\sigma) =supy2<λq1,<λq(y2,ξ2;σ)μ(x2y2)N100μ1μ1ξ2N100dξ2,\displaystyle=\sup_{y_{2}}\int\frac{\mathcal{M}_{<\lambda_{q}^{-1},<\lambda_{q}}(y_{2},\xi_{2};\sigma)}{\langle{\mu(x_{2}-y_{2})}\rangle^{\frac{N}{100}}}\frac{\mu^{-1}}{\langle{\mu^{-1}\xi_{2}}\rangle^{\frac{N}{100}}}\,\mathrm{d}\xi_{2},

and

<λq1,<λq(x2,ξ2;σ)\displaystyle\mathcal{M}_{<\lambda_{q}^{-1},<\lambda_{q}}(x_{2},\xi_{2};\sigma)
=M(σ(ζ+ξ2+x2Φ(x)x2Φ(x1,x2+z)12x22Φ(x)z)+x2Φ(x1,x2+z))\displaystyle=\iint M(\sigma(\zeta+\xi_{2}+\partial_{x_{2}}\Phi(x)-\partial_{x_{2}}\Phi(x_{1},x_{2}+z)-\tfrac{1}{2}\partial_{x_{2}}^{2}\Phi(x)z)+\partial_{x_{2}}\Phi(x_{1},x_{2}+z))
×λqλqzN100λq1λq1ζN100dζdz.\displaystyle\phantom{=\iint}\times\frac{\lambda_{q}}{\langle{\lambda_{q}z}\rangle^{\frac{N}{100}}}\frac{\lambda_{q}^{-1}}{\langle{\lambda_{q}^{-1}\zeta}\rangle^{\frac{N}{100}}}\,\mathrm{d}\zeta\mathrm{d}z.
Remark 5.2.

Note that the RHS of (5.5) involves a spatial weight g¯<μ1(x2)¯<μ1(x2;σ)\overline{g}_{<\mu^{-1}}(x_{2})\overline{\mathcal{M}}_{<\mu^{-1}}(x_{2};\sigma); this feature is important for exploiting the degeneracy of the principal symbol pθ̊p_{\mathring{\theta}} in the proof of the error bound in Section 6.3. We also point out that in (5.5), we are allowed to lose as many derivatives of aa (with weights in μ1\mu^{-1}) as we wish; this feature simplifies the proof (see Remark 5.4).

5.1 Oscillatory integral bounds for the symbol

In this section, we aim to derive key pointwise bounds for the symbol (Φ)q(x,ξ;σ){}^{(\Phi)}q(x,\xi;\sigma) and its ξ\xi-derivatives.

The main goal of this subsection is to prove the following pointwise bound for the symbol (Φ)q(x2,ξ2;σ){}^{(\Phi)}q(x_{2},\xi_{2};\sigma):

Lemma 5.3.

For any x1x_{1}\in\mathbb{R}, σ[0,1]\sigma\in[0,1], ,k0\ell,k\in\mathbb{N}_{0} and C0C_{0}\in\mathbb{N} such that +10C0N0\ell+10C_{0}\leq N_{0}, we have

|ξ2(Φ)q(x2,ξ2;σ)|\displaystyle|{\partial_{\xi_{2}}^{\ell}{}^{(\Phi)}q(x_{2},\xi_{2};\sigma)}| ,AΦ,10C0,Aq,+10C0λqg(x2)<λq1,<λq(x2,ξ2;σ)max{1,μλq1,λλq(μλq1+μ3λq3)}C0+2,\displaystyle\lesssim_{\ell,A_{\Phi,10C_{0}},A_{q,\ell+10C_{0}}}\lambda_{q}^{-\ell}g(x_{2})\mathcal{M}_{<\lambda_{q}^{-1},<\lambda_{q}}(x_{2},\xi_{2};\sigma)\max\left\{1,\mu\lambda_{q}^{-1},\frac{\lambda}{\lambda_{q}}\left(\mu\lambda_{q}^{-1}+\mu^{3}\lambda_{q}^{-3}\right)\right\}^{C_{0}+2}, (5.6)

where

<λq1,<λq(x2,ξ2;σ)\displaystyle\mathcal{M}_{<\lambda_{q}^{-1},<\lambda_{q}}(x_{2},\xi_{2};\sigma)
=M(σ(ζ+ξ2+x2Φ(x2)x2Φ(x2+z)12x22Φ(x2)z)+x2Φ(x2+z))\displaystyle=\iint M(\sigma(\zeta+\xi_{2}+\partial_{x_{2}}\Phi(x_{2})-\partial_{x_{2}}\Phi(x_{2}+z)-\tfrac{1}{2}\partial_{x_{2}}^{2}\Phi(x_{2})z)+\partial_{x_{2}}\Phi(x_{2}+z))
×λqλqzC0λq1λq1ζC0dζdz.\displaystyle\phantom{=\iint}\times\frac{\lambda_{q}}{\langle{\lambda_{q}z}\rangle^{C_{0}}}\frac{\lambda_{q}^{-1}}{\langle{\lambda_{q}^{-1}\zeta}\rangle^{C_{0}}}\,\mathrm{d}\zeta\mathrm{d}z.
Proof.

We begin by laying out some notational simplifications to be used in the proof. To simplify the notation, we shall write x=x2x=x_{2}, ξ=ξ2\xi=\xi_{2}, q(ξ)=q(x,ξ)q(\xi)=q(x,\xi) and Q(x,ξ)=(Φ)q(x,ξ;σ)Q(x,\xi)={}^{(\Phi)}q(x,\xi;\sigma). Then

Q(x,ξ)=q(ση+(1σ)xΦ(y))ei(Φ(y)Φ(x)+(ηξ)(xy))dydη2π.Q(x,\xi)=\iint q(\sigma\eta+(1-\sigma)\partial_{x}\Phi(y))e^{i(\Phi(y)-\Phi(x)+(\eta-\xi)(x-y))}\,\mathrm{d}y\frac{\mathrm{d}\eta}{2\pi}.

Moreover, the assumption (5.3) becomes

|ξnq(ξ)|n,Aq,nλqng(x)M(ξ).|{\partial_{\xi}^{n}q(\xi)}|\lesssim_{n,A_{q,n}}\lambda_{q}^{-n}g(x)M(\xi). (5.7)

Finally, by the second assumption on the phase, we have |xnxΦ(x)|n,AΦ,nμnλ|{\partial_{x}^{n}\partial_{x}\Phi(x)}|\lesssim_{n,A_{\Phi,n}}\mu^{n}\lambda.

In what follows, we first give a detailed proof of the case =0\ell=0, then indicate the necessary modifications for higher \ell’s.

The phase function ψ(x,ξ;y,η)\psi(x,\xi;y,\eta), defined by

ψ(x,ξ;y,η)=Φ(y)Φ(x)+(ηξ)(xy),\displaystyle\psi(x,\xi;y,\eta)=\Phi(y)-\Phi(x)+(\eta-\xi)(x-y),

is stationary at y=xy=x and η=ξ+xΦ(x)\eta=\xi+\partial_{x}\Phi(x). By Taylor expansion, we may write

ψ(x,ξ;y,η)\displaystyle\psi(x,\xi;y,\eta) =xΦ(x)(yx)+12x2Φ(x)(yx)2+(ηξ)(xy)+13!r(3)[Φ](y,x)(yx)3\displaystyle=\partial_{x}\Phi(x)(y-x)+\tfrac{1}{2}\partial_{x}^{2}\Phi(x)(y-x)^{2}+(\eta-\xi)(x-y)+\frac{1}{3!}r^{(3)}[\Phi](y,x)(y-x)^{3}
=(xy)(ηξxΦ(x)+12x2Φ(x)(xy))+13!r(3)[Φ](y,x)(yx)3.\displaystyle=(x-y)(\eta-\xi-\partial_{x}\Phi(x)+\tfrac{1}{2}\partial_{x}^{2}\Phi(x)(x-y))+\frac{1}{3!}r^{(3)}[\Phi](y,x)(y-x)^{3}.

We make the change of variables (y,η)(z,ζ)(y,\eta)\mapsto(z,\zeta), where

z\displaystyle z =yx,\displaystyle=y-x,
ζ\displaystyle\zeta =ηξxΦ(x)+12x2Φ(x)(xy),\displaystyle=\eta-\xi-\partial_{x}\Phi(x)+\tfrac{1}{2}\partial_{x}^{2}\Phi(x)(x-y),

so that

Q(x,ξ)\displaystyle Q(x,\xi) =q~(x,ξ;z,ζ)eir(3)[Φ](x+z,x)z33!eiζzdζdz,\displaystyle=\iint\widetilde{q}(x,\xi;z,\zeta)e^{ir^{(3)}[\Phi](x+z,x)\frac{z^{3}}{3!}}e^{-i\zeta z}\,\mathrm{d}\zeta\mathrm{d}z, (5.8)

where

q~(x,ξ;z,ζ)=12πq(σ(ζ+ξ+xΦ(x)xΦ(x+z)12x2Φ(x)z)+xΦ(x+z)).\widetilde{q}(x,\xi;z,\zeta)=\frac{1}{2\pi}q\left(\sigma(\zeta+\xi+\partial_{x}\Phi(x)-\partial_{x}\Phi(x+z)-\tfrac{1}{2}\partial_{x}^{2}\Phi(x)z)+\partial_{x}\Phi(x+z)\right).

Now observe that

|ζmznq~(x,ξ;z,ζ)|\displaystyle|{\partial_{\zeta}^{m}\partial_{z}^{n}\widetilde{q}(x,\xi;z,\zeta)}| (5.9)
m,n,AΦ,n,Aq,m+nλqmmax{λμλq,μ}n\displaystyle\lesssim_{m,n,A_{\Phi,n},A_{q,m+n}}\lambda_{q}^{-m}\max\left\{\frac{\lambda\mu}{\lambda_{q}},\mu\right\}^{n}
×g(x)M(σ(ζ+ξ+xΦ(x)xΦ(x+z)12x2Φ(x)z)+xΦ(x+z)).\displaystyle\phantom{\lesssim_{m,n}}\times g(x)M(\sigma(\zeta+\xi+\partial_{x}\Phi(x)-\partial_{x}\Phi(x+z)-\tfrac{1}{2}\partial_{x}^{2}\Phi(x)z)+\partial_{x}\Phi(x+z)).

We shall view (5.8) as an oscillatory integral with the simple phase ζz\zeta z. Accordingly, we introduce the dyadic decomposition

QZ,H(x,ξ)=χH(ζ)χZ(z)q~(x,ξ;z,ζ)eir(3)[Φ](x+z,x)z33!eiζzdζdz,Q_{Z,H}(x,\xi)=\iint\chi_{H}(\zeta)\chi_{Z}(z)\widetilde{q}(x,\xi;z,\zeta)e^{ir^{(3)}[\Phi](x+z,x)\frac{z^{3}}{3!}}e^{-i\zeta z}\,\mathrm{d}\zeta\mathrm{d}z,

as well as Q<Z,H(x,ξ)Q_{<Z,H}(x,\xi), QZ,<H(x,ξ)Q_{Z,<H}(x,\xi) and Q<Z,<H(x,ξ)Q_{<Z,<H}(x,\xi), which are similarly defined. We also introduce the shorthands

MZ,H=z(12Z,2Z)ζ(12H,2H)M(σ(ζ+ξ+xΦ(x)xΦ(x+z)12x2Φ(x)z)+xΦ(x+z))dζdz,M_{Z,H}=\int_{z\in(\frac{1}{2}Z,2Z)}\int_{\zeta\in(\frac{1}{2}H,2H)}M(\sigma(\zeta+\xi+\partial_{x}\Phi(x)-\partial_{x}\Phi(x+z)-\tfrac{1}{2}\partial_{x}^{2}\Phi(x)z)+\partial_{x}\Phi(x+z))\,\mathrm{d}\zeta\mathrm{d}z,

as well as M<H,ZM_{<H,Z}, MH,<ZM_{H,<Z} and M<H,<ZM_{<H,<Z}, which are similarly defined.

The core localized integral bounds are as follows: For dyadic numbers Zλq1Z\gtrsim\lambda_{q}^{-1} and HλqH\gtrsim\lambda_{q},

|QZ,H(x,ξ)|\displaystyle|{Q_{Z,H}(x,\xi)}| m,n,AΦ,n+2,Aq,m+ng(x)MZ,H(Z1λq1)m\displaystyle\lesssim_{m,n,A_{\Phi,n+2},A_{q,m+n}}g(x)M_{Z,H}\left(Z^{-1}\lambda_{q}^{-1}\right)^{m} (5.10)
×(H1max{Z1,λμλq,μ,λμ2Z2+λμ3Z3})n,\displaystyle\phantom{\lesssim_{m,n,A_{\Phi,n+2},A_{q,m+n}}}\times\left(H^{-1}\max\left\{Z^{-1},\frac{\lambda\mu}{\lambda_{q}},\mu,\lambda\mu^{2}Z^{2}+\lambda\mu^{3}Z^{3}\right\}\right)^{n},
|Q<λq1,H(x,ξ)|\displaystyle|{Q_{<\lambda_{q}^{-1},H}(x,\xi)}| n,AΦ,n+2,Aq,ng(x)M<λq1,H(H1max{λq,λμλq,μ,λλqμ2λq+λλqμ3λq2})n,\displaystyle\lesssim_{n,A_{\Phi,n+2},A_{q,n}}g(x)M_{<\lambda_{q}^{-1},H}\left(H^{-1}\max\left\{\lambda_{q},\frac{\lambda\mu}{\lambda_{q}},\mu,\frac{\lambda}{\lambda_{q}}\frac{\mu^{2}}{\lambda_{q}}+\frac{\lambda}{\lambda_{q}}\frac{\mu^{3}}{\lambda_{q}^{2}}\right\}\right)^{n}, (5.11)
|QZ,<λq(x,ξ)|\displaystyle|{Q_{Z,<\lambda_{q}}(x,\xi)}| m,Aq,mg(x)MZ,<λq(Z1λq1)m,\displaystyle\lesssim_{m,A_{q,m}}g(x)M_{Z,<\lambda_{q}}\left(Z^{-1}\lambda_{q}^{-1}\right)^{m}, (5.12)
|Q<λq1,<λq(x,ξ)|\displaystyle|{Q_{<\lambda_{q}^{-1},<\lambda_{q}}(x,\xi)}| g(x)M<λq1,<λq.\displaystyle\lesssim g(x)M_{<\lambda_{q}^{-1},<\lambda_{q}}. (5.13)

Indeed, (5.13) is trivial by the definition of M<λq1,<λqM_{<\lambda_{q}^{-1},<\lambda_{q}} and the fact that the volume of the support of χ<λq(ζ)χ<λq1(z)\chi_{<\lambda_{q}}(\zeta)\chi_{<\lambda_{q}^{-1}}(z) is O(1)O(1). To prove (5.12), we simply use eiζz=iz1ζeiζze^{-i\zeta z}=iz^{-1}\partial_{\zeta}e^{-i\zeta z} and integration by parts and estimate

|QZ,<λq|\displaystyle|{Q_{Z,<\lambda_{q}}}| =|χ<λq(ζ)χZ(z)q~(x,ξ;z,ζ)eir(3)[Φ](x+z,x)z33!znζneiζzdζdz|\displaystyle=\left|{\iint\chi_{<\lambda_{q}}(\zeta)\chi_{Z}(z)\widetilde{q}(x,\xi;z,\zeta)e^{ir^{(3)}[\Phi](x+z,x)\frac{z^{3}}{3!}}z^{-n}\partial_{\zeta}^{n}e^{-i\zeta z}\,\mathrm{d}\zeta\mathrm{d}z}\right|
=|ζm(χ<λq(ζ)q~(x,ξ;z,ζ))χZ(z)eir(3)[Φ](x+z,x)z33!zmeiζzdζdz|\displaystyle=\left|{\iint\partial_{\zeta}^{m}\left(\chi_{<\lambda_{q}}(\zeta)\widetilde{q}(x,\xi;z,\zeta)\right)\chi_{Z}(z)e^{ir^{(3)}[\Phi](x+z,x)\frac{z^{3}}{3!}}z^{-m}e^{-i\zeta z}\,\mathrm{d}\zeta\mathrm{d}z}\right|
m,Aq,mg(x)MZ,<λqZmλqm,\displaystyle\lesssim_{m,A_{q,m}}g(x)M_{Z,<\lambda_{q}}\,Z^{-m}\lambda_{q}^{-m},

where ZλqZ\lambda_{q} is the volume of the support of χ<λq(ζ)χZ(z)\chi_{<\lambda_{q}}(\zeta)\chi_{Z}(z). Next, to prove (5.11), we use eiζz=iζ1zeiζze^{-i\zeta z}=i\zeta^{-1}\partial_{z}e^{-i\zeta z} and integration by parts to bound

|Q<λq1,H|\displaystyle\left|{Q_{<\lambda_{q}^{-1},H}}\right| =|χH(ζ)zn(χ<λq1(z)q~(x,ξ;z,ζ)eir(3)[Φ](x+z,x)z33!)ζneiζzdζdz|.\displaystyle=\left|{\iint\chi_{H}(\zeta)\partial_{z}^{n}\left(\chi_{<\lambda_{q}^{-1}}(z)\widetilde{q}(x,\xi;z,\zeta)e^{ir^{(3)}[\Phi](x+z,x)\frac{z^{3}}{3!}}\right)\zeta^{-n}e^{-i\zeta z}\,\mathrm{d}\zeta\mathrm{d}z}\right|.

Then the desired estimate (5.11) follows, via the chain rule and the Leibniz rule, from (5.9) and

|zn(r(3)[Φ](x+z,x))|AΦ,n+2λμn+2.\left|{\partial_{z}^{n^{\prime}}\left(r^{(3)}[\Phi](x+z,x)\right)}\right|\lesssim_{A_{\Phi,n^{\prime}+2}}\lambda\mu^{n^{\prime}+2}. (5.14)

Finally, to prove (5.10), we use both eiζz=iz1ζeiζze^{-i\zeta z}=iz^{-1}\partial_{\zeta}e^{-i\zeta z} and eiζz=iζ1zeiζze^{-i\zeta z}=i\zeta^{-1}\partial_{z}e^{-i\zeta z} and integration by parts to estimate

|QZ,H|\displaystyle\left|{Q_{Z,H}}\right| =|ζm(χH(ζ)q~(x,ξ;z,ζ))χZ(z)eir(3)[Φ](x+z,x)z33!zmeiζzdζdz|\displaystyle=\left|{\iint\partial_{\zeta}^{m}\left(\chi_{H}(\zeta)\widetilde{q}(x,\xi;z,\zeta)\right)\chi_{Z}(z)e^{ir^{(3)}[\Phi](x+z,x)\frac{z^{3}}{3!}}z^{-m}e^{-i\zeta z}\,\mathrm{d}\zeta\mathrm{d}z}\right|
=|zn(ζm(χH(ζ)q~(x,ξ;z,ζ))χZ(z)zmeir(3)[Φ](x+z,x)z33!)ζneiζzdζdz|.\displaystyle=\left|{\iint\partial_{z}^{n}\left(\partial_{\zeta}^{m}\left(\chi_{H}(\zeta)\widetilde{q}(x,\xi;z,\zeta)\right)\chi_{Z}(z)z^{-m}e^{ir^{(3)}[\Phi](x+z,x)\frac{z^{3}}{3!}}\right)\zeta^{-n}e^{-i\zeta z}\,\mathrm{d}\zeta\mathrm{d}z}\right|.

From the last line, the desired estimate (5.10) follows, via the chain rule and the Leibniz rule, from (5.9) and (5.14).

We are now ready to prove (5.6). In what follows, we omit the dependence of implicit constants on AΦ,10C0A_{\Phi,10C_{0}} and Aq,10C0A_{q,10C_{0}}. Consider (5.10) with m=C0+2m=C_{0}+2 and n=3m+C0+2n=3m+C_{0}+2, and sum over Zλq1Z\gtrsim\lambda_{q}^{-1}, HλqH\gtrsim\lambda_{q}. Then

Zλq1,Hλq|QZ,H(x,ξ)|\displaystyle\sum_{Z\gtrsim\lambda_{q}^{-1},\,H\gtrsim\lambda_{q}}|{Q_{Z,H}(x,\xi)}|
g(x)Zλq1,HλqMZ,HHmmax{Z1,λμλq,μ,λμ2Z2+λμ3Z3}mZnλqn\displaystyle\lesssim g(x)\sum_{Z\gtrsim\lambda_{q}^{-1},\,H\gtrsim\lambda_{q}}M_{Z,H}\cdot H^{-m}\max\left\{Z^{-1},\frac{\lambda\mu}{\lambda_{q}},\mu,\lambda\mu^{2}Z^{2}+\lambda\mu^{3}Z^{3}\right\}^{m}Z^{-n}\lambda_{q}^{-n}
g(x)Hλq,(Zλq1MZ,H(Zλq)C0)Hmmax{λq,λμλq,μ,λμ2λq2+λμ3λq3}m\displaystyle\lesssim g(x)\sum_{H\gtrsim\lambda_{q},}\left(\sum_{Z\gtrsim\lambda_{q}^{-1}}M_{Z,H}(Z\lambda_{q})^{-C_{0}}\right)\,H^{-m}\max\left\{\lambda_{q},\frac{\lambda\mu}{\lambda_{q}},\mu,\lambda\mu^{2}\lambda_{q}^{-2}+\lambda\mu^{3}\lambda_{q}^{-3}\right\}^{m}
g(x)(Zλq1,HλqMZ,H(Zλq)C0(Hλq1)C0)max{1,μλq1,λλq(μλq1+μ3λq3)}C0+2\displaystyle\lesssim g(x)\left(\sum_{Z\gtrsim\lambda_{q}^{-1},\,H\gtrsim\lambda_{q}}M_{Z,H}(Z\lambda_{q})^{-C_{0}}(H\lambda_{q}^{-1})^{-C_{0}}\right)\max\left\{1,\mu\lambda_{q}^{-1},\frac{\lambda}{\lambda_{q}}\left(\mu\lambda_{q}^{-1}+\mu^{3}\lambda_{q}^{-3}\right)\right\}^{C_{0}+2}
g(x)<λq1,<λqmax{1,μλq1,λλq(μλq1+μ3λq3)}C0+2.\displaystyle\lesssim g(x)\mathcal{M}_{<\lambda_{q}^{-1},<\lambda_{q}}\max\left\{1,\mu\lambda_{q}^{-1},\frac{\lambda}{\lambda_{q}}\left(\mu\lambda_{q}^{-1}+\mu^{3}\lambda_{q}^{-3}\right)\right\}^{C_{0}+2}.

Next, invoking (5.11) with m=C0+2m=C_{0}+2 and summing over HλqH\gtrsim\lambda_{q}, we obtain

Hλq|Q<λq1,H(x,ξ)|\displaystyle\sum_{H\gtrsim\lambda_{q}}|{Q_{<\lambda_{q}^{-1},H}(x,\xi)}|
g(x)HλqM<λq1,HHmmax{λq,λμλq,μ,λλqμ2λq+λλqμ3λq2}m\displaystyle\lesssim g(x)\sum_{H\gtrsim\lambda_{q}}M_{<\lambda_{q}^{-1},H}\cdot H^{-m}\max\left\{\lambda_{q},\frac{\lambda\mu}{\lambda_{q}},\mu,\frac{\lambda}{\lambda_{q}}\frac{\mu^{2}}{\lambda_{q}}+\frac{\lambda}{\lambda_{q}}\frac{\mu^{3}}{\lambda_{q}^{2}}\right\}^{m}
g(x)(HλqM<λq1,H(Hλq1)C0)max{1,μλq1,λλq(μλq1+μ3λq3)}C0+2\displaystyle\lesssim g(x)\left(\sum_{H\gtrsim\lambda_{q}}M_{<\lambda_{q}^{-1},H}(H\lambda_{q}^{-1})^{-C_{0}}\right)\max\left\{1,\mu\lambda_{q}^{-1},\frac{\lambda}{\lambda_{q}}\left(\mu\lambda_{q}^{-1}+\mu^{3}\lambda_{q}^{-3}\right)\right\}^{C_{0}+2}
g(x)<λq1,<λqmax{1,μλq1,λλq(μλq1+μ3λq3)}C0+2.\displaystyle\lesssim g(x)\mathcal{M}_{<\lambda_{q}^{-1},<\lambda_{q}}\max\left\{1,\mu\lambda_{q}^{-1},\frac{\lambda}{\lambda_{q}}\left(\mu\lambda_{q}^{-1}+\mu^{3}\lambda_{q}^{-3}\right)\right\}^{C_{0}+2}.

From (5.12) with n=C0+2n=C_{0}+2 and summing over Zλq1Z\gtrsim\lambda_{q}^{-1}, we obtain

Zλq1|QZ,<λq(x,ξ)|\displaystyle\sum_{Z\gtrsim\lambda_{q}^{-1}}|{Q_{Z,<\lambda_{q}}(x,\xi)}| g(x)Zλq1MZ,<λq(Z1λq1)n\displaystyle\lesssim g(x)\sum_{Z\gtrsim\lambda_{q}^{-1}}M_{Z,<\lambda_{q}}\left(Z^{-1}\lambda_{q}^{-1}\right)^{n}
g(x)(Zλq1MZ,<λq(Zλq)C0)\displaystyle\lesssim g(x)\left(\sum_{Z\gtrsim\lambda_{q}^{-1}}M_{Z,<\lambda_{q}}(Z\lambda_{q})^{-C_{0}}\right)
g(x)<λq1,<λq.\displaystyle\lesssim g(x)\mathcal{M}_{<\lambda_{q}^{-1},<\lambda_{q}}.

Combining the preceding three bounds with (5.13), we obtain (5.6) with =0\ell=0.

To handle the case >0\ell>0, note in (5.8) that the ξ\xi-dependence of Q(x,ξ)Q(x,\xi) arises entirely from that of q~(x,ξ;z,ζ)\widetilde{q}(x,\xi;z,\zeta) in the integrand. By (5.7) and the definition of q~(x,ξ;z,ζ)\widetilde{q}(x,\xi;z,\zeta), observe that

|ξζmznq~(x,ξ;z,ζ)|,m,n,AΦ,n,Aq,+m+nλq×(RHS of (5.9)).|{\partial_{\xi}^{\ell}\partial_{\zeta}^{m}\partial_{z}^{n}\widetilde{q}(x,\xi;z,\zeta)}|\lesssim_{\ell,m,n,A_{\Phi,n},A_{q,\ell+m+n}}\lambda_{q}^{-\ell}\times(\hbox{RHS of \eqref{eq:Q-ampl-reg}}).

Using the preceding bound in place of (5.9) and repeating the remainder of the =0\ell=0 proof, we obtain (5.6) with higher \ell’s. ∎

Remark 5.4.

Under additional assumptions on xx-derivatives of q(x1,x,λ0,ξ)q(x_{1},x,\lambda_{0},\xi), the computation as above also yields bounds on xkQ\partial_{x}^{k}Q. For instance, if the symbol qq is independent of xx, then |xkξQ|(λλqμmax{1,μ2λq2})k×(RHS of (5.6) with g1)|{\partial_{x}^{k}\partial_{\xi}^{\ell}Q}|\lesssim\left(\frac{\lambda}{\lambda_{q}}\mu\max\{1,\mu^{2}\lambda_{q}^{-2}\}\right)^{k}\times(\hbox{RHS of \eqref{eq:osc-Q-key} with $g\equiv 1$}). In this case, by the Calderón–Vaillancourt theorem, Q(x,D)Q(x,D) is L2L^{2}-bounded with norm O(<λq,<λq1)O(\mathcal{M}_{<\lambda_{q},<\lambda_{q}^{-1}}) as long as λμλq2=O(1)\frac{\lambda\mu}{\lambda_{q}^{2}}=O(1) and λq1μ=O(1)\lambda_{q}^{-1}\mu=O(1). However, since we are able to exploit higher derivatives bounds for aa in (5.5) (which originate from Section 4), such spatial derivative bounds for QQ will not be necessary; see Section 5.2 below.

5.2 Kernel bounds and proof of Proposition 5.1

To complete the proof of Proposition 5.1, it remains to translate the symbol bound (5.6) to an operator bound for (Φ)q(x2,D2;σ){}^{(\Phi)}q(x_{2},D_{2};\sigma). As we are allowed to use high derivatives of aa on the RHS of (5.5), it suffices to utilize the following simple kernel bound:

Lemma 5.5.

Let n0n\in\mathbb{N}_{0} and p(x,ξ)C(×)p(x,\xi)\in C^{\infty}(\mathbb{R}\times\mathbb{R}) satisfy, for any 0n0\leq\ell\leq n,

|ξp(x,ξ)|Ap,λpG(x,ξ),|{\partial_{\xi}^{\ell}p(x,\xi)}|\leq A_{p,\ell}\lambda_{p}^{-\ell}G(x,\xi), (5.15)

where Ap,A_{p,\ell} are increasing. Let ν2\nu\in 2^{\mathbb{Z}} and k0k\in\mathbb{N}_{0}. Then the kernel Kp(Δ)kPν(x,y)K_{p(-\Delta)^{-k}P_{\nu}}(x,y) of the operator p(x,D)(Δ)kPν(D)up(x,D)(-\Delta)^{-k}P_{\nu}(D)u obeys, for any xyx\neq y, the pointwise bound

|Kp(Δ)kPν(x,y)|n,Ap,n,kν2k14ν<|ξ|<4νG(x,ξ)dξmin{ν,λp}(xy)n.|{K_{p(-\Delta)^{-k}P_{\nu}}(x,y)}|\lesssim_{n,A_{p,n},k}\frac{\nu^{-2k}\int_{\frac{1}{4}\nu<|{\xi}|<4\nu}G(x,\xi)\,\mathrm{d}\xi}{\langle{\min\{\nu,\lambda_{p}\}(x-y)}\rangle^{n}}. (5.16)

Moreover, the kernel KpP<ν(x,y)K_{pP_{<\nu}}(x,y) of the operator p(x,D)P<ν(D)up(x,D)P_{<\nu}(D)u obeys the same bound as KpPνK_{pP_{\nu}}.

Proof.

By definition,

p(x,D)Pν(D)u=p(x,ξ)Pν(ξ)u^(ξ)eixξdξ2π=KpPν(x,y)u(y)dy,p(x,D)P_{\nu}(D)u=\int p(x,\xi)P_{\nu}(\xi)\hat{u}(\xi)e^{ix\cdot\xi}\,\frac{\mathrm{d}\xi}{2\pi}=\int K_{pP_{\nu}}(x,y)u(y)\,\mathrm{d}y,

so formally,

Kp(Δ)2kPν(x,y)=p(x,ξ)|ξ|2kPν(ξ)ei(xy)ξdξ2π.K_{p(-\Delta)^{-2k}P_{\nu}}(x,y)=\int p(x,\xi)|{\xi}|^{-2k}P_{\nu}(\xi)e^{i(x-y)\cdot\xi}\,\frac{\mathrm{d}\xi}{2\pi}.

Thanks to the compact support property of Pν(ξ)P_{\nu}(\xi), this formal computation is readily justified. Moreover, by the identity ei(xy)ξ=(i(xjyj))1ξjei(xy)ξe^{i(x-y)\cdot\xi}=(i(x_{j}-y_{j}))^{-1}\partial_{\xi_{j}}e^{i(x-y)\cdot\xi}, integration by parts and using the bounds (5.15) and |ξPν(ξ)|ν|{\partial_{\xi}^{\ell}P_{\nu}(\xi)}|\lesssim_{\ell}\nu^{-\ell}, (5.16) follows. The case of KpP<νK_{pP_{<\nu}} is entirely analogous. ∎

We are now ready to complete the proof of Proposition 5.1.

Proof.

We introduce the shorthand Q(x2,ξ2)=(Φ)q(x2,D2;σ)Q(x_{2},\xi_{2})={}^{(\Phi)}q(x_{2},D_{2};\sigma). We begin by splitting

Q(x2,D2)a=Q(x2,D2)P<μ(D2)a+ν2,ν>μ(Φ)q(x2,D2;σ)Pν(D2)a.\displaystyle Q(x_{2},D_{2})a=Q(x_{2},D_{2})P_{<\mu}(D_{2})a+\sum_{\nu\in 2^{\mathbb{Z}},\nu>\mu}{}^{(\Phi)}q(x_{2},D_{2};\sigma)P_{\nu}(D_{2})a.

Consider the summand Q(x2,D2)Pν(D2)aQ(x_{2},D_{2})P_{\nu}(D_{2})a with ν>μ\nu>\mu. By Lemma 5.3 with C0=N100C_{0}=\frac{N}{100} and (5.4), Q(x2,ξ2)Q(x_{2},\xi_{2}) obeys (5.15) with G(x,ξ)=g(x2)<λq1,<λq(x2,ξ2;σ)G(x,\xi)=g(x_{2})\mathcal{M}_{<\lambda_{q}^{-1},<\lambda_{q}}(x_{2},\xi_{2};\sigma). By Lemma 5.5 with n=Nn=N, the kernel of Q(x2,D2)Pν(D2)Q(x_{2},D_{2})P_{\nu}(D_{2}) obeys the bound

|KQ(x22)kPν(x2,y2)|\displaystyle|{K_{Q(-\partial_{x_{2}}^{2})^{-k}P_{\nu}}(x_{2},y_{2})}| ν2kg(x2)|ξ2|ν<λq1,<λq(x2,ξ2;σ)dξ2min{ν,λq}(x2y2)N.\displaystyle\lesssim\frac{\nu^{-2k}g(x_{2})\int_{|{\xi_{2}}|\simeq\nu}\mathcal{M}_{<\lambda_{q}^{-1},<\lambda_{q}}(x_{2},\xi_{2};\sigma)\mathrm{d}\xi_{2}}{\langle{\min\{\nu,\lambda_{q}\}(x_{2}-y_{2})}\rangle^{N}}.

Note that, for νμ\nu\geq\mu,

g(x2)ν(y2x2)N100\displaystyle\frac{g(x_{2})}{\langle{\nu(y_{2}-x_{2})}\rangle^{\frac{N}{100}}} g(x2)μ(y2x2)N100supz2g(z2)μ(y2z2)C1=g¯<μ1(y2),\displaystyle\leq\frac{g(x_{2})}{\langle{\mu(y_{2}-x_{2})}\rangle^{\frac{N}{100}}}\lesssim\sup_{z_{2}}\frac{g(z_{2})}{\langle{\mu(y_{2}-z_{2})}\rangle^{C_{1}}}=\overline{g}_{<\mu^{-1}}(y_{2}),
|ξ2|ν<λq1,<λq(x2,ξ2;σ)dξ2ν(y2x2)N100\displaystyle\frac{\int_{|{\xi_{2}}|\simeq\nu}\mathcal{M}_{<\lambda_{q}^{-1},<\lambda_{q}}(x_{2},\xi_{2};\sigma)\mathrm{d}\xi_{2}}{\langle{\nu(y_{2}-x_{2})}\rangle^{\frac{N}{100}}} |ξ2|ν<λq1,<λq(x2,ξ2;σ)μ1μ1ξ2N100dξ2μ(y2x2)N100μ(νμ)N100\displaystyle\lesssim\frac{\int_{|{\xi_{2}}|\simeq\nu}\mathcal{M}_{<\lambda_{q}^{-1},<\lambda_{q}}(x_{2},\xi_{2};\sigma)\frac{\mu^{-1}}{\langle{\mu^{-1}\xi_{2}}\rangle^{\frac{N}{100}}}\mathrm{d}\xi_{2}}{\langle{\mu(y_{2}-x_{2})}\rangle^{\frac{N}{100}}}\mu\left(\frac{\nu}{\mu}\right)^{\frac{N}{100}}
¯<μ1(x2;σ)μ(νμ)N100,\displaystyle\leq\overline{\mathcal{M}}_{<\mu^{-1}}(x_{2};\sigma)\mu\left(\frac{\nu}{\mu}\right)^{\frac{N}{100}},
μ(x2X0)sν(y2x2)|s|\displaystyle{\frac{\langle{\mu(x_{2}-X_{0})}\rangle^{s}}{\langle{\nu(y_{2}-x_{2})}\rangle^{|{s}|}}} μ(x2X0)sμ(y2x2)|s|μ(y2X0)s.\displaystyle{\leq\frac{\langle{\mu(x_{2}-X_{0})}\rangle^{s}}{\langle{\mu(y_{2}-x_{2})}\rangle^{|{s}|}}\lesssim\langle{\mu(y_{2}-X_{0})}\rangle^{s}.}

Therefore,

|KQ(x22)kPν(x2,y2)|μ(x2X0)s\displaystyle|{K_{Q(-\partial_{x_{2}}^{2})^{-k}P_{\nu}}(x_{2},y_{2})}|{\langle{\mu(x_{2}-X_{0})}\rangle^{s}}
ν2k(μmin{ν,λq})(νμ)N100g¯<μ1(y2)¯<μ1(y2;σ)μ(y2X0)smin{ν,λq}min{ν,λq}(x2y2)N2,\displaystyle\lesssim\nu^{-2k}\left(\frac{\mu}{\min\{\nu,\lambda_{q}\}}\right)\left(\frac{\nu}{\mu}\right)^{\frac{N}{100}}\overline{g}_{<\mu^{-1}}(y_{2})\overline{\mathcal{M}}_{<\mu^{-1}}(y_{2};\sigma){\langle{\mu(y_{2}-X_{0})}\rangle^{s}}\frac{\min\{\nu,\lambda_{q}\}}{\langle{\min\{\nu,\lambda_{q}\}(x_{2}-y_{2})}\rangle^{\frac{N}{2}}},

Since N2>1\frac{N}{2}>1, the last factor defines a kernel that is L2L^{2}-bounded. Hence,

Q(x2,D2)Pν(D2)aμ(x2X0)sL2\displaystyle\|{Q(x_{2},D_{2})P_{\nu}(D_{2})a{\langle{\mu(x_{2}-X_{0})}\rangle^{s}}}\|_{L^{2}}
ν2k(μmin{ν,λq})(νμ)N100g¯<μ1(y2)¯<μ1(y2;σ)x22ka(y2)μ(y2X0)sL2y2\displaystyle\lesssim\nu^{-2k}\left(\frac{\mu}{\min\{\nu,\lambda_{q}\}}\right)\left(\frac{\nu}{\mu}\right)^{\frac{N}{100}}\|{\overline{g}_{<\mu^{-1}}(y_{2})\overline{\mathcal{M}}_{<\mu^{-1}}(y_{2};\sigma)\partial_{x_{2}}^{2k}a(y_{2}){\langle{\mu(y_{2}-X_{0})}\rangle^{s}}}\|_{L^{2}_{y_{2}}}

Choosing k=N2k=\lfloor\frac{N}{2}\rfloor and summing over all dyadic numbers ν>μ\nu>\mu (here, we use N100N\gg 100), we obtain

ν2,ν>μQ(x2,D2)Pν(D2)aμ(x2X0)sL2g¯<μ1(y2)¯<μ1(y2;σ)(μ1x2)2N2a(y2)μ(y2X0)sL2y2,\displaystyle\sum_{\nu\in 2^{\mathbb{Z}},\,\nu>\mu}\|{Q(x_{2},D_{2})P_{\nu}(D_{2})a{\langle{\mu(x_{2}-X_{0})}\rangle^{s}}}\|_{L^{2}}\lesssim\|{\overline{g}_{<\mu^{-1}}(y_{2})\overline{\mathcal{M}}_{<\mu^{-1}}(y_{2};\sigma)(\mu^{-1}\partial_{x_{2}})^{2\lfloor\frac{N}{2}\rfloor}a(y_{2})\langle{\mu(y_{2}-X_{0})}\rangle^{s}}\|_{L^{2}_{y_{2}}},

which is acceptable. Moreover, the following can be proved in an entirely analogous manner:

Q(x2,D2)P<μ(D2)aμ(x2X0)sL2g¯<μ1(y2)¯<μ1(y2;σ)a(y2)μ(y2X0)sL2y2.\displaystyle\|{Q(x_{2},D_{2})P_{<\mu}(D_{2})a{\langle{\mu(x_{2}-X_{0})}\rangle^{s}}}\|_{L^{2}}\lesssim\|{\overline{g}_{<\mu^{-1}}(y_{2})\overline{\mathcal{M}}_{<\mu^{-1}}(y_{2};\sigma)a(y_{2}){\langle{\mu(y_{2}-X_{0})}\rangle^{s}}}\|_{L^{2}_{y_{2}}}.

Combining the last two bounds, we obtain (5.5). ∎

6 Construction of degenerating wave packets

In this section, we finally put together all tools developed so far to construct degenerating wave packets for θ̊\mathcal{L}_{\mathring{\theta}}.

6.1 Specification of the construction

As in Section 4, we are given: a symbol γ\gamma and a function ff that satisfy the assumptions in Section 1.5; λ0\lambda_{0}\in\mathbb{N}, M>1M>1 and δ0>0\delta_{0}>0 that satisfy (1.18)–(1.20). When ff is time-dependent, we are given δ1>0\delta_{1}>0 such that (1.25) hold, as well as that ff is even. To apply the results of Section 4, we need to further specify δ2>0\delta_{2}>0 and N0N_{0}; δ2\delta_{2} will be specified below in Proposition 6.3, while N0N_{0} will be chosen in Section 7 (see Proposition 6.3 for the condition that N0N_{0} has to satisfy).

We apply the results of Section 4 with the above parameters, and obtain a global phase function Φ=Φ(t,x2)\Phi=\Phi(t,x_{2}) on [0,11ϵtf(τM)]×[0,\frac{1}{1-\epsilon}t_{f}(\tau_{M})]\times\mathbb{R} as in Section 4.6, which agrees with the solution to the Hamilton–Jacobi equation (4.1) constructed in Sections 4.24.3 in the region X(t,x0)xX(t,x1)X(t,x^{\prime}_{0})\leq x\leq X(t,x^{\prime}_{1}). We inherit the parameters δ3\delta_{3}, δ4\delta_{4}, δ5\delta_{5} and TT fixed in Section 4 (see Section 4.1). The parameters cx0c_{x_{0}} and Λ\Lambda, which were not fixed in Section 4, will be chosen in this section; we shall fix cx0c_{x_{0}} in this subsection, and Λ1\Lambda\geq 1 in Section 6.2.

Given the global phase function Φ\Phi from Section 4, the degenerating wave packet takes the form

φ~(t,x1,x2)=Re(eiλ0x1ψ~), where ψ~=a(t,x2)eiΦ(t,x2),\widetilde{\varphi}(t,x_{1},x_{2})=\operatorname{Re}\left(e^{i\lambda_{0}x_{1}}\widetilde{\psi}\right),\quad\hbox{ where }\widetilde{\psi}=a(t,x_{2})e^{i\Phi(t,x_{2})}, (6.1)

and aa solves (3.16). When Ω=𝕋×\Omega=\mathbb{T}\times\mathbb{R}, we fix a choice of cx0c_{x_{0}} so that it obeys all the requirements in Section 4, then we take as the initial data for aa the following:

a0(x2)=1(x1x0)12χ(x2x0x1x0),a_{0}(x_{2})=\frac{1}{(x^{\prime}_{1}-x^{\prime}_{0})^{\frac{1}{2}}}\chi\left(\frac{x_{2}-x^{\prime}_{0}}{x^{\prime}_{1}-x^{\prime}_{0}}\right), (6.2)

where χ()\chi(\cdot) is a smooth nonnegative function supported in [0,1][0,1], so that suppa0(x0,x1)\operatorname{supp}a_{0}\subseteq(x_{0},x_{1}) and a0L2=1\|{a_{0}}\|_{L^{2}}=1. Then suppa(t,)(X(t;x0),X(t;x1))\operatorname{supp}a(t,\cdot)\subseteq(X(t;x^{\prime}_{0}),X(t;x^{\prime}_{1})), so that Φ(t,x2)\Phi(t,x_{2}) solves (3.15) in the support of φ~\widetilde{\varphi}; hence, the computation in Section 3 is applicable. Moreover, recalling (4.11) or (4.27) and (4.76), as well as the fact that x0x0x^{\prime}_{0}\simeq x_{0} and x1x0Δx0x^{\prime}_{1}-x^{\prime}_{0}\simeq\Delta x_{0}, we have, for any k0k\geq 0,

(μ01x)ka0L2k1,\|{(\mu_{0}^{-1}\partial_{x})^{k}a_{0}}\|_{L^{2}}\lesssim_{k}1,

so that Proposition 4.12 is useful. Moreover, φ~(0,x1,x2)L2x1,x2(Ω)=1\|{\widetilde{\varphi}(0,x_{1},x_{2})}\|_{L^{2}_{x_{1},x_{2}}(\Omega)}=1, and for m0>0m_{0}>0 sufficiently large,

Pm01λ0<<m0λ0(D2)(eiΦ(0,x2)a0)L2x21,\|{P_{m_{0}^{-1}\lambda_{0}<\cdot<m_{0}\lambda_{0}}(D_{2})(e^{i\Phi(0,x_{2})}a_{0})}\|_{L^{2}_{x_{2}}}\gtrsim 1,

where the implicit constant is independent of λ0\lambda_{0}.

When Ω=𝕋2\Omega=\mathbb{T}^{2}, we simply periodize (6.1) and set

φ~(t,x1,x2)=n(𝕋×)φ~(t,x1,x2n).\widetilde{\varphi}(t,x_{1},x_{2})=\sum_{n\in\mathbb{Z}}{}^{(\mathbb{T}\times\mathbb{R})}\widetilde{\varphi}(t,x_{1},x_{2}-n). (6.3)

where (𝕋×)φ~{}^{(\mathbb{T}\times\mathbb{R})}\widetilde{\varphi} is the wave packet in the case Ω=𝕋×\Omega=\mathbb{T}\times\mathbb{R}. Choosing cx0>0c_{x_{0}}>0 to be sufficiently small, we may ensure that each summand (in particular, a0a_{0}) is supported in a fundamental domain.

6.2 Degeneration and initial estimates for the wave packet

The purpose of this subsection is to obtain sharp bounds on the HsH^{s}-norm of φ~\widetilde{\varphi} for a suitable range of ss. In what follows, the following convention is in effect: If Xx2\|{\cdot}\|_{X_{x_{2}}} is a norm for functions of x2x_{2}, then Xx2(Ω)\|{\cdot}\|_{X_{x_{2}}(\Omega)} denotes either Xx2()\|{\cdot}\|_{X_{x_{2}}(\mathbb{R})} or Xx2(𝕋)\|{\cdot}\|_{X_{x_{2}}(\mathbb{T})} depending on whether Ω=𝕋×\Omega=\mathbb{T}\times\mathbb{R} or 𝕋2\mathbb{T}^{2}, respectively. We also introduce, for any k0k\in\mathbb{N}_{0} and L>0L>0,

aH(L)k(Ω)2=k=0k(Lx2)kaL2x2(Ω)2.\|{a}\|_{H_{(L)}^{k}(\Omega)}^{2}=\sum_{k^{\prime}=0}^{k}\|{(L\partial_{x_{2}})^{k^{\prime}}a}\|_{L^{2}_{x_{2}}(\Omega)}^{2}.
Proposition 6.1 (Degeneration and initial estimates).

Let Ω=𝕋×\Omega=\mathbb{T}\times\mathbb{R} or 𝕋2\mathbb{T}^{2}, and let φ~\widetilde{\varphi} be given as in Section 6.1. For λ0Λ\lambda_{0}\geq\Lambda sufficiently large (depending on γ\gamma and the parameters fixed in Section 4), the following statements hold.

  1. 1.

    There exists a decomposition

    φ~=φ~main+φ~small\widetilde{\varphi}=\widetilde{\varphi}^{main}+\widetilde{\varphi}^{small}

    such that for any s,σs,\sigma\in\mathbb{R} and NN\in\mathbb{N} such that |s|+β0|σ|N1|{s}|+\beta_{0}|{\sigma}|\leq N-1 and 0t11ϵtf(τM)0\leq t\leq\frac{1}{1-\epsilon}t_{f}(\tau_{M}),

    Γσφ~main(t,x1,x2)Hsx2(Ω)\displaystyle\|{\Gamma^{\sigma}\widetilde{\varphi}^{main}(t,x_{1},x_{2})}\|_{H^{s}_{x_{2}}(\Omega)} γ,f,Nγλ0(λ)σλsa0H(μ01)N(Ω),\displaystyle\lesssim_{\gamma,f,N}\gamma_{\lambda_{0}}(\lambda)^{\sigma}\lambda^{s}\|{a_{0}}\|_{H_{(\mu_{0}^{-1})}^{N}(\Omega)}, (6.4)
    φ~small(t,x1,x2)L2x2(Ω)\displaystyle\|{\widetilde{\varphi}^{small}(t,x_{1},x_{2})}\|_{L^{2}_{x_{2}}(\Omega)} γ,fλ1μa0H(μ01)1(Ω).\displaystyle\lesssim_{\gamma,f}\lambda^{-1}\mu\|{a_{0}}\|_{H_{(\mu_{0}^{-1})}^{1}(\Omega)}. (6.5)
  2. 2.

    At t=0t=0, we additionally have, for any m0>0m_{0}>0 (and taking Λ\Lambda larger if necessary),

    Γσφ~(0,x1,x2)Hsx2(Ω)\displaystyle\|{\Gamma^{\sigma}\widetilde{\varphi}(0,x_{1},x_{2})}\|_{H^{s}_{x_{2}}(\Omega)} γ,f,Nγλ0(λ0)σλ0sa0H(μ01)N(Ω),\displaystyle\lesssim_{\gamma,f,N}\gamma_{\lambda_{0}}(\lambda_{0})^{\sigma}\lambda_{0}^{s}\|{a_{0}}\|_{H_{(\mu_{0}^{-1})}^{N}(\Omega)}, (6.6)
    Γσφ~(0,x1,x2)Hsx2(Ω)\displaystyle\|{\Gamma^{\sigma}\widetilde{\varphi}(0,x_{1},x_{2})}\|_{H^{s}_{x_{2}}(\Omega)} γ,f,m0γλ0(λ0)σλ0sPm01λ0<<m0λ0(D2)(eiΦ(0,x2)a0(x2))L2(Ω),\displaystyle\gtrsim_{\gamma,f,m_{0}}\gamma_{\lambda_{0}}(\lambda_{0})^{\sigma}\lambda_{0}^{s}\|{P_{m_{0}^{-1}\lambda_{0}<\cdot<m_{0}\lambda_{0}}(D_{2})(e^{i\Phi(0,x_{2})}a_{0}(x_{2}))}\|_{L^{2}(\Omega)}, (6.7)

Moreover, analogous conclusions hold for ψ~\widetilde{\psi}.

The core ingredient of the proof of Proposition 6.1 is the following simple lemma (on \mathbb{R}):

Lemma 6.2.

Assume n0n\in\mathbb{N}_{0}, λ|x2Φ(t,x2)|λ¯\lambda\leq|{\partial_{x_{2}}\Phi(t,x_{2})}|\leq\overline{\lambda} and |x2x2Φ|AΦ,μλ¯|{\partial_{x_{2}}^{\ell}\partial_{x_{2}}\Phi}|\leq A_{\Phi,\ell}\mu^{\ell}\overline{\lambda} on \mathbb{R} for 1n1\leq\ell\leq n, where AΦ,>0A_{\Phi,\ell}>0 is increasing. Then for any ν1\nu\geq 1 and function a=a(x2)a=a(x_{2}),

Pν(D2)(eiΦa)L2()\displaystyle\|{P_{\nu}(D_{2})(e^{i\Phi}a)}\|_{L^{2}(\mathbb{R})} n,AΦ,nmin{(ν+(λ¯/λ)μλ)n,(λ¯+μν)n}=0n(μ1x2)aL2(),\displaystyle\lesssim_{n,A_{\Phi,n}}\min\left\{\left(\frac{\nu+(\overline{\lambda}/\lambda)\mu}{\lambda}\right)^{n},\left(\frac{\overline{\lambda}+\mu}{\nu}\right)^{n}\right\}\sum_{\ell=0}^{n}\|{(\mu^{-1}\partial_{x_{2}})^{\ell}a}\|_{L^{2}(\mathbb{R})}, (6.8)
P<ν(D2)(eiΦa)L2()\displaystyle\|{P_{<\nu}(D_{2})(e^{i\Phi}a)}\|_{L^{2}(\mathbb{R})} n,AΦ,n(ν+(λ¯/λ)μλ)n=0n(μ1x2)aL2().\displaystyle\lesssim_{n,A_{\Phi,n}}\left(\frac{\nu+(\overline{\lambda}/\lambda)\mu}{\lambda}\right)^{n}\sum_{\ell=0}^{n}\|{(\mu^{-1}\partial_{x_{2}})^{\ell}a}\|_{L^{2}(\mathbb{R})}. (6.9)
Proof.

To prove (6.8) for n=1n=1, we write

Pν(D2)(aeiΦ)\displaystyle P_{\nu}(D_{2})\left(ae^{i\Phi}\right) =Pν(D2)(aix2Φx2eiΦ)=Pν(D2)D2(ax2ΦeiΦ)Pν(x2(aix2Φ)eiΦ), or\displaystyle=P_{\nu}(D_{2})\left(\tfrac{a}{i\partial_{x_{2}}\Phi}\partial_{x_{2}}e^{i\Phi}\right)=-P_{\nu}(D_{2})D_{2}\left(\tfrac{a}{\partial_{x_{2}}\Phi}e^{i\Phi}\right)-P_{\nu}\left(\partial_{x_{2}}(\tfrac{a}{i\partial_{x_{2}}\Phi})e^{i\Phi}\right),\hbox{ or }
Pν(D2)(aeiΦ)\displaystyle P_{\nu}(D_{2})\left(ae^{i\Phi}\right) =Pν(D2)D2|D2|2ix2(aeiΦ),\displaystyle=\frac{P_{\nu}(D_{2})D_{2}}{|{D_{2}}|^{2}}i\partial_{x_{2}}\left(ae^{i\Phi}\right),

and use the hypotheses, as well as the bounds Pν(D2)D2L2L2ν\|{P_{\nu}(D_{2})D_{2}}\|_{L^{2}\to L^{2}}\lesssim\nu and Pν(D2)D2|D2|2L2L2ν1\|{\tfrac{P_{\nu}(D_{2})D_{2}}{|{D_{2}}|^{2}}}\|_{L^{2}\to L^{2}}\lesssim\nu^{-1}. Higher nn’s are handled by iteration. Finally, (6.9) is proved using an analogous identity as the first one for P<νP_{<\nu}. ∎

Proof of Proposition 6.1.

To begin with, observe that, thanks to the support conditions for φ~\widetilde{\varphi} and a0a_{0}, we may assume that Ω=𝕋×\Omega=\mathbb{T}\times\mathbb{R} without any loss of generality. Note the identity

DsΓσφ~=DsΓσRe(eiλ0x1ψ~)=12(eiλ0x1D2λ0sγλ0(D2)σψ~+eiλ0x1D2λ0sγλ0(D2)σψ~¯),\langle{D}\rangle^{s}\Gamma^{\sigma}\widetilde{\varphi}=\langle{D}\rangle^{s}\Gamma^{\sigma}\operatorname{Re}(e^{i\lambda_{0}x_{1}}\widetilde{\psi})=\tfrac{1}{2}(e^{i\lambda_{0}x_{1}}\langle{D_{2}}\rangle_{\lambda_{0}}^{s}\gamma_{\lambda_{0}}(D_{2})^{\sigma}\widetilde{\psi}+e^{-i\lambda_{0}x_{1}}\langle{D_{2}}\rangle_{-\lambda_{0}}^{s}\gamma_{-\lambda_{0}}(D_{2})^{\sigma}\overline{\widetilde{\psi}}), (6.10)

where ξ2λ0s=(λ02+ξ22)s2\langle{\xi_{2}}\rangle_{\lambda_{0}}^{s}=(\lambda_{0}^{2}+\xi_{2}^{2})^{\frac{s}{2}}. In view of the symmetries ξ2λ0=ξ2λ0\langle{\xi_{2}}\rangle_{-\lambda_{0}}=\langle{\xi_{2}}\rangle_{\lambda_{0}} and γλ0(ξ2)=γλ0(ξ2)\gamma_{-\lambda_{0}}(\xi_{2})=\gamma_{\lambda_{0}}(-\xi_{2}), the two terms on the RHS are always treated similarly. In what follows, we focus on D2λ0sγλ0(D2)σψ~\langle{D_{2}}\rangle_{\lambda_{0}}^{s}\gamma_{\lambda_{0}}(D_{2})^{\sigma}\widetilde{\psi}, except in the proof of (6.7). To simplify the notation, we shall abbreviate Pν=Pν(D2)P_{\nu}=P_{\nu}(D_{2}), L2=L2x2L^{2}=L^{2}_{x_{2}} and Hs=Hsx2H^{s}=H^{s}_{x_{2}} in this proof. We shall also often suppress the dependence of constants on γ\gamma and ff.

Note that x2Φ\partial_{x_{2}}\Phi obeys the hypothesis of Lemma 6.2 with λ¯λ\overline{\lambda}\lesssim\lambda. Applying Lemma 6.2 for any 0nN0\leq n\leq N, then using Proposition 4.12 to bound aH(μ1)na0H(μ01)n\|{a}\|_{H_{(\mu^{-1})}^{n}}\lesssim\|{a_{0}}\|_{H_{(\mu_{0}^{-1})}^{n}}, we obtain

Pν(D2)ψ~L2\displaystyle\|{P_{\nu}(D_{2})\widetilde{\psi}}\|_{L^{2}} Nmin{(ν+μλ)n,(λ+μν)n}a0Hn(μ01),\displaystyle\lesssim_{N}\min\left\{\left(\frac{\nu+\mu}{\lambda}\right)^{n},\left(\frac{\lambda+\mu}{\nu}\right)^{n}\right\}\|{a_{0}}\|_{H^{n}_{(\mu_{0}^{-1})}}, (6.11)
P<ν(D2)ψ~L2\displaystyle\|{P_{<\nu}(D_{2})\widetilde{\psi}}\|_{L^{2}} N(ν+μλ)na0Hn(μ01).\displaystyle\lesssim_{N}\left(\frac{\nu+\mu}{\lambda}\right)^{n}\|{a_{0}}\|_{H^{n}_{(\mu_{0}^{-1})}}. (6.12)

From (4.48), note moreover that

μλλ01+2δ3N0(logλ0)2ϵ(λ0)1x01=λ01+2δ3N0(logλ0)2ϵ(λ0)2cx01,\frac{\mu}{\lambda}\lesssim\lambda_{0}^{-1+2\delta_{3}N_{0}}(\log\lambda_{0})^{-2}\epsilon(\lambda_{0})^{-1}x_{0}^{-1}=\lambda_{0}^{-1+2\delta_{3}N_{0}}(\log\lambda_{0})^{-2}\epsilon(\lambda_{0})^{-2}c_{x_{0}}^{-1}, (6.13)

where the implicit constant depends only on γ\gamma. Recall (4.3) and that δ3δ2N0δ0N0\delta_{3}\ll\frac{\delta_{2}}{N_{0}}\ll\frac{\delta_{0}}{N_{0}} (see Proposition 4.10). Hence, by taking λ0Λ\lambda_{0}\geq\Lambda sufficiently large, we may ensure that

μλλ013110δ0.\frac{\mu}{\lambda}\leq\lambda_{0}^{-\frac{1}{3}-\frac{1}{10}\delta_{0}}. (6.14)

We also observe that, in view of Assumption 1 for Γ\Gamma, we have, for any bL2()b\in L^{2}(\mathbb{R}) and ν>0\nu>0,

D2λ0γλ0(D2)PνbL2\displaystyle\|{\langle{D_{2}}\rangle_{\lambda_{0}}\gamma_{\lambda_{0}}(D_{2})P_{\nu}b}\|_{L^{2}} γ,s,σ(ν+λ0)sγλ0(ν)σPνbL2,\displaystyle\lesssim_{\gamma,s,\sigma}(\nu+\lambda_{0})^{s}\gamma_{\lambda_{0}}(\nu)^{\sigma}\|{P_{\nu}b}\|_{L^{2}}, (6.15)
D2λ0γλ0(D2)P<λ0bL2\displaystyle\|{\langle{D_{2}}\rangle_{\lambda_{0}}\gamma_{\lambda_{0}}(D_{2})P_{<\lambda_{0}}b}\|_{L^{2}} γ,s,σλ0sγλ0(λ0)σP<λ0bL2.\displaystyle\lesssim_{\gamma,s,\sigma}\lambda_{0}^{s}\gamma_{\lambda_{0}}(\lambda_{0})^{\sigma}\|{P_{<\lambda_{0}}b}\|_{L^{2}}. (6.16)

Now we are ready to begin the proof of the proposition in earnest. To prove the first statement, we begin by defining

ψ~main=νμPνψ~,ψ~small=ψ~ψ~main,\widetilde{\psi}^{main}=\sum_{\nu\geq\mu}P_{\nu}\widetilde{\psi},\qquad\widetilde{\psi}^{small}=\widetilde{\psi}-\widetilde{\psi}^{main},

and, accordingly, φ~main=Re(eiλ0x1ψ~main)\widetilde{\varphi}^{main}=\operatorname{Re}(e^{i\lambda_{0}x_{1}}\widetilde{\psi}^{main}) and φ~small=φ~φ~main\widetilde{\varphi}^{small}=\widetilde{\varphi}-\widetilde{\varphi}^{main}. Applying (6.11) with n=Nn=N, as well as (6.15) and (1.11), we obtain

γλ0(λ)σλsD2λ0γλ0(D2)νμPνψ~L2\displaystyle\gamma_{\lambda_{0}}(\lambda)^{-\sigma}\lambda^{-s}\|{\langle{D_{2}}\rangle_{\lambda_{0}}\gamma_{\lambda_{0}}(D_{2})\sum_{\nu\geq\mu}P_{\nu}\widetilde{\psi}}\|_{L^{2}} Nμν<max{μ,λ0}(λλ0)β0|σ|(λ0λ)s(νλ)Na0HN(μ01)\displaystyle\lesssim_{N}\sum_{\mu\leq\nu<\max\{\mu,\lambda_{0}\}}\left(\frac{\lambda}{\lambda_{0}}\right)^{\beta_{0}|{\sigma}|}\left(\frac{\lambda_{0}}{\lambda}\right)^{s}\left(\frac{\nu}{\lambda}\right)^{N}\|{a_{0}}\|_{H^{N}_{(\mu_{0}^{-1})}}
+max{μ,λ0}ν<λ(λν)β0|σ|(νλ)s(νλ)Na0HN(μ01)\displaystyle\mathrel{\phantom{=}}+\sum_{\max\{\mu,\lambda_{0}\}\leq\nu<\lambda}\left(\frac{\lambda}{\nu}\right)^{\beta_{0}|{\sigma}|}\left(\frac{\nu}{\lambda}\right)^{s}\left(\frac{\nu}{\lambda}\right)^{N}\|{a_{0}}\|_{H^{N}_{(\mu_{0}^{-1})}}
+ν>λ(νλ)β0|σ|(νλ)s(λν)Na0HN(μ01).\displaystyle\mathrel{\phantom{=}}+\sum_{\nu>\lambda}\left(\frac{\nu}{\lambda}\right)^{\beta_{0}|{\sigma}|}\left(\frac{\nu}{\lambda}\right)^{s}\left(\frac{\lambda}{\nu}\right)^{N}\|{a_{0}}\|_{H^{N}_{(\mu_{0}^{-1})}}.

By the assumed lower bound on NN, the RHS may be bounded by a0HN(μ01)\|{a_{0}}\|_{H^{N}_{(\mu_{0}^{-1})}} (up to a constant), which implies (6.4). On the other hand, for ψ~small=P<μψ~\widetilde{\psi}^{small}=P_{<\mu}\widetilde{\psi}, (6.5) follows quickly from (6.12) with n=1n=1 and Proposition 4.12.

Next, we turn to the second statement. Note that (λ,μ,a)=(λ0,μ0,a0)(\lambda,\mu,a)=(\lambda_{0},\mu_{0},a_{0}) at t=0t=0. In order to establish (6.6), in view of (6.4), it suffices to prove

D2λ0Γλ0σ(D2)ψ~small(t=0)L2γλ0(λ0)σλ0sa0H(μ01)1.\|{\langle{D_{2}}\rangle_{\lambda_{0}}\Gamma_{\lambda_{0}}^{\sigma}(D_{2})\widetilde{\psi}^{small}(t=0)}\|_{L^{2}}\lesssim\gamma_{\lambda_{0}}(\lambda_{0})^{\sigma}\lambda_{0}^{s}\|{a_{0}}\|_{H_{(\mu_{0}^{-1})}^{1}}.

Since ψ~small(t=0)=P<μ0ψ~0\widetilde{\psi}^{small}(t=0)=P_{<\mu_{0}}\widetilde{\psi}_{0} and μ0<λ0\mu_{0}<\lambda_{0}, the desired bound follows from (6.12) and (6.16). To prove (6.7), observe that, by (6.10) and orthogonality in the frequency space (first in ξ1\xi_{1} and then in ξ2\xi_{2}), we have

DsΓσφ~0L2x1,x212eiλ0x1D2λ0sγλ0(D2)σψ~0L2x1,x2D2λ0sγλ0(D2)σPm01λ0<<m0λ0ψ~0L2.\displaystyle\|{\langle{D}\rangle^{s}\Gamma^{\sigma}\widetilde{\varphi}_{0}}\|_{L^{2}_{x_{1},x_{2}}}\geq\tfrac{1}{2}\|{e^{i\lambda_{0}x_{1}}\langle{D_{2}}\rangle_{\lambda_{0}}^{s}\gamma_{\lambda_{0}}(D_{2})^{\sigma}\widetilde{\psi}_{0}}\|_{L^{2}_{x_{1},x_{2}}}\gtrsim\|{\langle{D_{2}}\rangle_{\lambda_{0}}^{s}\gamma_{\lambda_{0}}(D_{2})^{\sigma}P_{m_{0}^{-1}\lambda_{0}<\cdot<m_{0}\lambda_{0}}\widetilde{\psi}_{0}}\|_{L^{2}}.

Hence, it suffices to prove

Pm01λ0<<m0λ0ψ~0L2m0λ0sγλ0(λ0)σD2λ0sγλ0(D2)σPm01λ0<<m0λ0ψ~0L2,\displaystyle\|{P_{m_{0}^{-1}\lambda_{0}<\cdot<m_{0}\lambda_{0}}\widetilde{\psi}_{0}}\|_{L^{2}}\lesssim_{m_{0}}\lambda_{0}^{-s}\gamma_{\lambda_{0}}(\lambda_{0})^{-\sigma}\|{\langle{D_{2}}\rangle_{\lambda_{0}}^{s}\gamma_{\lambda_{0}}(D_{2})^{\sigma}P_{m_{0}^{-1}\lambda_{0}<\cdot<m_{0}\lambda_{0}}\widetilde{\psi}_{0}}\|_{L^{2}},

which follows from (6.15). This completes the proof. ∎

We may now fix our choice of Λ\Lambda so that it obeys all the requirements that appeared so far.

6.3 Error estimate

The main result of this subsection is the following:

Proposition 6.3 (Error estimate).

Let φ~\widetilde{\varphi} be constructed as in Section 6.1, with δ2=δ010\delta_{2}=\delta_{0}^{10} and N0104δ2(1+β0)N_{0}\geq\frac{10^{4}}{\delta_{2}}(1+\beta_{0}). Then there exists δ6>0\delta_{6}>0 such that

011ϵtf(τM)θ̊φ~L2x1,x2(Ω)dtλ0δ6a0HN0(μ01)(Ω).\int_{0}^{\frac{1}{1-\epsilon}t_{f}(\tau_{M})}\|{\mathcal{L}_{\mathring{\theta}}\widetilde{\varphi}}\|_{L^{2}_{x_{1},x_{2}}(\Omega)}\,\mathrm{d}t\lesssim\lambda_{0}^{-\delta_{6}}\|{a_{0}}\|_{H^{N_{0}}_{(\mu_{0}^{-1})}(\Omega)}. (6.17)
Proof.

By translation in x2x_{2} and rescaling tt, we assume, without loss of generality, that x̊2=0\mathring{x}_{2}=0 and x22f(0,0)=1\partial_{x_{2}}^{2}f(0,0)=-1. We let all implicit constants in this proof to depend on γ\gamma and ff. Note that

θ̊φ~\displaystyle\mathcal{L}_{\mathring{\theta}}\widetilde{\varphi} =Re(θ̊(eiλ0x1ψ~))\displaystyle=\operatorname{Re}\left(\mathcal{L}_{\mathring{\theta}}(e^{i\lambda_{0}x_{1}}\widetilde{\psi})\right)
=Re(eiλ0x1(ipθ̊,λ0(x2,D2)+sθ̊,λ0(x2,D2)+rθ̊,λ0(x2,D2))ψ~).\displaystyle=\operatorname{Re}\left(e^{i\lambda_{0}x_{1}}\left(ip_{\mathring{\theta},\lambda_{0}}(x_{2},D_{2})+s_{\mathring{\theta},\lambda_{0}}(x_{2},D_{2})+r_{\mathring{\theta},\lambda_{0}}(x_{2},D_{2})\right)\widetilde{\psi}\right).

Hence, in what follows, we focus on estimating the L1tL2x2L^{1}_{t}L^{2}_{x_{2}}-norm of

(ipθ̊,λ0(x2,D2)+sθ̊,λ0(x2,D2)+rθ̊,λ0(x2,D2))ψ~.\left(ip_{\mathring{\theta},\lambda_{0}}(x_{2},D_{2})+s_{\mathring{\theta},\lambda_{0}}(x_{2},D_{2})+r_{\mathring{\theta},\lambda_{0}}(x_{2},D_{2})\right)\widetilde{\psi}.

We first treat the case Ω=𝕋×\Omega=\mathbb{T}\times\mathbb{R}.

Step 1: Contribution of ipθ̊,λ0+sθ̊,λ0ip_{\mathring{\theta},\lambda_{0}}+s_{\mathring{\theta},\lambda_{0}}, low frequency input. By Assumption 1 on γ\gamma, it follows that

|x2kξ2((ipθ̊,λ0(x2,ξ2)+sθ̊,λ0(x2,ξ2))P<λ1δ2(ξ2))|k,ξ2λ0γ(λ0,λ1δ2).\displaystyle\left|{\partial_{x_{2}}^{k}\partial_{\xi_{2}}^{\ell}\left(\left(ip_{\mathring{\theta},\lambda_{0}}(x_{2},\xi_{2})+s_{\mathring{\theta},\lambda_{0}}(x_{2},\xi_{2})\right)P_{<\lambda^{1-\delta_{2}}}(\xi_{2})\right)}\right|\lesssim_{k,\ell}\langle{\xi_{2}}\rangle^{-\ell}\lambda_{0}\gamma(\lambda_{0},\lambda^{1-\delta_{2}}).

By the standard L2L^{2}-boundedness of a pseudo-differential operator with a classical symbol and Lemma 6.2, we have

(ipθ̊,λ0(x2,D2)+sθ̊,λ0(x2,D2))P<λ1δ2(D2)ψ~L2λ0(max{λ0,λ1δ2})β0(μ+λ1δ2λ)NaH(μ1)N.\displaystyle\left\|{\left(ip_{\mathring{\theta},\lambda_{0}}(x_{2},D_{2})+s_{\mathring{\theta},\lambda_{0}}(x_{2},D_{2})\right)P_{<\lambda^{1-\delta_{2}}}(D_{2})\widetilde{\psi}}\right\|_{L^{2}}\lesssim\lambda_{0}(\max\{\lambda_{0},\lambda^{1-\delta_{2}}\})^{\beta_{0}}\left(\frac{\mu+\lambda^{1-\delta_{2}}}{\lambda}\right)^{N}\|{a}\|_{H_{(\mu^{-1})}^{N}}.

By (6.14), we see that μ+λ1δ2λmax{λ013110δ0,λδ2}\frac{\mu+\lambda^{1-\delta_{2}}}{\lambda}\leq\max\{\lambda_{0}^{-\frac{1}{3}-\frac{1}{10}\delta_{0}},\lambda^{-\delta_{2}}\}. Choosing N=δ21(β0+2)N=\delta_{2}^{-1}(\beta_{0}+2), which bounded by N0N_{0}, then applying Proposition 4.12, the RHS is O(λ01a0HN0(μ01))O(\lambda_{0}^{-1}\|{a_{0}}\|_{H^{N_{0}}_{(\mu_{0}^{-1})}}), which is sufficient.

Step 2: Decomposition of ipθ̊,λ0+sθ̊,λ0ip_{\mathring{\theta},\lambda_{0}}+s_{\mathring{\theta},\lambda_{0}} Thanks to our pointwise bound λx2Φ2λ\lambda\leq\partial_{x_{2}}\Phi\leq 2\lambda, observe that

pθ̊,λ0(x2,x2Φ)P>λ1δ2(x2Φ)\displaystyle p_{\mathring{\theta},\lambda_{0}}(x_{2},\partial_{x_{2}}\Phi)P_{>\lambda^{1-\delta_{2}}}(\partial_{x_{2}}\Phi) =pθ̊,λ0(x2,x2Φ),\displaystyle=p_{\mathring{\theta},\lambda_{0}}(x_{2},\partial_{x_{2}}\Phi),
sθ̊,λ0(x2,x2Φ)P>λ1δ2(x2Φ)\displaystyle s_{\mathring{\theta},\lambda_{0}}(x_{2},\partial_{x_{2}}\Phi)P_{>\lambda^{1-\delta_{2}}}(\partial_{x_{2}}\Phi) =sθ̊,λ0(x2,x2Φ),\displaystyle=s_{\mathring{\theta},\lambda_{0}}(x_{2},\partial_{x_{2}}\Phi),

so that by definition,

(Φ)rpθ̊,λ0,2=(Φ)rpθ̊,λ0P>λ1δ2,2,(Φ)rsθ̊,λ0,1=(Φ)rsθ̊,λ0P>λ1δ2,1.{}^{(\Phi)}r_{p_{\mathring{\theta},\lambda_{0}},-2}={}^{(\Phi)}r_{p_{\mathring{\theta},\lambda_{0}}P_{>\lambda^{1-\delta_{2}}},-2},\quad{}^{(\Phi)}r_{s_{\mathring{\theta},\lambda_{0}},-1}={}^{(\Phi)}r_{s_{\mathring{\theta},\lambda_{0}}P_{>\lambda^{1-\delta_{2}}},-1}.

To proceed, we furthermore split pθ̊,λ0p_{\mathring{\theta},\lambda_{0}} into pθ̊,λ0main+pθ̊,λ0remp_{\mathring{\theta},\lambda_{0}}^{main}+p_{\mathring{\theta},\lambda_{0}}^{rem}

pθ̊,λ0main(x2,ξ2)\displaystyle p_{\mathring{\theta},\lambda_{0}}^{main}(x_{2},\xi_{2}) =x2f(t,x2)λ0γλ0(ξ2),\displaystyle=\partial_{x_{2}}f(t,x_{2})\lambda_{0}\gamma_{\lambda_{0}}(\xi_{2}),
pθ̊,λ0rem(x2,ξ2)\displaystyle p_{\mathring{\theta},\lambda_{0}}^{rem}(x_{2},\xi_{2}) =Γx2f(t,x2)λ0.\displaystyle=\Gamma\partial_{x_{2}}f(t,x_{2})\lambda_{0}.

By the linear dependence of (Φ)rp,k{}^{(\Phi)}r_{p,-k} on pp, the remaining task reduces to estimating

(i(Φ)rpθ̊,λ0mainP>λ1δ2,2+i(Φ)rpθ̊,λ0remP>λ1δ2,2+(Φ)rsθ̊,λ0P>λ1δ2,1)(x2,D2)ψ~.\displaystyle\left(i{}^{(\Phi)}r_{p_{\mathring{\theta},\lambda_{0}}^{main}P_{>\lambda^{1-\delta_{2}}},-2}+i{}^{(\Phi)}r_{p_{\mathring{\theta},\lambda_{0}}^{rem}P_{>\lambda^{1-\delta_{2}}},-2}+{}^{(\Phi)}r_{s_{\mathring{\theta},\lambda_{0}}P_{>\lambda^{1-\delta_{2}}},-1}\right)(x_{2},D_{2})\widetilde{\psi}.

Step 3: Contribution of ipθ̊,λ0mainip_{\mathring{\theta},\lambda_{0}}^{main}. In this step, we estimate the contribution of pθ̊,λ0mainP>λ1δ2p_{\mathring{\theta},\lambda_{0}}^{main}P_{>\lambda^{1-\delta_{2}}}. Observe that this symbol obeys

ξ2(pθ̊,λ0main(x2,ξ2)P>λ1δ2(ξ2))λ(1δ2)|x2f(x2)|λ0m(λ0,ξ2).\partial_{\xi_{2}}^{\ell}\left(p_{\mathring{\theta},\lambda_{0}}^{main}(x_{2},\xi_{2})P_{>\lambda^{1-\delta_{2}}}(\xi_{2})\right)\lesssim_{\ell}\lambda^{-\ell(1-\delta_{2})}|{\partial_{x_{2}}f(x_{2})}|\lambda_{0}m(\lambda_{0},\xi_{2}).

As a result, each term in (3.8) for p=pθ̊,λ0main(x2,ξ2)P>λ1δ2(ξ2)p=p_{\mathring{\theta},\lambda_{0}}^{main}(x_{2},\xi_{2})P_{>\lambda^{1-\delta_{2}}}(\xi_{2}) is of the form (after rewriting ξ2=μ(ξ2/μ)\xi_{2}=\mu\cdot(\xi_{2}/\mu))

01(Φ)q(x2,ξ2;σ)dσ(ξ2μ),=0,1,2,\begin{split}\int_{0}^{1}{}^{(\Phi)}q(x_{2},\xi_{2};\sigma)\,\mathrm{d}\sigma\left(\frac{\xi_{2}}{\mu}\right)^{\ell},\qquad\ell=0,1,2,\end{split}

where qq obeys the hypothesis of Proposition 5.1 with

λq=λ1δ2,g(x2)=|f(x2)|,M(ξ2)=λ0γλ0(ξ2)(μ2λ2(1δ2)+μ2λλ3(1δ2)+μ2λ2λ4(1δ2)).\lambda_{q}=\lambda^{1-\delta_{2}},\quad g(x_{2})=|{f^{\prime}(x_{2})}|,\quad M(\xi_{2})=\lambda_{0}\gamma_{\lambda_{0}}(\xi_{2})\left(\frac{\mu^{2}}{{\lambda^{2(1-\delta_{2})}}}+\frac{\mu^{2}\lambda}{{\lambda^{3(1-\delta_{2})}}}+\frac{\mu^{2}\lambda^{2}}{{\lambda^{4(1-\delta_{2})}}}\right).

In order to apply Proposition 5.1, note that, since μλqλ013120δ0\frac{\mu}{\lambda_{q}}\lesssim\lambda_{0}^{-\frac{1}{3}-\frac{1}{20}\delta_{0}} by (6.14), ff^{\prime} is smooth and bounded and γ\gamma obeys (1.11), we have

g¯<μ1x2+μ1,¯<μ1M(λ),\overline{g}_{<\mu^{-1}}\lesssim x_{2}+\mu^{-1},\qquad\overline{\mathcal{M}}_{<\mu^{-1}}\lesssim M(\lambda),

where we choose N=1000(1+β0)N=1000(1+\beta_{0}). We remark that the second inequality may be readily verified using the fact that M(ξ2)M(\xi_{2}) is increasing and slowly varying (i.e., Assumptions 1 and 2 for Γ\Gamma). Putting together the above observations with Propositions 3.3 and 5.1 (with s=0s=0), we arrive at

(Φ)rpθ̊,λ0mainP>λ1δ2,2ψ~L2\displaystyle\|{{}^{(\Phi)}r_{p_{\mathring{\theta},\lambda_{0}}^{main}P_{>\lambda^{1-\delta_{2}}},-2}\widetilde{\psi}}\|_{L^{2}} (X(t;x0)+μ1)λ0γλ0(λ)(μ2λ2(1δ2)+μ2λλ3(1δ2)+μ2λ2λ4(1δ2))aH(μ1)N\displaystyle\lesssim(X(t;x_{0})+\mu^{-1})\lambda_{0}\gamma_{\lambda_{0}}(\lambda)\left(\frac{\mu^{2}}{{\lambda^{2(1-\delta_{2})}}}+\frac{\mu^{2}\lambda}{{\lambda^{3(1-\delta_{2})}}}+\frac{\mu^{2}\lambda^{2}}{{\lambda^{4(1-\delta_{2})}}}\right)\|{a}\|_{H_{(\mu^{-1})}^{N}}
(X(t;x0)+μ1)λ0γλ0(λ)(μλ)2λ010δ2δ0a0H(μ1)N0.\displaystyle\lesssim(X(t;x_{0})+\mu^{-1})\lambda_{0}\gamma_{\lambda_{0}}(\lambda)\left(\frac{\mu}{\lambda}\right)^{2}\lambda_{0}^{\frac{10\delta_{2}}{\delta_{0}}}\|{a_{0}}\|_{H_{(\mu^{-1})}^{N_{0}}}.

We need to estimate the integral of the preceding expression on t[0,11ϵtf(τM)]t\in[0,\frac{1}{1-\epsilon}t_{f}(\tau_{M})]. By (4.67) from the proof of Proposition 4.10, it follows that μ1X(t;x0)\mu^{-1}\lesssim X(t;x_{0}); hence it suffices to estimate the contribution of X(t;x0)X(t;x_{0}). Using (4.3), (4.20) or (4.31) and (6.13), we have

011ϵtf(τM)X(t;x0)λ0γλ0(λ)(μλ)2λ010δ2δ0a0HN0(μ01)dt\displaystyle\int_{0}^{\frac{1}{1-\epsilon}t_{f}(\tau_{M})}X(t;x_{0})\lambda_{0}\gamma_{\lambda_{0}}(\lambda)\left(\frac{\mu}{\lambda}\right)^{2}\lambda_{0}^{\frac{10\delta_{2}}{\delta_{0}}}\|{a_{0}}\|_{H^{N_{0}}_{(\mu_{0}^{-1})}}\,\mathrm{d}t
τMλ0γλ0(λ0)λ02+4δ3N0+10δ2δ0(logλ0)4ϵ(λ0)3cx01a0HN0(μ01)\displaystyle\lesssim\tau_{M}\lambda_{0}\gamma_{\lambda_{0}}(\lambda_{0})\lambda_{0}^{-2+4\delta_{3}N_{0}+\frac{10\delta_{2}}{\delta_{0}}}(\log\lambda_{0})^{-4}\epsilon(\lambda_{0})^{-3}c_{x_{0}}^{-1}\|{a_{0}}\|_{H^{N_{0}}_{(\mu_{0}^{-1})}}
τMγλ0(λ0)λ01δ03σ0λ0δ0+4δ3N0+10δ2δ0+12δ0(logλ0)4cx01a0HN0(μ01)\displaystyle\lesssim\tau_{M}\frac{\gamma_{\lambda_{0}}(\lambda_{0})}{\lambda_{0}^{1-\delta_{0}-3\sigma_{0}}}\lambda_{0}^{-\delta_{0}+4\delta_{3}N_{0}+\frac{10\delta_{2}}{\delta_{0}}+\frac{1}{2}\delta_{0}}(\log\lambda_{0})^{-4}c_{x_{0}}^{-1}\|{a_{0}}\|_{H^{N_{0}}_{(\mu_{0}^{-1})}}

which is tightly acceptable thanks to (1.19).

Step 4: Contribution of ipθ̊,λ0ip_{\mathring{\theta},\lambda_{0}}, remainder. Now, we estimate the contribution of ipθ̊,λ0remP>λ1δ2ip_{\mathring{\theta},\lambda_{0}}^{rem}P_{>\lambda^{1-\delta_{2}}}. Each term in (3.8) for p=pθ̊,λ0rem(x2,ξ2)P>λ1δ2(ξ2)p=p_{\mathring{\theta},\lambda_{0}}^{rem}(x_{2},\xi_{2})P_{>\lambda^{1-\delta_{2}}}(\xi_{2}) is of the form 0σ(Φ)q(x2,ξ2;σ)dσ\int_{0}^{\sigma}{}^{(\Phi)}q(x_{2},\xi_{2};\sigma)\,\mathrm{d}\sigma, where qq obeys the hypothesis of Proposition 5.1 with

λq=λ1δ2,g(x2)=Γx2fL,M(ξ2)=λ0(μ2λ2(1δ2)+μ2λλ3(1δ2)+μ2λ2λ4(1δ2)).\lambda_{q}=\lambda^{1-\delta_{2}},\quad g(x_{2})=\|{\Gamma\partial_{x_{2}}f}\|_{L^{\infty}},\quad M(\xi_{2})=\lambda_{0}\left(\frac{\mu^{2}}{{\lambda^{2(1-\delta_{2})}}}+\frac{\mu^{2}\lambda}{{\lambda^{3(1-\delta_{2})}}}+\frac{\mu^{2}\lambda^{2}}{{\lambda^{4(1-\delta_{2})}}}\right).

Since g(x2)g(x_{2}) and M(ξ2)M(\xi_{2}) are constant, Proposition 5.1 (with s=0s=0) immediately implies

(Φ)rpθ̊,λ0remP>λ1δ2,2ψ~L2\displaystyle\|{{}^{(\Phi)}r_{p_{\mathring{\theta},\lambda_{0}}^{rem}P_{>\lambda^{1-\delta_{2}}},-2}\widetilde{\psi}}\|_{L^{2}} Γx2fLλ0(μ2λ2(1δ2)+μ2λλ3(1δ2)+μ2λ2λ4(1δ2))aHμ110000\displaystyle\lesssim\|{\Gamma\partial_{x_{2}}f}\|_{L^{\infty}}\lambda_{0}\left(\frac{\mu^{2}}{{\lambda^{2(1-\delta_{2})}}}+\frac{\mu^{2}\lambda}{{\lambda^{3(1-\delta_{2})}}}+\frac{\mu^{2}\lambda^{2}}{{\lambda^{4(1-\delta_{2})}}}\right)\|{a}\|_{H_{\mu^{-1}}^{10000}}
Γx2fLλ0λ010δ2+N0δ3λ02a0Hμ1N0,\displaystyle\lesssim\|{\Gamma\partial_{x_{2}}f}\|_{L^{\infty}}\lambda_{0}\frac{\lambda_{0}^{10\delta_{2}+N_{0}\delta_{3}}}{\lambda_{0}^{2}}\|{a_{0}}\|_{H_{\mu^{-1}}^{N_{0}}},

which is acceptable.

Step 5: Contribution of sθ̊,λ0s_{\mathring{\theta},\lambda_{0}}. In this step, we handle the contribution of sθ̊,λ0P>λ1δ2s_{\mathring{\theta},\lambda_{0}}P_{>\lambda^{1-\delta_{2}}}. Each term in (3.7) for p=sθ̊,λ0rem(x2,ξ2)P>λ1δ2(ξ2)p=s_{\mathring{\theta},\lambda_{0}}^{rem}(x_{2},\xi_{2})P_{>\lambda^{1-\delta_{2}}}(\xi_{2}) is of the form 0σ(Φ)q(x2,ξ2;σ)dσ\int_{0}^{\sigma}{}^{(\Phi)}q(x_{2},\xi_{2};\sigma)\,\mathrm{d}\sigma, where qq obeys the hypothesis of Proposition 5.1 with

λq=λ1δ2,g(x2)=g:=x22fL+Γx22fL,M(ξ2)=λ0ξ2|ξ2|γλ0(ξ2)(μλ1δ2+μλλ2(1δ2)).\lambda_{q}=\lambda^{1-\delta_{2}},\quad g(x_{2})=g:=\|{\partial_{x_{2}}^{2}f}\|_{L^{\infty}}+\|{\Gamma\partial_{x_{2}}^{2}f}\|_{L^{\infty}},\quad M(\xi_{2})=\lambda_{0}\frac{\xi_{2}}{|{\xi_{2}}|}\gamma_{\lambda_{0}}^{\prime}(\xi_{2})\left(\frac{\mu}{{\lambda^{1-\delta_{2}}}}+\frac{\mu\lambda}{{\lambda^{2(1-\delta_{2})}}}\right).

In order to apply Proposition 5.1, note that, by (1.11) and the almost comparability of γ\gamma and ξ2ξ2γ\xi_{2}\partial_{\xi_{2}}\gamma, we have ¯<μ1M(λ)\overline{\mathcal{M}}_{<\mu^{-1}}\lesssim M(\lambda) if we choose N=1000(1+β0)N=1000(1+\beta_{0}). On the other hand, gg is constant, so g¯<μ1=g\overline{g}_{<\mu^{-1}}=g. Putting together the above observations with Propositions 3.3 and 5.1 (with s=0s=0), we arrive at

(Φ)rsθ̊,λ0P>λ1δ2,1ψ~L2\displaystyle\|{{}^{(\Phi)}r_{s_{\mathring{\theta},\lambda_{0}}P_{>\lambda^{1-\delta_{2}}},-1}\widetilde{\psi}}\|_{L^{2}} gλ0γλ0(λ)(μλ1δ2+μλλ2(1δ2))aHμ1N\displaystyle\lesssim g\lambda_{0}\gamma_{\lambda_{0}}^{\prime}(\lambda)\left(\frac{\mu}{{\lambda^{1-\delta_{2}}}}+\frac{\mu\lambda}{{\lambda^{2(1-\delta_{2})}}}\right)\|{a}\|_{H_{\mu^{-1}}^{N}}
gλ0γλ0(λ)(μλ)λ010δ2δ0a0Hμ1N0.\displaystyle\lesssim g\lambda_{0}\gamma_{\lambda_{0}}^{\prime}(\lambda)\left(\frac{\mu}{\lambda}\right)\lambda_{0}^{\frac{10\delta_{2}}{\delta_{0}}}\|{a_{0}}\|_{H_{\mu^{-1}}^{N_{0}}}.

To estimate the integral of the preceding expression on t[0,11ϵtf(τM)]t\in[0,\frac{1}{1-\epsilon}t_{f}(\tau_{M})], we use (4.3) and (6.14) to estimate

011ϵtf(τM)λ0γλ0(λ)(μλ)λ010δ2δ0dt\displaystyle\int_{0}^{\frac{1}{1-\epsilon}t_{f}(\tau_{M})}\lambda_{0}\gamma_{\lambda_{0}}^{\prime}(\lambda)\left(\frac{\mu}{\lambda}\right)\lambda_{0}^{\frac{10\delta_{2}}{\delta_{0}}}\,\mathrm{d}t (μλ)λ010δ2δ0λ0Mλ0γλ0(λ)γλ0(λ)dλ\displaystyle\lesssim\left(\frac{\mu}{\lambda}\right)\lambda_{0}^{\frac{10\delta_{2}}{\delta_{0}}}\int_{\lambda_{0}}^{M\lambda_{0}}\frac{\gamma_{\lambda_{0}}^{\prime}(\lambda)}{\gamma_{\lambda_{0}}(\lambda)}\,\mathrm{d}\lambda
λ013110δ0+10δ2δ0β0logMδ01β0λ013110δ0+10δ2δ0logλ0,\displaystyle\lesssim\lambda_{0}^{-\frac{1}{3}-\frac{1}{10}\delta_{0}+\frac{10\delta_{2}}{\delta_{0}}}\beta_{0}\log M\lesssim\delta_{0}^{-1}\beta_{0}\lambda_{0}^{-\frac{1}{3}-\frac{1}{10}\delta_{0}+\frac{10\delta_{2}}{\delta_{0}}}\log\lambda_{0},

which is acceptable.

Step 6: Contribution of rθ̊,λ0r_{\mathring{\theta},\lambda_{0}}. We begin by splitting

rθ̊,λ0(x2,D2)ψ~L2\displaystyle\|{r_{\mathring{\theta},\lambda_{0}}(x_{2},D_{2})\widetilde{\psi}}\|_{L^{2}} rθ̊,λ0(x2,D2)P<max{μ,λ0}(D2)ψ~L2\displaystyle\lesssim\|{r_{\mathring{\theta},\lambda_{0}}(x_{2},D_{2})P_{<\max\{\mu,\lambda_{0}\}}(D_{2})\widetilde{\psi}}\|_{L^{2}}
+ν>max{μ,λ0}rθ̊,λ0(x2,D2)Pν(D2)ψ~L2.\displaystyle\mathrel{\phantom{=}}+\sum_{\nu>\max\{\mu,\lambda_{0}\}}\|{r_{\mathring{\theta},\lambda_{0}}(x_{2},D_{2})P_{\nu}(D_{2})\widetilde{\psi}}\|_{L^{2}}.

For the first term, note that λ0γ(λ0,max{μ,λ0})rθ̊,λ0(x2,ξ2)P<max{μ,λ0}(ξ2)\frac{\lambda_{0}}{\gamma(\lambda_{0},\max\{\mu,\lambda_{0}\})}r_{\mathring{\theta},\lambda_{0}}(x_{2},\xi_{2})P_{<\max\{\mu,\lambda_{0}\}}(\xi_{2}) is a classical symbol of order 0. By the L2L^{2}-boundedness of its quantization, as well as Lemma 6.2, we obtain the bound

rθ̊,λ0(x2,D2)P<max{μ,λ0}(D2)ψ~L2\displaystyle\|{r_{\mathring{\theta},\lambda_{0}}(x_{2},D_{2})P_{<\max\{\mu,\lambda_{0}\}}(D_{2})\widetilde{\psi}}\|_{L^{2}} 1λ0γλ0(max{μ,λ0})(max{μ,λ0}λ)NaHμ1N,\displaystyle\lesssim\frac{1}{\lambda_{0}}\gamma_{\lambda_{0}}(\max\{\mu,\lambda_{0}\})\left(\frac{\max\{\mu,\lambda_{0}\}}{\lambda}\right)^{N}\|{a}\|_{H_{\mu^{-1}}^{N}},

which is acceptable if we take N=δ21(β0+2)N=\delta_{2}^{-1}(\beta_{0}+2), which is bounded by N0N_{0}.

For each summand in the second term, note that ν2λ0γλ0(ν)rθ̊,λ0(x2,ξ2)Pν(ξ2)\frac{\nu^{2}}{\lambda_{0}\gamma_{\lambda_{0}}(\nu)}r_{\mathring{\theta},\lambda_{0}}(x_{2},\xi_{2})P_{\nu}(\xi_{2}) is a classical symbol of order 0. As before, we therefore have

rθ̊,λ0(x2,D2)Pν(D2)ψ~L2\displaystyle\|{r_{\mathring{\theta},\lambda_{0}}(x_{2},D_{2})P_{\nu}(D_{2})\widetilde{\psi}}\|_{L^{2}} λ0ν2γλ0(ν)min{(νλ)N,(λν)N}aHμ1N.\displaystyle\lesssim\frac{\lambda_{0}}{\nu^{2}}\gamma_{\lambda_{0}}(\nu)\min\left\{\left(\frac{\nu}{\lambda}\right)^{N},\left(\frac{\lambda}{\nu}\right)^{N}\right\}\|{a}\|_{H_{\mu^{-1}}^{N}}.

By the slow variance assumption on γ\gamma, we have γλ0(ν)γλ0(λ)max{(νλ)β0,(λν)β0}\gamma_{\lambda_{0}}(\nu)\lesssim\gamma_{\lambda_{0}}(\lambda)\max\{\left(\frac{\nu}{\lambda}\right)^{\beta_{0}},\left(\frac{\lambda}{\nu}\right)^{\beta_{0}}\}. Hence, applying the preceding bound with N=1000(1+β0)N=1000(1+\beta_{0}), we obtain the summed bound

ν>μrθ̊,λ0(x2,D2)Pν(D2)ψ~L2λ0δ2λ0λ2γλ0(λ)a0Hμ1N.\displaystyle\sum_{\nu>\mu}\|{r_{\mathring{\theta},\lambda_{0}}(x_{2},D_{2})P_{\nu}(D_{2})\widetilde{\psi}}\|_{L^{2}}\lesssim\lambda_{0}^{\delta_{2}}\frac{\lambda_{0}}{\lambda^{2}}\gamma_{\lambda_{0}}(\lambda)\|{a_{0}}\|_{H_{\mu^{-1}}^{N}}.

Then

011ϵtf(τM)ν>μrθ̊,λ0(x2,D2)Pν(D2)ψ~L2dt\displaystyle\int_{0}^{\frac{1}{1-\epsilon}t_{f}(\tau_{M})}\sum_{\nu>\mu}\|{r_{\mathring{\theta},\lambda_{0}}(x_{2},D_{2})P_{\nu}(D_{2})\widetilde{\psi}}\|_{L^{2}}\,\mathrm{d}t λ0δ2011ϵtf(τM)λ0λ2γλ0(λ)dt\displaystyle\lesssim\lambda_{0}^{\delta_{2}}\int_{0}^{\frac{1}{1-\epsilon}t_{f}(\tau_{M})}\frac{\lambda_{0}}{\lambda^{2}}\gamma_{\lambda_{0}}(\lambda)\,\mathrm{d}t
λ0δ2λ0Mλ01λ2dλλ01+δ2,\displaystyle\lesssim\lambda_{0}^{\delta_{2}}\int_{\lambda_{0}}^{M\lambda_{0}}\frac{1}{\lambda^{2}}\mathrm{d}\lambda\lesssim\lambda_{0}^{-1+\delta_{2}},

which is acceptable.

Finally, we consider the case Ω=𝕋2\Omega=\mathbb{T}^{2}, in which case, by the x2x_{2}-periodicity of θ̊\mathcal{L}_{\mathring{\theta}},

θ̊φ~(x1,x2)=k(θ̊((𝕋×)φ~))(x1,x2k).\mathcal{L}_{\mathring{\theta}}\widetilde{\varphi}(x_{1},x_{2})=\sum_{k\in\mathbb{Z}}(\mathcal{L}_{\mathring{\theta}}({}^{(\mathbb{T}\times\mathbb{R})}\widetilde{\varphi}))(x_{1},x_{2}-k).

We repeat the above procedure, but now use the boundedness on L2(x24dx2)L^{2}(\langle{x_{2}}\rangle^{4}\mathrm{d}x_{2}) for classical pseudo-differential operators (for Step 1), and Proposition 5.1 with X0=X(t;x0)X_{0}=X(t;x_{0}) and s=2s=2 for Φ\Phi-conjugated remainders. Then we obtain the unit-scale-localized bounds

011ϵtf(τM)θ̊((𝕋×)φ~)L2(k,k+1)dtmin{1,1k2}λ0δ6.\int_{0}^{\frac{1}{1-\epsilon}t_{f}(\tau_{M})}\|{\mathcal{L}_{\mathring{\theta}}({}^{(\mathbb{T}\times\mathbb{R})}\widetilde{\varphi})}\|_{L^{2}(k,k+1)}\,\mathrm{d}t\lesssim\min\left\{1,\frac{1}{k^{2}}\right\}\lambda_{0}^{-\delta_{6}}. (6.18)

Summing up in kk, the desired conclusion in the periodic case follows. ∎

7 Proof of illposedness

We now establish the illposedness theorems stated in Section 1.5.

7.1 Linear illposedness

In this section, we prove Theorem A. As it is assumed in the statement of Theorem A, we are given a quadratically degenerate function ff together with parameters λ0\lambda_{0}\in\mathbb{N}, M>1M>1, 0<δ0<11000<\delta_{0}<\frac{1}{100}, 0σ013(12δ0)0\leq\sigma_{0}\leq\frac{1}{3}(1-2\delta_{0}), so that the conditions (1.18)–(1.20) hold. We also set c1=13+110δ0c_{1}=\frac{1}{3}+\frac{1}{10}\delta_{0} (cf. (6.14)). Without loss of generality, we may assume that x̊2=0\mathring{x}_{2}=0 and f(0)<0f^{\prime\prime}(0)<0. We introduce parameters Λ01\Lambda_{0}\geq 1 and 0<T010<T_{0}\leq 1, to be determined in the course of the proof, and require λ0Λ0\lambda_{0}\geq\Lambda_{0} and τMT0\tau_{M}\leq T_{0}. We also fix the regularity exponents s,ss,s^{\prime}\in\mathbb{R}.

We apply the construction of a degenerating wave packet in Sections 36 with N0=10000max{1+β0,1+|s|,1+|s|}N_{0}=10000\max\{1+\beta_{0},1+|{s}|,1+|{s^{\prime}}|\} and the above parameters; we take Λ0Λ\Lambda_{0}\geq\Lambda and T0TT_{0}\leq T. As a result, we obtain999Observe from Section 6.1 that while the constant μ\mu and bounds depend on the choice of N0N_{0}, the wave packet φ\varphi itself is independent of N0N_{0}. a degenerating wave packet φ~=Re(aei𝚽)\widetilde{\varphi}=\operatorname{Re}\left(ae^{i\mathbf{\Phi}}\right) satisfying φ~0L2(Ω)=1\|{\widetilde{\varphi}_{0}}\|_{L^{2}(\Omega)}=1, where φ~0=φ~|t=0\widetilde{\varphi}_{0}=\left.\widetilde{\varphi}\right|_{t=0}. Introducing the shorthand

|||φ~0|||:=a0H(μ01)N0,{|\kern-1.07639pt|\kern-1.07639pt|\widetilde{\varphi}_{0}|\kern-1.07639pt|\kern-1.07639pt|}:=\|{a_{0}}\|_{H_{(\mu_{0}^{-1})}^{N_{0}}},

note that |||φ~0|||1{|\kern-1.07639pt|\kern-1.07639pt|\widetilde{\varphi}_{0}|\kern-1.07639pt|\kern-1.07639pt|}\simeq 1. Moreover, in view of Proposition 6.1.(2) and the choice of m0m_{0} in Section 6.1, we have, for any s,σs^{\prime\prime},\sigma\in\mathbb{R},

λ0sγλ0(λ0)σΓσφ~0Hss,σ1.\lambda_{0}^{-s^{\prime\prime}}\gamma_{\lambda_{0}}(\lambda_{0})^{-\sigma}\|{\Gamma^{\sigma}\widetilde{\varphi}_{0}}\|_{H^{s^{\prime\prime}}}\simeq_{s^{\prime\prime},\sigma}1. (7.1)

We introduce ϕ0\phi_{0} and φ0\varphi_{0} defined by the relations

ϕ0(x)=Γ12φ0(x),φ0(x)=φ~0(x).\phi_{0}(x)=\Gamma^{-\frac{1}{2}}\varphi_{0}(x),\quad\varphi_{0}(x)=\widetilde{\varphi}_{0}(x). (7.2)

We now proceed to prove (1.22). Let ϕ\phi be an Γ12L2\Gamma^{-\frac{1}{2}}L^{2}-solution to Lθ̊ϕ=0L_{\mathring{\theta}}\phi=0 on [0,10099tf(τM)]\left[0,\frac{100}{99}t_{f}(\tau_{M})\right] with ϕ|t=0=ϕ0\left.\phi\right|_{t=0}=\phi_{0}. Then φ=Γ12ϕ\varphi=\Gamma^{\frac{1}{2}}\phi is an L2L^{2}-solution to θ̊φ=0\mathcal{L}_{\mathring{\theta}}\varphi=0 on the same interval with φ|t=0=Γ12ϕ0=φ~0(x)\left.\varphi\right|_{t=0}=\Gamma^{\frac{1}{2}}\phi_{0}=\widetilde{\varphi}_{0}(x). With the simplified notation t=11ϵtf(τM)t^{\ast}=\frac{1}{1-\epsilon}t_{f}(\tau_{M}), note that t10099tf(τM)t^{\ast}\leq\frac{100}{99}t_{f}(\tau_{M}) by (4.3). In view of (7.1) and (7.2), it suffices to establish

supt[0,t]Γ12φ(t,)HsCs1γλ0(Mλ0)12Msλ0s\begin{split}\sup_{t\in[0,t^{\ast}]}\|{\Gamma^{-\frac{1}{2}}\varphi(t,\cdot)}\|_{H^{s^{\prime}}}\geq C_{s^{\prime}}\frac{1}{\gamma_{\lambda_{0}}(M\lambda_{0})^{\frac{1}{2}}}M^{s^{\prime}}\lambda_{0}^{s^{\prime}}\end{split}

for any s>0s^{\prime}>0.

Recall that φ~=Re(ei𝚽a)\widetilde{\varphi}=\operatorname{Re}(e^{i\mathbf{\Phi}}a) is well-defined on [0,t]\left[0,t^{\ast}\right] and belongs to Cc(Ω)C^{\infty}_{c}(\Omega). Therefore, the generalized energy inequality

|ddtφ,φ~θ̊φ,φ~φ,θ̊φ~|φL2φ~L2\begin{split}\left|\frac{\mathrm{d}}{\mathrm{d}t}\langle{\varphi,\widetilde{\varphi}}\rangle-\langle{\mathcal{L}_{\mathring{\theta}}\varphi,\widetilde{\varphi}}\rangle-\langle{\varphi,\mathcal{L}_{\mathring{\theta}}\widetilde{\varphi}}\rangle\right|\lesssim\|{\varphi}\|_{L^{2}}\|{\widetilde{\varphi}}\|_{L^{2}}\end{split}

can be justified on the time interval [0,t]\left[0,t^{\ast}\right]. Therefore, we immediately obtain the inequality

|ddtφ,φ~|(φ~L2+θ̊φ~L2)φL2.\begin{split}\left|\frac{\mathrm{d}}{\mathrm{d}t}\langle{\varphi,\widetilde{\varphi}}\rangle\right|\lesssim(\|{\widetilde{\varphi}}\|_{L^{2}}+\|{\mathcal{L}_{\mathring{\theta}}\widetilde{\varphi}}\|_{L^{2}})\|{\varphi}\|_{L^{2}}.\end{split}

Integrating in time and applying the error estimate (6.17) gives

|φ,φ~(t)φ,φ~(0)|0t(φ~L2+θ̊φ~L2)φL2dt(tφ~0L2+λ0δ6|||φ~0|||)φ0L2\begin{split}\left|\langle{\varphi,\widetilde{\varphi}}\rangle(t)-\langle{\varphi,\widetilde{\varphi}}\rangle(0)\right|&\lesssim\int_{0}^{t}(\|{\widetilde{\varphi}}\|_{L^{2}}+\|{\mathcal{L}_{\mathring{\theta}}\widetilde{\varphi}}\|_{L^{2}})\|{\varphi}\|_{L^{2}}\,\mathrm{d}t^{\prime}\\ &\lesssim(t\|{\widetilde{\varphi}_{0}}\|_{L^{2}}+\lambda_{0}^{-\delta_{6}}{|\kern-1.07639pt|\kern-1.07639pt|\widetilde{\varphi}_{0}|\kern-1.07639pt|\kern-1.07639pt|})\|{\varphi_{0}}\|_{L^{2}}\end{split}

for any 0<tt0<t\leq t^{\ast}. In the last inequality we have used φ~L2φ~0L2\|{\widetilde{\varphi}}\|_{L^{2}}\lesssim\|{\widetilde{\varphi}_{0}}\|_{L^{2}} and φL([0,t];L2)φ0L2\|{\varphi}\|_{L^{\infty}([0,t^{\ast}];L^{2})}\lesssim\|{\varphi_{0}}\|_{L^{2}}. Note that in the above inequalities, the implicit constants depend only on f,γf,\gamma. Therefore, taking Λ0\Lambda_{0} larger and the constant T0T_{0} smaller if necessary (the latter in a way depending only on f,γf,\gamma), we may guarantee that

|φ,φ~(t)φ,φ~(0)|14φ0L2φ~0L2,\begin{split}\left|\langle{\varphi,\widetilde{\varphi}}\rangle(t^{\ast})-\langle{\varphi,\widetilde{\varphi}}\rangle(0)\right|\leq\frac{1}{4}\|{\varphi_{0}}\|_{L^{2}}\|{\widetilde{\varphi}_{0}}\|_{L^{2}},\end{split}

which gives, recalling φ|t=0=φ~|t=0=φ~0\varphi|_{t=0}=\widetilde{\varphi}|_{t=0}=\widetilde{\varphi}_{0},

φ,φ~(t)>14φ0L2φ~0L2.\begin{split}\langle{\varphi,\widetilde{\varphi}}\rangle(t^{\ast})>\frac{1}{4}\|{\varphi_{0}}\|_{L^{2}}\|{\widetilde{\varphi}_{0}}\|_{L^{2}}.\end{split}

Furthermore, from φ~=φ~main+φ~small\widetilde{\varphi}=\widetilde{\varphi}^{main}+\widetilde{\varphi}^{small} and the estimate for φ~small\widetilde{\varphi}^{small} in (6.5) and (6.14),

φ,φ~main(t)14φ0L2φ~0L2φ0L2φ~small(t)L214φ0L2φ~0L2Cλ0c1φ0L2|||φ~0|||18φ0L2φ~0L2,\begin{split}\langle{\varphi,\widetilde{\varphi}^{main}}\rangle(t^{\ast})&\geq\frac{1}{4}\|{\varphi_{0}}\|_{L^{2}}\|{\widetilde{\varphi}_{0}}\|_{L^{2}}-\|{\varphi_{0}}\|_{L^{2}}\|{\widetilde{\varphi}^{small}(t^{\ast})}\|_{L^{2}}\\ &\geq\frac{1}{4}\|{\varphi_{0}}\|_{L^{2}}\|{\widetilde{\varphi}_{0}}\|_{L^{2}}-C\lambda_{0}^{-c_{1}}\|{\varphi_{0}}\|_{L^{2}}{|\kern-1.07639pt|\kern-1.07639pt|\widetilde{\varphi}_{0}|\kern-1.07639pt|\kern-1.07639pt|}\geq\frac{1}{8}\|{\varphi_{0}}\|_{L^{2}}\|{\widetilde{\varphi}_{0}}\|_{L^{2}},\end{split}

by again taking Λ0\Lambda_{0} larger if necessary. Combining this inequality with φ~0L2=1\|{\widetilde{\varphi}_{0}}\|_{L^{2}}=1,

φ,φ~main(t)Γ12φ(t)HsΓ12φ~main(t)Hs,\begin{split}\langle{\varphi,\widetilde{\varphi}^{main}}\rangle(t^{\ast})\leq\|{\Gamma^{-\frac{1}{2}}\varphi(t^{\ast})}\|_{H^{s^{\prime}}}\|{\Gamma^{\frac{1}{2}}\widetilde{\varphi}^{main}(t^{\ast})}\|_{H^{-s^{\prime}}},\end{split}

and (6.4), we conclude that

ϕ(t)Hs=Γ12φ(t)HsCs1γλ0(λ(t))12λ(t)sφ0L2Cs1γλ0(Mλ0)12Msλ0s.\begin{split}\|{\phi(t^{\ast})}\|_{H^{s^{\prime}}}=\|{\Gamma^{-\frac{1}{2}}\varphi(t^{\ast})}\|_{H^{s^{\prime}}}\geq C_{s^{\prime}}\frac{1}{\gamma_{\lambda_{0}}(\lambda(t^{\ast}))^{\frac{1}{2}}}\lambda(t^{\ast})^{s^{\prime}}\|{\varphi_{0}}\|_{L^{2}}\geq C_{s^{\prime}}\frac{1}{\gamma_{\lambda_{0}}(M\lambda_{0})^{\frac{1}{2}}}M^{s^{\prime}}\lambda_{0}^{s^{\prime}}.\end{split}

In the last inequality, we have used (1.21) and that λ(t)=Mλ0\lambda(t^{\ast})=M\lambda_{0} (see Lemma 4.3). This completes the proof of Theorem A. ∎

7.2 Linear illposedness, dissipative case

In this section, we proceed to prove Theorem B. Let f0f_{0} be a smooth function satisfying the assumptions in the statement of Theorem B. Without loss of generality, we further assume that f0(0)<0f^{\prime\prime}_{0}(0)<0. Recall that f(t,x2)f(t,x_{2}) for t0t\geq 0 was defined to be the solution of tf+κΥf=0\partial_{t}f+\kappa\Upsilon f=0. From the assumption that the symbol of Υ\Upsilon is even, we have that f(t,)f(t,\cdot) is even as well. In particular, f(t,0)=0f^{\prime}(t,0)=0 for all t0t\geq 0. We are also given parameters λ0\lambda_{0}\in\mathbb{N}, M>1M>1, 0<δ1δ0<11000<\delta_{1}\leq\delta_{0}<\frac{1}{100} and 0σ013(12δ0)0\leq\sigma_{0}\leq\frac{1}{3}(1-2\delta_{0}), so that the conditions (1.18)–(1.20) as well as (1.24)–(1.25) hold. We also set c1=13+110δ0c_{1}=\frac{1}{3}+\frac{1}{10}\delta_{0} (cf. (6.14)). Let Λ11\Lambda_{1}\geq 1 and 0<T110<T_{1}\leq 1 be parameters to be determined below, and we also require that λ0Λ1\lambda_{0}\geq\Lambda_{1} and τMT1\tau_{M}\leq T_{1}. Fix also the regularity exponents s,ss,s^{\prime}\in\mathbb{R}.

We apply the construction in Sections 36 with N0=104max{1+β0,1+α0,1+|s|,1+|s|}N_{0}=10^{4}\max\{1+\beta_{0},1+\alpha_{0},1+|{s}|,1+|{s^{\prime}}|\} and the above parameters; we take Λ1Λ\Lambda_{1}\leq\Lambda and T1TT_{1}\leq T. As in the proof of Theorem A, we obtain a L2L^{2}-normalized degenerating wave packet φ~=Re(aei𝚽)\widetilde{\varphi}=\operatorname{Re}(ae^{i\mathbf{\Phi}}); we set ϕ0=Γ12φ~0\phi_{0}=\Gamma^{-\frac{1}{2}}\widetilde{\varphi}_{0}, where φ~0=φ~|t=0\widetilde{\varphi}_{0}=\left.\widetilde{\varphi}\right|_{t=0}.

Let ϕ\phi be a Γ12L2\Gamma^{-\frac{1}{2}}L^{2}-solution to Lθ̊(κ)ϕ=0L_{\mathring{\theta}}^{(\kappa)}\phi=0 on [0,10099tf(τM)]\left[0,\frac{100}{99}t_{f}(\tau_{M})\right] with ϕ|t=0=ϕ0\phi|_{t=0}=\phi_{0}, so that φ:=Γ12ϕ\varphi:=\Gamma^{\frac{1}{2}}\phi is an L2L^{2}-solution to θ̊(κ)φ=0\mathcal{L}_{\mathring{\theta}}^{(\kappa)}\varphi=0 with φ|t=0=φ0\left.\varphi\right|_{t=0}=\varphi_{0}. We also introduce t=11ϵtf(τM)t^{\ast}=\frac{1}{1-\epsilon}t_{f}(\tau_{M}), which obeys t10099tf(τM)t^{\ast}\leq\frac{100}{99}t_{f}(\tau_{M}). As before, we apply the generalized energy inequality to φ\varphi and φ~\widetilde{\varphi} and obtain

|ddtφ,φ~(κ)θ̊φ,φ~φ,θ̊(κ)φ~|φL2φ~L2,\begin{split}\left|\frac{\mathrm{d}}{\mathrm{d}t}\langle{\varphi,\widetilde{\varphi}}\rangle-\langle{\mathcal{L}^{(\kappa)}_{\mathring{\theta}}\varphi,\widetilde{\varphi}}\rangle-\langle{\varphi,\mathcal{L}_{\mathring{\theta}}^{(\kappa)}\widetilde{\varphi}}\rangle\right|\lesssim\|{\varphi}\|_{L^{2}}\|{\widetilde{\varphi}}\|_{L^{2}},\end{split}

which is valid on the time interval [0,t]\left[0,t^{\ast}\right]. This time, we obtain the inequality

|ddtφ,φ~|(φ~L2+θ̊φ~L2+κΥφ~L2)φL2.\begin{split}\left|\frac{\mathrm{d}}{\mathrm{d}t}\langle{\varphi,\widetilde{\varphi}}\rangle\right|\lesssim(\|{\widetilde{\varphi}}\|_{L^{2}}+\|{\mathcal{L}_{\mathring{\theta}}\widetilde{\varphi}}\|_{L^{2}}+\kappa\|{\Upsilon\widetilde{\varphi}}\|_{L^{2}})\|{\varphi}\|_{L^{2}}.\end{split} (7.3)

Since N0104(1+α0)N_{0}\geq 10^{4}(1+\alpha_{0}), arguing as in the proof of Proposition 6.1, we have

κΥφ~L2κυ(λ0,λ)|||φ~0|||,\kappa\|{\Upsilon\widetilde{\varphi}}\|_{L^{2}}\lesssim\kappa\upsilon(\lambda_{0},\lambda){|\kern-1.07639pt|\kern-1.07639pt|\widetilde{\varphi}_{0}|\kern-1.07639pt|\kern-1.07639pt|},

where |||φ~0|||{|\kern-1.07639pt|\kern-1.07639pt|\widetilde{\varphi}_{0}|\kern-1.07639pt|\kern-1.07639pt|} is defined as in the proof of Theorem A. Applying the previous bound to (7.3) and proceeding as in the inviscid case gives

|φ,φ~(t)φ,φ~(0)|(tφ~0L2+(λ0δ6+0t|υ(λ0,λ(t))|dt)|||φ~0|||)φ0L2\begin{split}\left|\langle{\varphi,\widetilde{\varphi}}\rangle(t)-\langle{\varphi,\widetilde{\varphi}}\rangle(0)\right|&\lesssim\left(t\|{\widetilde{\varphi}_{0}}\|_{L^{2}}+\left(\lambda_{0}^{-\delta_{6}}+\int_{0}^{t}|\upsilon(\lambda_{0},\lambda(t^{\prime}))|\,\mathrm{d}t^{\prime}\right){|\kern-1.07639pt|\kern-1.07639pt|\widetilde{\varphi}_{0}|\kern-1.07639pt|\kern-1.07639pt|}\right)\|{\varphi_{0}}\|_{L^{2}}\end{split}

for any 0<tt0<t\leq t^{\ast}. Taking t=tt=t^{\ast} and using λ˙=(1ϵ)|f(0,0)|λ0γ(λ0,λ)\dot{\lambda}=(1-\epsilon)|{f^{\prime\prime}(0,0)}|\lambda_{0}\gamma(\lambda_{0},\lambda) with a change of variables, we have that

0t|υ(λ0,λ(t))|dtλ0Mλ0υ(λ0,λ)γ(λ0,λ)dλλ0.\begin{split}\int_{0}^{t^{\ast}}|\upsilon(\lambda_{0},\lambda(t))|\,\mathrm{d}t\lesssim\int_{\lambda_{0}}^{M\lambda_{0}}\frac{\upsilon(\lambda_{0},\lambda)}{\gamma(\lambda_{0},\lambda)}\frac{\mathrm{d}\lambda}{\lambda_{0}}.\end{split}

Under the assumption (1.24), taking Λ1\Lambda_{1} larger if necessary, we can conclude that

φ,φ~(t)116φ0L2φ~0L2.\begin{split}\langle{\varphi,\widetilde{\varphi}}\rangle(t^{\ast})\geq\frac{1}{16}\|{\varphi_{0}}\|_{L^{2}}\|{\widetilde{\varphi}_{0}}\|_{L^{2}}.\end{split}

The rest of the proof is identical to the inviscid case. ∎

7.3 Nonlinear illposedness

In this section, we will establish Theorem C. We only consider the inviscid case, as the dissipative case can be treated similarly.

We assume the same hypothesis and conventions as in the proof of Theorem A in the beginning of Section 7.1. Given ϵ>0\epsilon>0 and θ̊=f(x2)\mathring{\theta}=f(x_{2}) in the statement of the Theorem C, we consider the sequence

ϑ0(λ0):=cϵγ(λ0,λ0)12λ0sΓ12φ~0Cc(Ω),\begin{split}{\vartheta_{0}^{(\lambda_{0})}}:=c\epsilon{\gamma(\lambda_{0},\lambda_{0})^{\frac{1}{2}}\lambda_{0}^{-s}\Gamma^{-\frac{1}{2}}\widetilde{\varphi}_{0}}\in C^{\infty}_{c}(\Omega),\end{split}

where φ~0=Re(ei𝚽a)|t=0\widetilde{\varphi}_{0}=\operatorname{Re}(e^{i\mathbf{\Phi}}a)|_{t=0} is the L2L^{2}-normalized degenerating wave packet at the initial time with frequency λ0\lambda_{0} from the proof of Theorem A. Here, c>0c>0 is an absolute constant inserted to guarantee that

ϑ0(λ0)Hsϵ\begin{split}\|{{\vartheta_{0}^{(\lambda_{0})}}}\|_{H^{s}}\leq\epsilon\end{split}

uniformly for λ0Λ0\lambda_{0}\geq\Lambda_{0} (the same Λ0\Lambda_{0} from Theorem A), as it is required in the statement of the Theorem C. We then consider the sequence initial data

θ(λ0)|t=0=θ̊+ϑ0(λ0)\begin{split}\theta^{(\lambda_{0})}|_{t=0}=\mathring{\theta}+{\vartheta_{0}^{(\lambda_{0})}}\end{split}

with λ0Λ0\lambda_{0}\geq\Lambda_{0} for (1.1). Recalling the parameters δ,A>0\delta,A>0 given in the statement, we may assume towards a contradiction that for any sufficiently large λ0\lambda_{0}, there exists a solution θ(λ0)\theta^{(\lambda_{0})} to (1.1) with initial data θ(λ0)|t=0\theta^{(\lambda_{0})}|_{t=0} satisfying θ(λ0)θ̊L([0,δ];Hs)\theta^{(\lambda_{0})}-\mathring{\theta}\in L^{\infty}([0,\delta];H^{s^{\prime}}) and

supt[0,δ](θ(λ0)θ̊)(t,)HsA.\begin{split}\sup_{t\in[0,\delta]}\|{(\theta^{(\lambda_{0})}-\mathring{\theta})(t,\cdot)}\|_{H^{s^{\prime}}}\leq A.\end{split}

From now on, we shall fix some large λ0Λ0\lambda_{0}\geq\Lambda_{0} and omit writing out the dependence of λ0\lambda_{0} on the solution. On the time interval [0,δ][0,\delta], we simply define

φ(t,)=Γ12(θ(λ0)θ̊)(t,).\begin{split}\varphi^{\star}(t,\cdot)=\Gamma^{\frac{1}{2}}(\theta^{(\lambda_{0})}-\mathring{\theta})(t,\cdot).\end{split}

Since θ(λ0)\theta^{(\lambda_{0})} and θ̊\mathring{\theta} are solutions to (1.1), we see that the equation for φ\varphi^{\star} is given by

θ̊φ=Γ12(Γ12φΓ12φ)=:𝒩[φ].\begin{split}\mathcal{L}_{\mathring{\theta}}\varphi^{\star}=-\Gamma^{\frac{1}{2}}\left(\nabla^{\perp}\Gamma^{\frac{1}{2}}\varphi^{\star}\cdot\nabla\Gamma^{-\frac{1}{2}}\varphi^{\star}\right)=:\mathcal{N}[\varphi^{\star}].\end{split}

We claim that

φ0,φ~0\displaystyle\langle{\varphi^{\star}_{0},\widetilde{\varphi}_{0}}\rangle φ0L2,\displaystyle\simeq\|{\varphi^{\star}_{0}}\|_{L^{2}}, (7.4)
𝒩[φ]L2\displaystyle\|{\mathcal{N}[\varphi^{\star}]}\|_{L^{2}} C(1+A)φL2.\displaystyle\leq C(1+A)\|{\varphi^{\star}}\|_{L^{2}}. (7.5)

Bound (7.4) follows from Proposition 6.1.(2). To prove (7.5), we use s>32β0+3s^{\prime}>\frac{3}{2}\beta_{0}+3. We take some 0<ε<s(3+β0)0<\varepsilon<s^{\prime}-(3+\beta_{0}) and estimate

𝒩[φ]L2CΓ12φΓ12φHβ02C(Γ12φHβ02Γ12φL+Γ12φLΓ12φHβ02)CεφL21β0+1sφHsβ0+1sφL212+εsφHs2+εs+CεφL21β02+1sφHsβ02+1sφL212+ε+β02sφHs2+ε+β02sCε(1+A)φL2.\begin{split}\|{\mathcal{N}[\varphi^{\star}]}\|_{L^{2}}&\leq C\|{\nabla^{\perp}\Gamma^{\frac{1}{2}}\varphi^{\star}\cdot\nabla\Gamma^{-\frac{1}{2}}\varphi^{\star}}\|_{H^{\frac{\beta_{0}}{2}}}\\ &\leq C(\|{\nabla^{\perp}\Gamma^{\frac{1}{2}}\varphi^{\star}}\|_{H^{\frac{\beta_{0}}{2}}}\|{\nabla\Gamma^{-\frac{1}{2}}\varphi^{\star}}\|_{L^{\infty}}+\|{\nabla^{\perp}\Gamma^{\frac{1}{2}}\varphi^{\star}}\|_{L^{\infty}}\|{\nabla\Gamma^{-\frac{1}{2}}\varphi^{\star}}\|_{H^{\frac{\beta_{0}}{2}}})\\ &\leq C_{\varepsilon}\|{\varphi^{\star}}\|_{L^{2}}^{1-\frac{\beta_{0}+1}{s^{\prime}}}\|{\varphi^{\star}}\|_{H^{s^{\prime}}}^{\frac{\beta_{0}+1}{s^{\prime}}}\|{\varphi^{\star}}\|_{L^{2}}^{1-\frac{2+\varepsilon}{s^{\prime}}}\|{\varphi^{\star}}\|_{H^{s^{\prime}}}^{\frac{2+\varepsilon}{s^{\prime}}}\\ &\qquad+C_{\varepsilon}\|{\varphi^{\star}}\|_{L^{2}}^{1-\frac{\frac{\beta_{0}}{2}+1}{s^{\prime}}}\|{\varphi^{\star}}\|_{H^{s^{\prime}}}^{\frac{\frac{\beta_{0}}{2}+1}{s^{\prime}}}\|{\varphi^{\star}}\|_{L^{2}}^{1-\frac{2+\varepsilon+\frac{\beta_{0}}{2}}{s^{\prime}}}\|{\varphi^{\star}}\|_{H^{s^{\prime}}}^{\frac{2+\varepsilon+\frac{\beta_{0}}{2}}{s^{\prime}}}\\ &\leq C_{\varepsilon}(1+A)\|{\varphi^{\star}}\|_{L^{2}}.\end{split}

In this chain of estimates, we used Γ12hHsshHs+β2\|{\Gamma^{\frac{1}{2}}h}\|_{H^{s^{\prime\prime}}}\lesssim_{s^{\prime\prime}}\|{h}\|_{H^{s^{\prime\prime}+\frac{\beta}{2}}} and Γ12hHsshHs\|{\Gamma^{-\frac{1}{2}}h}\|_{H^{s^{\prime\prime}}}\lesssim_{s^{\prime\prime}}\|{h}\|_{H^{s^{\prime\prime}}}, which follow the assumptions on Γ\Gamma and Littlewood–Paley decomposition, as well as the Sobolev product estimate and interpolation estimate. This verifies the claim.

Using the energy structure of θ̊\mathcal{L}_{\mathring{\theta}}, as well as the cancellation 𝒩[φ],φ=0\langle{\mathcal{N}[\varphi^{\star}],\varphi^{\star}}\rangle=0, we obtain

|ddtφ,φ|φL22.\begin{split}\left|\frac{\mathrm{d}}{\mathrm{d}t}\langle{\varphi^{\star},\varphi^{\star}}\rangle\right|\lesssim{\|{\varphi^{\star}}\|_{L^{2}}^{2}}.\end{split}

Hence, by Gronwall’s inquality,

φ(t)L2φ0L2exp(Ct)\begin{split}\|{\varphi^{\star}(t)}\|_{L^{2}}\leq\|{\varphi^{\star}_{0}}\|_{L^{2}}\exp({Ct})\end{split} (7.6)

for 0tδ0\leq t\leq\delta with C>0C>0 depending only on ff and Γ\Gamma. Using the generalized energy inequality together with (7.5) and (7.6),

|ddtφ,φ~|(φ~L2+θ̊φ~L2)(φL2+𝒩[φ]L2)(1+A)exp(Ct)(φ~L2+θ̊φ~L2)φ0L2.\begin{split}\left|\frac{\mathrm{d}}{\mathrm{d}t}\langle{\varphi^{\star},\widetilde{\varphi}}\rangle\right|&\lesssim(\|{\widetilde{\varphi}}\|_{L^{2}}+\|{\mathcal{L}_{\mathring{\theta}}\widetilde{\varphi}}\|_{L^{2}})(\|{\varphi^{\star}}\|_{L^{2}}+\|{\mathcal{N}[\varphi^{\star}]}\|_{L^{2}})\\ &\lesssim(1+A)\exp({Ct})(\|{\widetilde{\varphi}}\|_{L^{2}}+\|{\mathcal{L}_{\mathring{\theta}}\widetilde{\varphi}}\|_{L^{2}})\|{\varphi^{\star}_{0}}\|_{L^{2}}.\end{split} (7.7)

Taking tδt\leq{\delta} and integrating in time, we obtain similarly as in the proof of linear illposedness that

φ,φ~(t)φ0,φ~0C(1+A)(tφ~0L2+λ0δ6|||φ~0|||)φ0L2.\begin{split}\langle{\varphi^{\star},\widetilde{\varphi}}\rangle(t)\geq\langle{\varphi^{\star}_{0},\widetilde{\varphi}_{0}}\rangle-C(1+A)\left(t\|{\widetilde{\varphi}_{0}}\|_{L^{2}}+\lambda_{0}^{-\delta_{6}}{|\kern-1.07639pt|\kern-1.07639pt|\widetilde{\varphi}_{0}|\kern-1.07639pt|\kern-1.07639pt|}\right)\|{\varphi^{\star}_{0}}\|_{L^{2}}.\end{split} (7.8)

At this point, from the definition of HsH^{s}-HsH^{s^{\prime}} instability (Definition 1.12), we have a sequence (M(n),λ0(n))(M_{(n)},\lambda_{0(n)}) satisfying the growth conditions (1.18)–(1.20) (for some σ0(n)\sigma_{0(n)}), τM(n)0\tau_{M_{(n)}}\to 0, and γ(λ0(n),λ0(n))12γ(λ0(n),M(n)λ0(n))12M(n)sλ0(n)ss\frac{\gamma(\lambda_{0(n)},\lambda_{0(n)})^{\frac{1}{2}}}{\gamma(\lambda_{0(n)},M_{(n)}\lambda_{0(n)})^{\frac{1}{2}}}M_{(n)}^{s^{\prime}}\lambda_{0(n)}^{s^{\prime}-s}\to\infty. Therefore, by taking nn to be sufficiently large, we have

t(n):=11ϵtf(τM(n))<min{δ,110C(1+A)},C(1+A)λ0(n)δ6|||φ~0|||<110.\begin{split}t^{\ast}_{(n)}:=\frac{1}{1-\epsilon}t_{f}(\tau_{M_{(n)}})<\min\{\delta,\tfrac{1}{10C(1+A)}\},\quad C(1+A)\lambda_{0(n)}^{-\delta_{6}}{|\kern-1.07639pt|\kern-1.07639pt|\widetilde{\varphi}_{0}|\kern-1.07639pt|\kern-1.07639pt|}<\frac{1}{10}.\end{split}

Here, CC is the constant from (7.8) which is independent of λ0\lambda_{0}. From (7.8), we now deduce that

φ,φ~(t(n))12φ0L2.\begin{split}\langle{\varphi^{\star},\widetilde{\varphi}}\rangle(t^{\ast}_{(n)})\geq\frac{1}{2}\|{\varphi^{\star}_{0}}\|_{L^{2}}.\end{split}

By Proposition 6.1, Cauchy–Schwartz and the HsH^{s^{\prime}}-HsH^{-s^{\prime}} duality, we may decompose

φ,φ~(t(n))Γ12φ(t(n))HsΓ12φ~main(t(n))Hs+φ(t(n))L2φ~small(t(n))L2.\begin{split}\langle{\varphi^{\star},\widetilde{\varphi}}\rangle(t^{\ast}_{(n)})\leq\|{\Gamma^{-\frac{1}{2}}\varphi^{\star}(t^{\ast}_{(n)})}\|_{H^{s}}\|{\Gamma^{\frac{1}{2}}\widetilde{\varphi}^{main}(t^{\ast}_{(n)})}\|_{H^{-s}}+\|{\varphi^{\star}(t^{\ast}_{(n)})}\|_{L^{2}}\|{\widetilde{\varphi}^{small}(t^{\ast}_{(n)})}\|_{L^{2}}.\end{split}

where

φ(t(n))L2φ~small(t(n))L2<110φ0L2.\begin{split}\|{\varphi^{\star}(t^{\ast}_{(n)})}\|_{L^{2}}\|{\widetilde{\varphi}^{small}(t^{\ast}_{(n)})}\|_{L^{2}}<\frac{1}{10}\|{\varphi^{\star}_{0}}\|_{L^{2}}.\end{split}

By (6.4) in Proposition 6.1, we obtain

Γ12φ(t(n))Hsφ0L2Γ12φ~main(t(n))Hsϵγ(λ0(n),λ0(n))12γ(λ0(n),M(n)λ0(n))12M(n)sλ0(n)ss.\begin{split}\|{\Gamma^{-\frac{1}{2}}\varphi^{\star}(t^{\ast}_{(n)})}\|_{H^{s^{\prime}}}&\gtrsim\frac{\|{\varphi^{\star}_{0}}\|_{L^{2}}}{\|{\Gamma^{\frac{1}{2}}\widetilde{\varphi}^{main}(t^{\ast}_{(n)})}\|_{H^{-{s^{\prime}}}}}\gtrsim\epsilon\frac{\gamma(\lambda_{0(n)},\lambda_{0(n)})^{\frac{1}{2}}}{\gamma(\lambda_{0(n)},M_{(n)}\lambda_{0(n)})^{\frac{1}{2}}}M_{(n)}^{s^{\prime}}\lambda_{0(n)}^{s^{\prime}-s}.\end{split} (7.9)

By HsH^{s}-HsH^{s^{\prime}} instability (see Definition 1.12), the RHS becomes arbitrary large as nn\to\infty. On the other hand, using the definition of φ\varphi^{\star}, we obtain

Aθ(t(n))Hs=Γ12φ(t(n))+θ̊HsΓ12φ(t(n))Hsθ̊Hs\begin{split}A\geq\|{\theta(t^{\ast}_{(n)})}\|_{H^{s^{\prime}}}=\|{\Gamma^{-\frac{1}{2}}\varphi^{\star}(t^{\ast}_{(n)})+\mathring{\theta}}\|_{H^{s^{\prime}}}\geq\|{\Gamma^{-\frac{1}{2}}\varphi^{\star}(t^{\ast}_{(n)})}\|_{H^{s^{\prime}}}-\|{\mathring{\theta}}\|_{H^{s^{\prime}}}\end{split}

which is a contradiction with (7.9) as we take nn\to\infty. The proof is now complete. ∎

7.4 Nonexistence

The goal of this section is to prove Theorem D. We proceed in several steps.

1. Choice of the initial data. To begin with, we fix some f0(x2)f_{0}(x_{2}) satisfying the following properties:

  • f0f_{0} is CC^{\infty} smooth and supported in [2,2][-2,2];

  • f0(x2)=12x22f_{0}(x_{2})=-\frac{1}{2}x_{2}^{2} for |x2|1|x_{2}|\leq 1.

For λ0Λ0\lambda_{0}\geq\Lambda_{0}, let us denote by φ~(λ0)(t)\widetilde{\varphi}^{(\lambda_{0})}(t) the degenerate wave packet solution adapted to f0f_{0} with frequency λ0\lambda_{0}, normalized in L2L^{2}. Furthermore, we assume that the wave packet at the initial time is supported in (12,1](\frac{1}{2},1]. For convenience, we recall a few essential properties of φ~(λ0)(t)\widetilde{\varphi}^{(\lambda_{0})}(t). Define M(λ0):=min{γ(λ0,λ0)1δ0,λ0δ0,γ(λ0,λ0)12δ0β0,λ015δ03β0}M(\lambda_{0}):=\min\left\{\gamma(\lambda_{0},\lambda_{0})^{1-\delta_{0}},\lambda_{0}^{\delta_{0}},\,\gamma(\lambda_{0},\lambda_{0})^{\frac{1-2\delta_{0}}{\beta_{0}}},\lambda_{0}^{\frac{1-5\delta_{0}}{3\beta_{0}}}\right\} and τ(λ0):=λ0M(λ0)λ01γ(λ0,λ)dλλ0\tau(\lambda_{0}):=\int_{\lambda_{0}}^{M(\lambda_{0})\lambda_{0}}\frac{1}{\gamma(\lambda_{0},\lambda)}\frac{\mathrm{d}\lambda}{\lambda_{0}}.

  • Degeneration estimate: there is a decomposition φ~(λ0)=φ~(λ0),main+φ~(λ0),small\widetilde{\varphi}^{(\lambda_{0})}=\widetilde{\varphi}^{(\lambda_{0}),main}+\widetilde{\varphi}^{(\lambda_{0}),small} such that for t[0,τ(λ0)]t\in[0,\tau(\lambda_{0})],

    Γ12φ~(λ0),main(t)HsCsγ(λ0,λ(t))12λ(t)s,φ~(λ0),small(t)L2Cλ(t)c1.\begin{split}\|{\Gamma^{\frac{1}{2}}\widetilde{\varphi}^{(\lambda_{0}),main}(t)}\|_{H^{-s}}\leq C_{-s}\gamma(\lambda_{0},\lambda(t))^{\frac{1}{2}}\lambda(t)^{-s},\qquad\|{\widetilde{\varphi}^{(\lambda_{0}),small}(t)}\|_{L^{2}}\leq C\lambda(t)^{-c_{1}}.\end{split}
  • Error estimate: f0φ~(λ0)L1([0,10099tf0(τM)];L2)Cλ0δ6\|{\mathcal{L}_{f_{0}}\widetilde{\varphi}^{(\lambda_{0})}}\|_{L^{1}([0,\frac{100}{99}t_{f_{0}}(\tau_{M})];L^{2})}\leq C\lambda_{0}^{-\delta_{6}} and f0φ~(λ0)L1([0,10099tf0(τM)];L2[k,k+1])Cmin{1,k2}λ0δ6\|{\mathcal{L}_{f_{0}}\widetilde{\varphi}^{(\lambda_{0})}}\|_{L^{1}([0,\frac{100}{99}t_{f_{0}}(\tau_{M})];L^{2}[k,k+1])}\leq C\min\{1,k^{-2}\}\lambda_{0}^{-\delta_{6}} (see (6.17) and (6.18)).

By λ(t)\lambda(t), we mean the solution of (4.16) with f=f0f=f_{0}, which verifies λ(τ(λ0))M(λ0)λ0\lambda(\tau(\lambda_{0}))\geq M(\lambda_{0})\lambda_{0}. In the above, it is important that the positive constants CC, c1c_{1} and δ6\delta_{6} do not depend on λ0\lambda_{0}.

We now take the following shear steady state:

θ̊(x2)=k=k0fk:=k=k0akf0(x2yk)\begin{split}\mathring{\theta}(x_{2})=\sum_{k=k_{0}}^{\infty}f_{k}:=\sum_{k=k_{0}}^{\infty}a_{k}f_{0}(x_{2}-y_{k})\end{split}

where {ak}k1\{a_{k}\}_{k\geq 1} can be any square summable sequence; we fix it to be ak=k2a_{k}=k^{-2} for simplicity. On the other hand, {yk}k1\{y_{k}\}_{k\geq 1} is a strictly increasing sequence to be determined; y1=1y_{1}=1 and we shall take yky_{k} sufficiently large with respect to yk1y_{k-1} for k2k\geq 2. By taking k01k_{0}\geq 1 large, it is guaranteed that θ̊Hs<ϵ2\|{\mathring{\theta}}\|_{H^{s}}<\frac{\epsilon}{2} for any given ϵ>0\epsilon>0 and ss.

Next, we consider the perturbation

ϑ0=Γ12φ0:=k=k0akγ(λ0,k,λ0,k)12λ0,ksΓ12φ~0,k,\begin{split}{\vartheta_{0}}={\Gamma^{-\frac{1}{2}}\varphi^{\star}_{0}}:=\sum_{k=k_{0}}^{\infty}a_{k}\gamma(\lambda_{0,k},\lambda_{0,k})^{\frac{1}{2}}\lambda_{0,k}^{-s}\Gamma^{-\frac{1}{2}}{\widetilde{\varphi}_{0,k}},\end{split}

where

φ~0,k:=φ~(λ0,k)(t=0,x2yk),{\widetilde{\varphi}_{0,k}:=\widetilde{\varphi}^{(\lambda_{0,k})}(t=0,x_{2}-y_{k}),}

and λ0,kΛ0\lambda_{0,k}\geq\Lambda_{0} is a strictly increasing sequence to be determined below. Observe that (using a simple rescaling in time) φ~(λ0,k)(akt,x2yk)\widetilde{\varphi}^{(\lambda_{0,k})}(a_{k}t,x_{2}-y_{k}) is simply the wave packet adapted to the rescaled and translated profile fkf_{k} with frequency λ0,k\lambda_{0,k}. By taking larger k0k_{0} if necessary, we can guarantee that ϑ0Hs<ϵ2\|{\vartheta_{0}}\|_{H^{s}}<\frac{\epsilon}{2}.

We then take the initial data

θ0=θ̊0+ϑ0\begin{split}\theta_{0}=\mathring{\theta}_{0}+\vartheta_{0}\end{split}

for (1.1), which satisfies θ0Hs<ϵ\|{\theta_{0}}\|_{H^{s}}<\epsilon. Towards a contradiction, we assume that there exist δ>0\delta>0 and a solution θL([0,δ];Hs)\theta\in L^{\infty}([0,\delta];H^{s}) to (1.1) with θ(t=0)=θ0\theta(t=0)=\theta_{0}. We denote

A=supt[0,δ]θ(t)Hs\begin{split}A=\sup_{t\in[0,\delta]}\|{\theta(t)}\|_{H^{s}}\end{split}

and define on t[0,δ]t\in[0,\delta]

ϑ(t):=θ(t)θ̊,φ(t):=Γ12ϑ(t).\begin{split}\vartheta(t):=\theta(t)-\mathring{\theta},\qquad\varphi^{\star}(t):=\Gamma^{\frac{1}{2}}\vartheta(t).\end{split}

In what follows, we shall often suppress the dependence of implicit constants on Γ\Gamma, f0f_{0} and ss. We remark that, logically, the sequence {λ0,k}\{\lambda_{0,k}\} shall be fixed first, and then shall {y0,k}\{y_{0,k}\} be fixed – the last part crucially uses the unboundedness of Ω=𝕋×\Omega=\mathbb{T}\times\mathbb{R}.

2. Localization of the energy identity. We first introduce some cutoff functions. Let χ0\chi\geq 0 be a smooth function supported on [1,1][-1,1] and satisfies χ=1\chi=1 on [12,12][-\frac{1}{2},\frac{1}{2}]. Assuming that yk1y_{k-1} is given for some k2k\geq 2, we take yk8yk1y_{k}\geq 8y_{k-1} and χk(x2):=χ(2yk1(x2yk))\chi_{k}(x_{2}):=\chi(2y_{k}^{-1}(x_{2}-y_{k})). It is then guaranteed that the support of χk\chi_{k} is disjoint from each other.

Recall that the equation for φ\varphi^{\star} is given by

θ̊φ+Γ12(Γ12φΓ12φ)=0.\begin{split}\mathcal{L}_{\mathring{\theta}}\varphi^{\star}+\Gamma^{\frac{1}{2}}(\nabla^{\perp}\Gamma^{\frac{1}{2}}\varphi^{\star}\cdot\nabla\Gamma^{-\frac{1}{2}}\varphi^{\star})=0.\end{split}

As in Section 7.3, testing the equation against φ\varphi^{\star}, we immediately obtain

ddtφL22φL22\frac{\mathrm{d}}{\mathrm{d}t}\|{\varphi^{\star}}\|_{L^{2}}^{2}\lesssim\|{\varphi^{\star}}\|_{L^{2}}^{2}

which implies, by Gronwall’s inequality,

φL2φ0L2exp(Ct).\|{\varphi^{\star}}\|_{L^{2}}\lesssim\|{\varphi^{\star}_{0}}\|_{L^{2}}\exp(Ct). (7.10)

Moreover, multiplying the equation by χk\chi_{k} and testing against χkφ\chi_{k}\varphi^{\star}, we have from

χkΓ12θ̊Γ12φ+χkΓ12Γθ̊Γ12φ,χkφ=Γ12θ̊Γ12(χkφ)+Γ12Γθ̊Γ12(χkφ),χkφ+Γ12θ̊[χk,Γ12]φ+Γ12Γθ̊[χk,Γ12]φ,χkφ+[χk,Γ12]θ̊Γ12φ+[χk,Γ12]Γθ̊Γ12φ,χkφ,\begin{split}&\langle{-\chi_{k}\Gamma^{\frac{1}{2}}\nabla^{\perp}\mathring{\theta}\cdot\nabla\Gamma^{\frac{1}{2}}\varphi^{\star}+\chi_{k}\Gamma^{\frac{1}{2}}\nabla^{\perp}\Gamma\mathring{\theta}\cdot\nabla\Gamma^{-\frac{1}{2}}\varphi^{\star},\chi_{k}\varphi^{\star}}\rangle\\ &\qquad=\langle{-\Gamma^{\frac{1}{2}}\nabla^{\perp}\mathring{\theta}\cdot\nabla\Gamma^{\frac{1}{2}}(\chi_{k}\varphi^{\star})+\Gamma^{\frac{1}{2}}\nabla^{\perp}\Gamma\mathring{\theta}\cdot\nabla\Gamma^{-\frac{1}{2}}(\chi_{k}\varphi^{\star}),\chi_{k}\varphi^{\star}}\rangle\\ &\qquad\quad+\langle{-\Gamma^{\frac{1}{2}}\nabla^{\perp}\mathring{\theta}\cdot[\chi_{k},\nabla\Gamma^{\frac{1}{2}}]\varphi^{\star}+\Gamma^{\frac{1}{2}}\nabla^{\perp}\Gamma\mathring{\theta}\cdot[\chi_{k},\nabla\Gamma^{-\frac{1}{2}}]\varphi^{\star},\chi_{k}\varphi^{\star}}\rangle\\ &\qquad\quad+\langle{-[\chi_{k},\Gamma^{\frac{1}{2}}]\nabla^{\perp}\mathring{\theta}\cdot\nabla\Gamma^{\frac{1}{2}}\varphi^{\star}+[\chi_{k},\Gamma^{\frac{1}{2}}]\nabla^{\perp}\Gamma\mathring{\theta}\cdot\nabla\Gamma^{-\frac{1}{2}}\varphi^{\star},\chi_{k}\varphi^{\star}}\rangle,\end{split}

and

χkΓ12(Γ12φΓ12φ),χkφ=[χk,Γ12](Γ12φΓ12φ),χkφ+[χk,Γ12]φΓ12φ,Γ12(χkφ)\begin{split}\langle{\chi_{k}\Gamma^{\frac{1}{2}}(\nabla^{\perp}\Gamma^{\frac{1}{2}}\varphi^{\star}\cdot\nabla\Gamma^{-\frac{1}{2}}\varphi^{\star}),\chi_{k}\varphi^{\star}}\rangle&=\langle{[\chi_{k},\Gamma^{\frac{1}{2}}](\nabla^{\perp}\Gamma^{\frac{1}{2}}\varphi^{\star}\cdot\nabla\Gamma^{-\frac{1}{2}}\varphi^{\star}),\chi_{k}\varphi^{\star}}\rangle\\ &\qquad+\langle{[\chi_{k},\nabla^{\perp}\Gamma^{\frac{1}{2}}]\varphi^{\star}\cdot\nabla\Gamma^{-\frac{1}{2}}\varphi^{\star},\Gamma^{\frac{1}{2}}(\chi_{k}\varphi^{\star})}\rangle\end{split}

that

ddtχkφL22θ̊,ΓχkφL22+AχkLχkφL2.\begin{split}\frac{\mathrm{d}}{\mathrm{d}t}\|{\chi_{k}\varphi^{\star}}\|_{L^{2}}^{2}\lesssim_{\mathring{\theta},\Gamma}\|{\chi_{k}\varphi^{\star}}\|_{L^{2}}^{2}+{A}\|{\chi_{k}^{\prime}}\|_{L^{\infty}}\|{\chi_{k}\varphi^{\star}}\|_{L^{2}}.\end{split}

Here, it is important that the implicit constant depends only on θ̊\mathring{\theta} and Γ\Gamma and is independent of kk. We remark that we have used the commutator bounds [g,Γ12]hHss,ΓgLhHs+β01\|{[g,\Gamma^{\frac{1}{2}}]h}\|_{H^{s^{\prime}}}\lesssim_{s^{\prime},\Gamma}\|{g^{\prime}}\|_{L^{\infty}}\|{h}\|_{H^{s^{\prime}+\beta_{0}-1}} and [g,Γ12]hHss,ΓgLhHs\|{[g,\Gamma^{-\frac{1}{2}}]h}\|_{H^{s^{\prime}}}\lesssim_{s^{\prime},\Gamma}\|{g^{\prime}}\|_{L^{\infty}}\|{h}\|_{H^{s^{\prime}}}, which may be established using the assumptions on Γ\Gamma, Littlewood–Paley decomposition and writing out the commutator using the integral kernel of Γ±12Pν\Gamma^{\pm\frac{1}{2}}P_{\nu}; we omit the standard details. Regarding θ̊\mathring{\theta}, this estimate requires Γθ̊L<\|{\Gamma\nabla\mathring{\theta}}\|_{L^{\infty}}<\infty, and here it suffices to have that s>β0+2s>\beta_{0}+2.

Requiring that yky_{k} satisfies

yk1akγ(λ0,k,λ0,k)12λ0,ksy_{k}^{-1}\lesssim a_{k}\gamma(\lambda_{0,k},\lambda_{0,k})^{\frac{1}{2}}\lambda_{0,k}^{-s}

for suitably chosen implicit constant independent of kk, we may ensure that

χkLχkφ0L2.\|{\chi_{k}^{\prime}}\|_{L^{\infty}}\leq\|{\chi_{k}\varphi^{\star}_{0}}\|_{L^{2}}.

By Gronwall’s inequality, we conclude the following localized energy estimate:

supt[0,δ]χkφ(t)L2χkφ0L2exp(C(1+A)t).\begin{split}\sup_{t\in[0,\delta]}\|{\chi_{k}\varphi^{\star}(t)}\|_{L^{2}}\lesssim\|{\chi_{k}\varphi^{\star}_{0}}\|_{L^{2}}{\exp(C(1+A)t)}.\end{split} (7.11)

3. Localization of the generalized energy identity. We denote

φ~k(t,x2):=φ~(λ0,k)(akt,x2yk).\begin{split}\widetilde{\varphi}_{k}(t,x_{2}):=\widetilde{\varphi}^{(\lambda_{0,k})}(a_{k}t,x_{2}-y_{k}).\end{split}

From the properties of φ~(λ0)\widetilde{\varphi}^{(\lambda_{0})} summarized in the above, we have that with

Mk:=M(λ0,k),τk:=τ(λ0,k)\begin{split}M_{k}:=M(\lambda_{0,k}),\qquad\tau_{k}:=\tau(\lambda_{0,k})\end{split}
  • for t[0,ak1τk]t\in[0,a_{k}^{-1}\tau_{k}], φ~k(t)\widetilde{\varphi}_{k}(t) is supported on [yk,yk+1][y_{k},y_{k}+1] and satisfies φ~(t)Lt([0,ak1τk];L2)C\|{\widetilde{\varphi}(t)}\|_{L^{\infty}_{t}([0,a_{k}^{-1}\tau_{k}];L^{2})}\leq C;

  • there is a decomposition φ~k=φ~maink+φ~smallk\widetilde{\varphi}_{k}=\widetilde{\varphi}^{main}_{k}+\widetilde{\varphi}^{small}_{k} such that for t[0,ak1τk]t\in[0,a_{k}^{-1}\tau_{k}],

    φ~maink(t)HsCsλks(akt),φ~smallk(t)L2Cλkc0(akt);\begin{split}\|{\widetilde{\varphi}^{main}_{k}(t)}\|_{H^{-s}}\leq C_{-s}\lambda_{k}^{-s}(a_{k}t),\qquad\|{\widetilde{\varphi}^{small}_{k}(t)}\|_{L^{2}}\leq C\lambda_{k}^{-c_{0}}(a_{k}t);\end{split}
  • we have θ̊φ~kL1([0,ak1τk];L2)Cλ0,kc0\|{\mathcal{L}_{\mathring{\theta}}\widetilde{\varphi}_{k}}\|_{L^{1}([0,a_{k}^{-1}\tau_{k}];L^{2})}\leq C\lambda_{0,k}^{-c_{0}} and θ̊φ~kL1([0,ak1τk];L2[,+1])Cmin{1,2}λ0,kc0\|{\mathcal{L}_{\mathring{\theta}}\widetilde{\varphi}_{k}}\|_{L^{1}([0,a_{k}^{-1}\tau_{k}];L^{2}[\ell,\ell+1])}\leq C\min\{1,\ell^{-2}\}\lambda_{0,k}^{-c_{0}}.

Here, λk(t)\lambda_{k}(t) is the solution to (4.16) with λ0=λ0,k\lambda_{0}=\lambda_{0,k} and f=θ̊f=\mathring{\theta}. The constants C,CsC,C_{-s} do not depend on kk; in particular, the L1tL2L^{1}_{t}L^{2} estimates follow directly from scaling property of θ̊\mathcal{L}_{\mathring{\theta}} (2.1). In particular, on t[0,min{δ,ak1τk}]t\in[0,\min\{\delta,a_{k}^{-1}\tau_{k}\}], we have χk2φ~k(t)=χkφ~k(t)=φ~k(t)\chi_{k}^{2}\widetilde{\varphi}_{k}(t)=\chi_{k}\widetilde{\varphi}_{k}(t)=\widetilde{\varphi}_{k}(t). Before we proceed, note

τkMkγ(λ0,k,λ0,k)1γδ0(λ0,k,λ0,k)\begin{split}\tau_{k}{\leq\frac{M_{k}}{\gamma(\lambda_{0,k},\lambda_{0,k})}\leq}\frac{1}{\gamma^{\delta_{0}}(\lambda_{0,k},\lambda_{0,k})}\end{split}

and therefore by taking λ0,k\lambda_{0,k} sufficiently large, we may ensure that ak1τkδa_{k}^{-1}\tau_{k}\leq\delta. Choosing λ0,k\lambda_{0,k} larger if necessary, we may also ensure that akMksβ0a_{k}M_{k}^{s-\beta_{0}}\to\infty as kk\to\infty.

Denoting ϵk:=θ̊φ~k\boldsymbol{\epsilon}_{k}:=\mathcal{L}_{\mathring{\theta}}\widetilde{\varphi}_{k}, we deduce from applying (2.4) with φ=φ\varphi=\varphi^{\star} and ψ=φ~k\psi=\widetilde{\varphi}_{k} that

ddtχkφ,φ~k=φ,ϵk+φ,([Γ12,Γθ̊]Γ12+Γ12[Γ12,Γθ̊])(χkφ~k)Γ12(Γ12φΓ12φ),χk2φ~k.\begin{split}\frac{\mathrm{d}}{\mathrm{d}t}\langle{\chi_{k}\varphi^{\star},\widetilde{\varphi}_{k}}\rangle&=\langle{\varphi^{\star},\boldsymbol{\epsilon}_{k}}\rangle+\langle{\varphi^{\star},\left([\Gamma^{\frac{1}{2}},\nabla^{\perp}\Gamma\mathring{\theta}\cdot\nabla]\Gamma^{-\frac{1}{2}}+\Gamma^{-\frac{1}{2}}[\Gamma^{\frac{1}{2}},\nabla^{\perp}\Gamma\mathring{\theta}\cdot\nabla]\right)(\chi_{k}\widetilde{\varphi}_{k})}\rangle\\ &\qquad-\langle{\Gamma^{\frac{1}{2}}(\nabla^{\perp}\Gamma^{\frac{1}{2}}\varphi^{\star}\cdot\nabla\Gamma^{-\frac{1}{2}}\varphi^{\star}),\chi_{k}^{2}\widetilde{\varphi}_{k}}\rangle.\end{split} (7.12)

We write

φ,ϵk=χkφ,ϵk+φ,(1χk)ϵk\begin{split}\langle{\varphi^{\star},\boldsymbol{\epsilon}_{k}}\rangle=\langle{\chi_{k}\varphi^{\star},\boldsymbol{\epsilon}_{k}}\rangle+\langle{\varphi^{\star},(1-\chi_{k})\boldsymbol{\epsilon}_{k}}\rangle\end{split}

and apply Cauchy–Schwartz. Furthermore, modulo several commutators involving χk\chi_{k}, the other two terms on the right hand side of (7.12) can be written as

χkφ,([Γ12,Γθ̊]Γ12+Γ12[Γ12,Γθ̊])φ~kΓ12(Γ12(χkφ)Γ12(χkφ)),φ~k.\begin{split}\langle{\chi_{k}\varphi^{\star},\left([\Gamma^{\frac{1}{2}},\nabla^{\perp}\Gamma\mathring{\theta}\cdot\nabla]\Gamma^{-\frac{1}{2}}+\Gamma^{-\frac{1}{2}}[\Gamma^{\frac{1}{2}},\nabla^{\perp}\Gamma\mathring{\theta}\cdot\nabla]\right)\widetilde{\varphi}_{k}}\rangle-\langle{\Gamma^{\frac{1}{2}}(\nabla^{\perp}\Gamma^{\frac{1}{2}}(\chi_{k}\varphi^{\star})\cdot\nabla\Gamma^{-\frac{1}{2}}(\chi_{k}\varphi^{\star})),\widetilde{\varphi}_{k}}\rangle.\end{split}

Now, it is important that we have

Γ12(Γ12(χkφ)Γ12(χkφ))L2(1+A)χkφL2.\begin{split}\|{\Gamma^{\frac{1}{2}}(\nabla^{\perp}\Gamma^{\frac{1}{2}}(\chi_{k}\varphi^{\star})\cdot\nabla\Gamma^{-\frac{1}{2}}(\chi_{k}\varphi^{\star}))}\|_{L^{2}}\lesssim(1+A)\|{\chi_{k}\varphi^{\star}}\|_{L^{2}}.\end{split}

Observe that the terms involving commutators can be bounded in absolute value by

(1+A2)χkLφ~kL2(1+A2)χkφ0L2φ~kL2.\begin{split}\lesssim(1+A^{2})\|{\chi_{k}^{\prime}}\|_{L^{\infty}}\|{\widetilde{\varphi}_{k}}\|_{L^{2}}\lesssim(1+A^{2})\|{\chi_{k}\varphi^{\star}_{0}}\|_{L^{2}}\|{\widetilde{\varphi}_{k}}\|_{L^{2}}.\end{split}

Therefore, we have arrived at the following localized and generalized energy estimate:

|ddtχkφ,φ~k|(1+A2)(φ~kL2+ϵkL2)(χkφL2+χkφ0L2)+φL2(1χk)ϵkL2.\begin{split}\left|\frac{\mathrm{d}}{\mathrm{d}t}\langle{\chi_{k}\varphi^{\star},\widetilde{\varphi}_{k}}\rangle\right|&\lesssim(1+A^{2})(\|{\widetilde{\varphi}_{k}}\|_{L^{2}}+\|{\boldsymbol{\epsilon}_{k}}\|_{L^{2}})(\|{\chi_{k}\varphi^{\star}}\|_{L^{2}}+\|{\chi_{k}\varphi^{\star}_{0}}\|_{L^{2}}){+\|{\varphi^{\star}}\|_{L^{2}}\|{(1-\chi_{k})\boldsymbol{\epsilon}_{k}}\|_{L^{2}}.}\end{split} (7.13)

4. Conclusion. Applying (7.10), (7.12) to (7.13) and integrating in time, we obtain that

χkφ,φ~k(t)χkφ,φ~k(0)C(1+A2)exp(C(1+A)t)χkφ0L2(tφ~0,kL2+0tϵk(τ)L2dτ)Cexp(Ct)φ0L20t(1χk)ϵk(τ)L2dτ.\begin{split}\langle{\chi_{k}\varphi^{\star},\widetilde{\varphi}_{k}}\rangle(t)&\geq\langle{\chi_{k}\varphi^{\star},\widetilde{\varphi}_{k}}\rangle(0)\\ &\mathrel{\phantom{=}}-C(1+A^{2}){\exp(C(1+A)t)}\|{\chi_{k}\varphi^{\star}_{0}}\|_{L^{2}}(t\|{\widetilde{\varphi}_{0,k}}\|_{L^{2}}+\int_{0}^{t}\|{\boldsymbol{\epsilon}_{k}(\tau)}\|_{L^{2}}\,\mathrm{d}\tau)\\ &\mathrel{\phantom{=}}{-C\exp(Ct)\|{\varphi^{\star}_{0}}\|_{L^{2}}\int_{0}^{t}\|{(1-\chi_{k})\boldsymbol{\epsilon}_{k}(\tau)}\|_{L^{2}}\,\mathrm{d}\tau.}\end{split}

By Proposition 6.1.(2) and our construction, observe that

χkφ,φ~k(0)χkφ0L2akγ(λ0,k,λ0,k)12λ0,ks.\langle{\chi_{k}\varphi^{\star},\widetilde{\varphi}_{k}}\rangle(0)\simeq\|{\chi_{k}\varphi^{\star}_{0}}\|_{L^{2}}\lesssim a_{k}\gamma(\lambda_{0,k},\lambda_{0,k})^{\frac{1}{2}}\lambda_{0,k}^{-s}.

We fix t=ak1τkt=a_{k}^{-1}\tau_{k}. Using the localized error bound, if we take yky_{k} large enough depending on λ0,k\lambda_{0,k}, we may ensure that

φ0L20t(1χk)ϵk(τ)L2dτλ0,kc1χkφ,φ~k(0).\|{\varphi^{\star}_{0}}\|_{L^{2}}\int_{0}^{t}\|{(1-\chi_{k})\boldsymbol{\epsilon}_{k}(\tau)}\|_{L^{2}}\,\mathrm{d}\tau\leq\lambda_{0,k}^{-c_{1}}\langle{\chi_{k}\varphi^{\star},\widetilde{\varphi}_{k}}\rangle(0).

For kk large enough so that τk(1+A2)1\tau_{k}\ll(1+A^{2})^{-1} and λ0,kc1(1+A2)1\lambda_{0,k}^{-c_{1}}\ll(1+A^{2})^{-1}, we obtain

χkφ,φ~k(ak1τk)12χkφ,φ~k(0).\begin{split}\langle{\chi_{k}\varphi^{\star},\widetilde{\varphi}_{k}}\rangle(a_{k}^{-1}\tau_{k})\geq\frac{1}{2}\langle{\chi_{k}\varphi^{\star},\widetilde{\varphi}_{k}}\rangle(0).\end{split}

On the other hand, we may bound

χkφ,φ~k(ak1τk)Γ12χkφ(ak1τk)HsΓ12φ~maink(ak1τk)Hs+χkφ(ak1τk)L2φ~smallk(ak1τk)L2.\begin{split}\langle{\chi_{k}\varphi^{\star},\widetilde{\varphi}_{k}}\rangle(a_{k}^{-1}\tau_{k})\leq\|{\Gamma^{-\frac{1}{2}}\chi_{k}\varphi^{\star}(a_{k}^{-1}\tau_{k})}\|_{H^{s}}\|{\Gamma^{\frac{1}{2}}\widetilde{\varphi}^{main}_{k}(a_{k}^{-1}\tau_{k})}\|_{H^{-{s}}}+\|{\chi_{k}\varphi^{\star}(a_{k}^{-1}\tau_{k})}\|_{L^{2}}\|{\widetilde{\varphi}^{small}_{k}(a_{k}^{-1}\tau_{k})}\|_{L^{2}}.\end{split}

By taking kk larger if necessary, we may guarantee that

χkφ(ak1τk)L2φ~smallk(ak1τk)L2<14χkφ,φ~k(0).\begin{split}\|{\chi_{k}\varphi^{\star}(a_{k}^{-1}\tau_{k})}\|_{L^{2}}\|{\widetilde{\varphi}^{small}_{k}(a_{k}^{-1}\tau_{k})}\|_{L^{2}}<\frac{1}{4}\langle{\chi_{k}\varphi^{\star},\widetilde{\varphi}_{k}}\rangle(0).\end{split}

This gives

Γ12χkφ(ak1τk)Hsχkφ0L2φ~k,0L2Γ12φ~kmain(ak1τk)Hsakγ(λ0,k)12λ0,ksλsk(τk)γ(λk(τk))12akγ(λ0,k)12γ(λk(τk))12MksakMksβ0.\begin{split}\|{\Gamma^{-\frac{1}{2}}\chi_{k}\varphi^{\star}(a_{k}^{-1}\tau_{k})}\|_{H^{s}}&\gtrsim\frac{\|{\chi_{k}\varphi^{\star}_{0}}\|_{L^{2}}\|{\widetilde{\varphi}_{k,0}}\|_{L^{2}}}{\|{\Gamma^{\frac{1}{2}}\widetilde{\varphi}_{k}^{main}(a_{k}^{-1}\tau_{k})}\|_{H^{-{s}}}}\gtrsim{a_{k}}\frac{\gamma(\lambda_{0,k})^{\frac{1}{2}}}{\lambda_{0,k}^{s}}\frac{\lambda^{s}_{k}(\tau_{k})}{\gamma(\lambda_{k}(\tau_{k}))^{\frac{1}{2}}}\\ &\gtrsim{a_{k}}\frac{\gamma(\lambda_{0,k})^{\frac{1}{2}}}{\gamma(\lambda_{k}(\tau_{k}))^{\frac{1}{2}}}M_{k}^{s}\gtrsim{a_{k}}M_{k}^{s-\beta_{0}}.\end{split}

For kk large, [Γ12,χk]φHs(1+A)χkL1\|{[\Gamma^{-\frac{1}{2}},\chi_{k}]\varphi^{\star}}\|_{H^{s}}\lesssim(1+A)\|{\chi_{k}^{\prime}}\|_{L^{\infty}}\lesssim 1 and therefore we obtain

χk(θ(ak1τk)θ̊)Hs=χkΓ12φ(ak1τk)HsakMksβ0.\begin{split}\|{\chi_{k}(\theta(a_{k}^{-1}\tau_{k})-\mathring{\theta})}\|_{H^{s}}=\|{\chi_{k}\Gamma^{-\frac{1}{2}}\varphi^{\star}(a_{k}^{-1}\tau_{k})}\|_{H^{s}}\gtrsim{a_{k}}M_{k}^{s-\beta_{0}}.\end{split}

On the other hand, with a constant independent of kk,

χk(θ(ak1τk)θ̊)Hs(θ(ak1τk)θ̊)Hs1+A.\begin{split}\|{\chi_{k}(\theta(a_{k}^{-1}\tau_{k})-\mathring{\theta})}\|_{H^{s}}\lesssim\|{(\theta(a_{k}^{-1}\tau_{k})-\mathring{\theta})}\|_{H^{s}}\lesssim 1+A.\end{split}

We obtain a contradiction as kk\to\infty since akMksβ0{a_{k}M_{k}^{s-\beta_{0}}\to\infty}. This finishes the proof. ∎

References

  • [1] David M. Ambrose, Gideon Simpson, J. Douglas Wright, and Dennis G. Yang, Ill-posedness of degenerate dispersive equations, Nonlinearity 25 (2012), no. 9, 2655–2680. MR 2967120
  • [2] F. Betancourt, R. Bürger, K. H. Karlsen, and E. M. Tory, On nonlocal conservation laws modelling sedimentation, Nonlinearity 24 (2011), no. 3, 855–885. MR 2772627
  • [3] J. L. Bona, M. Chen, and J.-C. Saut, Boussinesq equations and other systems for small-amplitude long waves in nonlinear dispersive media. I. Derivation and linear theory, J. Nonlinear Sci. 12 (2002), no. 4, 283–318. MR 1915939
  • [4] B. H. Burgess, Thin-jet scaling in large-scale shallow water quasigeostrophic flow, Geophys. Astrophys. Fluid Dyn. 114 (2020), no. 4-5, 481–503. MR 4150285
  • [5] B. H. Burgess and D. G. Dritschel, Potential vorticity fronts and the late-time evolution of large-scale quasi-geostrophic flows, J. Fluid Mech. 939 (2022), Paper No. A40, 16. MR 4402682
  • [6] R. Caflisch and G. C. Papanicolaou, Dynamic theory of suspensions with Brownian effects, SIAM J. Appl. Math. 43 (1983), no. 4, 885–906. MR 709743
  • [7] Roberto Camassa and Darryl D. Holm, An integrable shallow water equation with peaked solitons, Phys. Rev. Lett. 71 (1993), no. 11, 1661–1664. MR 1234453
  • [8] Dongho Chae, Peter Constantin, Diego Córdoba, Francisco Gancedo, and Jiahong Wu, Generalized surface quasi-geostrophic equations with singular velocities, Comm. Pure Appl. Math. 65 (2012), no. 8, 1037–1066. MR 2928091
  • [9] Dongho Chae, Pierre Degond, and Jian-Guo Liu, Well-posedness for Hall-magnetohydrodynamics, Ann. Inst. H. Poincaré Anal. Non Linéaire 31 (2014), no. 3, 555–565. MR 3208454
  • [10] Dongho Chae, In-Jee Jeong, Jungkyoung Na, and Sung-Jin Oh, Well-posedness for Ohkitani model and long-time existence for surface quasi-geostrophic equations.
  • [11] Dongho Chae, Renhui Wan, and Jiahong Wu, Local well-posedness for the Hall-MHD equations with fractional magnetic diffusion, J. Math. Fluid Mech. 17 (2015), no. 4, 627–638. MR 3412271
  • [12] Peter Constantin, Andrew J. Majda, and Esteban Tabak, Formation of strong fronts in the 22-D quasigeostrophic thermal active scalar, Nonlinearity 7 (1994), no. 6, 1495–1533. MR 1304437
  • [13] Peter Constantin, Andrew J. Majda, and Esteban G. Tabak, Singular front formation in a model for quasigeostrophic flow, Phys. Fluids 6 (1994), no. 1, 9–11. MR 1252829
  • [14] Walter Craig and Jonathan Goodman, Linear dispersive equations of Airy type, J. Differential Equations 87 (1990), no. 1, 38–61. MR 1070026
  • [15] 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 (1995), no. 2, 248–252. MR 1349459
  • [16] Pierre Germain, Benjamin Harrop-Griffiths, and Jeremy L. Marzuola, Existence and uniqueness of solutions for a quasilinear KdV equation with degenerate dispersion, Comm. Pure Appl. Math. 72 (2019), no. 11, 2449–2484. MR 4011864
  • [17] A. E. Green, N. Laws, and P. M. Naghdi, On the theory of water waves, Proc. Roy. Soc. London Ser. A 338 (1974), 43–55. MR 349127
  • [18] A. Hasegawa and K. Mima, Pseudo-three-dimensional turbulence in a magnetized nonuniform plasma, Phys. Fluids 21 (1978).
  • [19] M. S. Ismail and T. R. Taha, A numerical study of compactons, Math. Comput. Simulation 47 (1998), no. 6, 519–530. MR 1662379
  • [20] In-Jee Jeong and Sung-Jin Oh, On the Cauchy problem for the Hall and electron magnetohydrodynamic equations without resistivity I: illposedness near degenerate stationary solutions, arXiv:1902.02025.
  • [21] In-Jee Jeong and Sung-Jin Oh, Strong illposedness for quasilinear dispersive equations via degenerate dispersion.
  • [22] Vitaly D. Larichev and James C. McWilliams, Weakly decaying turbulence in an equivalent‐barotropic fluid, Physics of Fluids A: Fluid Dynamics 3 (1991), no. 5, 938–950.
  • [23] Yuri E. Litvinenko and Liam C. McMahon, Finite-time singularity formation at a magnetic neutral line in Hall magnetohydrodynamics, Appl. Math. Lett. 45 (2015), 76–80. MR 3316965
  • [24] Jeremy L. Marzuola, Jason Metcalfe, and Daniel Tataru, Quasilinear Schrödinger equations I: Small data and quadratic interactions, Adv. Math. 231 (2012), no. 2, 1151–1172. MR 2955206
  • [25] Sigeru Mizohata, Some remarks on the Cauchy problem, J. Math. Kyoto Univ. 1 (1961/62), 109–127. MR 170112
  • [26] Koji Ohkitani, Growth rate analysis of scalar gradients in generalized surface quasigeostrophic equations of ideal fluids, Phys. Rev. E (3) 83 (2011), no. 3, 036317, 8. MR 2788267
  • [27]  , Asymptotics and numerics of a family of two-dimensional generalized surface quasi-geostrophic equations, Physics of Fluids 24 (2012), no. 095101.
  • [28] Hans Pécseli, Waves and oscillations in plasmas, Series in Plasma Physics (2012).
  • [29] J. Pedlosky, Geophysical fluid dynamics, Springer, 1987.
  • [30] Philip Rosenau, On solitons, compactons, and Lagrange maps, Phys. Lett. A 211 (1996), no. 5, 265–275. MR 1377202
  • [31] Philip Rosenau and James M. Hyman, Compactons: Solitons with finite wavelength, Phys. Rev. Lett. 70 (1993), 564–567.
  • [32] Jacob Rubinstein, Evolution equations for stratified dilute suspensions, Phys. Fluids A 2 (1990), no. 1, 3–6. MR 1030168
  • [33] Jacob Rubinstein and Joseph B. Keller, Sedimentation of a dilute suspension, Phys. Fluids A 1 (1989), no. 4, 637–643. MR 1021642
  • [34] Melvin E. Stern and Lawrence J. Pratt, Dynamics of vorticity fronts, J. Fluid Mech. 161 (1985), 513–532. MR 828157
  • [35] Michael E. Taylor, Pseudodifferential operators, Princeton Mathematical Series, vol. 34, Princeton University Press, Princeton, N.J., 1981. MR 618463
  • [36] Chuong V. Tran and David G. Dritschel, Impeded inverse energy transfer in the Charney-Hasegawa-Mima model of quasi-geostrophic flows, J. Fluid Mech. 551 (2006), 435–443. MR 2263703
  • [37] Kevin Zumbrun, On a nonlocal dispersive equation modeling particle suspensions, Quart. Appl. Math. 57 (1999), no. 3, 573–600. MR 1704419